VehicleDynamics Compendium 2020
VehicleDynamics Compendium 2020
COMPENDIUM
unit
rear
coupling on
front unit
coupling on
unit
Control algorithm:
Version 2020 for course academic year autumn 2020 and spring 2021
Latest draft available at https://tinyurl.com/VehDynCompDraft
Bengt Jacobson et al
Vehicle Dynamics Group, Division Vehicle and Autonomous Systems,
Department of Mechanics and Maritime Sciences, Chalmers University of Technology, www.chalmers.se
2
Preface 2020
From and including 2020, Chalmers does not print the compendium on paper. Therefore, the two vari-
ants, “Printed on paper” and “Digital only” introduced 2019, are not equally motivated. So, the material
marked with “#Digital only” from 2019 is kept, and still with blue text, but now marked “§ ”. The
reader should understand the extra material as possible to jump over for a basic understanding but
recommended if aiming for more advanced knowledge.
Many contribute to this compendium, and new for version 2020 is that contributors are mentioned in
the beginning of the section where their contribution was primarily done:
Jelena Andric and Majid Astaneh, Chalmers
Pinar Boyraz, Vehicle Safety at Chalmers
Adam Brandt, Vehicle AeroDynamics at Chalmers and CEVT
Fredrik Bruzelius, Vehicle Dynamics at Chalmers and VTI
Niklas Fröjd, Volvo Trucks
Toheed Ghandriz, Vehicle Dynamics at Chalmers
Anders Hedman, Volvo Trucks
Inge Johansson, Volvo Trucks
Ingemar Johansson, CEVT and Vehicle Engineering at Chalmers
Mats Jonasson, Volvo Cars and Vehicle Dynamics at Chalmers
Mikko Karisaari, Oulo University, Finland
Waltteri Koskinen, student at Tampere University, Finland
Leo Laine, Volvo Trucks and Chalmers
Mathias Lidberg, Vehicle Dynamics at Chalmers
Luigi Romano, Vehicle Dynamics at Chalmers
Dragan Sekulić, Vehicle Dynamics at Chalmers
Alexey Vdodin, Vehicle AeroDynamics at Chalmers
My apologies to contributors I have forgotten. Remind me, and I’ll add you to next version.
/Bengt Jacobson, Göteborg, Last Modification: 2020-10-31 10:19
Preface 2019
The compendium 2019 is published in 2 variants: “Printed on paper” and “Digital only”. Both are available as pdf-file. The
“Digital only” variant contains some additional material; search for “#DigitalOnly”. The numbered items (figures, equations,
etc) in the “Digital only” variant does not have numbering, in order to keep same numbers of each item between the variants.
However, note that page numbering varies between the variants.
Thanks to Tobias Brandin, Fredrik Bruzelius, Edo Drenth, Niklas Fröjd, Toheed Ghandriz, Patrick Gruber, Mats Jonasson, Ma-
thias Lidberg, Anders Lindström, Oscar Ljungcrantz, Peter Nilsson, Luigi Romano, Juliette Utbult and errata reporting from
students. Sorry, if I forgot some contributor! /Bengt Jacobson, Göteborg, October 2019
Preface 2018
One large rearrangement is done: Subsystem descriptions has been collected from Chapter 3, 4 and 5 to 2.3-2.7. Also, minor
changes and additions has been done throughout all chapters. Many thanks to, among other, Niklas Fröjd Volvo GTT, Toheed
and Fatemeh Ghandriz, Ingemar Johansson CEVT, Mats Jonasson VCC, Mathias Lidberg, Simone Sebben, Alexey Vdovin.
Thanks also to many students that have found and reported errors in the previous edition.
/Bengt Jacobson, Göteborg, October 2018
Preface 2017
This edition has various smaller changes and additions. Thanks to Fredrik Bruzelius (VTI), Tobias Brandin (VCC), Niklas Fröjd
(Volvo GTT), Assar Jarlsson (Kinnarps), Pär Pettersson (Chalmers), among others. Thanks also to many students that have
found and reported errors in the previous edition. /Bengt Jacobson, Göteborg, October 2017
Preface 2016
This edition has various changes and additions. Some of these are: Chapter 1: Control engineering, Chapter 2: Tyre models,
Driver models, Chapter 3: Propulsion systems, Varying road pitch, Non-reactive truck suspensions, Chapter 4: Track-ability,
Articulated vehicles, and Cambering vehicles.
Thanks to Cornelia Lex (TU Graz), Fredrik Bruzelius (VTI), Niklas Fröjd, Anders Hedman, Kristoffer Tagesson, Peter Nilsson,
Sixten Berglund (Volvo GTT), Tobias Brandin, Edo Drenth, Mats Jonasson (VCC), Mathias Lidberg, Artem Kusachov, Anton
Albinsson, Manjurul Islam, Pär Pettersson, Ola Benderius (Chalmers), Mats Sabelström, and Roland Svensson among others.
3
/Bengt Jacobson, Göteborg, October 2016
Preface 2015
This edition has various changes and additions. Some of these are: brush model with parabolic pressure distribution, typical
numerical data for heavy vehicle, added “2.2.3 Tyre”, “4.5.3.2 Example of explicit form model”, more about tyre relaxation,
introduction of neutral steering point, introduction of steady state roll-over wheel lift diagram. Thanks to Anton Albinsson,
Edo Drenth (VCC), Gunnar Olsson (LeanNova), Manjurul Islam, Mathias Lidberg, Mats Jonasson (VCC), Niklas Fröjd (Volvo
GTT), Ola Benderius, Pär Pettersson, and Zuzana Nedelkova among other. /Bengt Jacobson, Göteborg, 2015
Preface 2014
This edition has various small changes and additions. The largest changes are: Function definitions added and major update
of sections 2.3, 4.1.1, 0, 6.1.1.
Thanks to Lars Almefelt from Chalmers, Jan Andersson from VCC, Kristoffer Tagesson from Volvo GTT and Gunnar Olsson
from Leannova and Karthik Venkataraman. /Bengt Jacobson, Göteborg, 2014
Preface 2013
This edition has various small changes and additions. The largest additions were in: Functional architecture, Smaller vehicles,
Roll-over, Pendulum effect in lateral load transfer and Step steer.
Thanks to Gunnar Olsson from LeanNova, Mathias Lidberg, Marco Dozza, Andrew Dawkes from Chalmers, Erik Coelingh from
Volvo Cars, Fredrik Bruzelius from VTI, Edo Drenth from Modelon, Mats Sabelström, Martin Petersson and Leo Laine from
Volvo GTT. /Bengt Jacobson, Göteborg, 2013
Preface 2012
A major revision is done. The material is renamed from “Lecture notes” to “Compendium”. Among the changes it is worth
mentioning: 1) the chapters about longitudinal, lateral and vertical are more organised around design for vehicle functions,
2) a common notation list is added, 3) brush tyre model added, 4) more organised and detailed about different load transfer
models, and 5) road spectral density roughness model is added.
Thanks to Adithya Arikere, John Aurell, Andrew Dawkes, Edo Drenth, Mathias Lidberg, Peter Nilsson, Gunnar Olsson, Mats
Sabelström, Ulrich Sander, Simone Sebben, Kristoffer Tagesson, Alexey Vdovin and Derong Yang for review reading.
/Bengt Jacobson, Göteborg, 2012
Preface 2011
Material on heavy vehicles is added with help from John Aurell. Coordinate system is changed from SAE to ISO. Minor addi-
tions and changes are also done. /Bengt Jacobson, Göteborg, 2011
Preface 2007
This document was developed as a result of the reorganization of the Automotive Engineering Master’s Programme at
Chalmers in 2007. The course content has been modified in response to the redistribution of vehicle dynamics and power
train education.
These lecture notes are based on the original documents developed by Dr Bengt Jacobson. The text and examples have been
reformatted and edited but the author is indebted to the contribution of Dr Jacobson. /Rob Thomson, Gothenburg, 2007
Keywords: Vehicle Engineering, Automotive Engineering, Vehicle Dynamics, Vehicle Motion, Modelling, Modelica, Simulation
Cover:
Left column, from top: Steering control testing in Jokkmokk, Testing A-double combination vehicle in real transport operation
in the Autofreight research project (photo: Borås Stad), and Alexander Rasch, Chalmers, tests slalom with e-scooter. Right
column: Figures from compendium additions during 2020.
4
CONTENTS
1 INTRODUCTION 9
1.1 Definition of Vehicle Dynamics 9
1.2 About this compendium 9
1.3 Automotive engineering 10
1.3.1 Vehicle Dynamics Engineers’ Industry Roles 10
1.4 Requirement Setting 11
1.4.1 Attributes 11
1.4.2 Functions 12
1.4.3 Requirements 14
1.4.4 Models, Methods and Tools 17
1.5 Engineering 18
1.5.1 Model Based Engineering 18
1.5.2 Mechanical/Machine Engineering 38
1.5.3 Control Engineering 47
1.5.4 Tools 54
1.6 Vehicle Engineering 61
1.6.1 Vehicle Motions and Coordinate Systems 61
1.6.2 Complete Vehicle Modelling Concepts 67
1.6.3 Vehicle Dynamics Terms 71
1.6.4 Vehicle Architectures 73
1.6.5 Verification Methods with Real Vehicle 77
1.6.6 Verification Methods with Virtual Vehicle 78
1.7 Heavy Trucks 79
1.7.1 General Differences 80
1.7.2 Vehicle Dynamics Differences 80
1.7.3 Definitions 80
1.8 Smaller Vehicles 81
1.9 Notation List 83
1.10 Typical Numerical Data 86
1.10.1 For Passenger Vehicles 87
1.10.2 For Heavy Vehicles 88
5
2.3.3 Suspension -- Heave and Pitch 148
2.3.4 Suspension -- Heave and Roll 148
2.4 Propulsion System 150
2.4.1 Modelling Concepts 151
2.4.2 Prime movers 152
2.4.3 Transmissions 156
2.4.4 Clutches and Brakes in Transmission 158
2.4.5 Hydrodynamic Torque Converters 163
2.4.6 Energy Storages and Energy Buffers 163
2.4.7 Special Topology Propulsion Systems 164
2.5 (Wheel) Braking System 167
2.6 (Wheel) Steering System 170
2.6.1 Chassis Steering Geometry 170
2.6.2 Steering System Forces 171
2.7 Environment Sensing System 174
2.8 Vehicle Aerodynamics 174
2.8.1 Longitudinal Relative Wind Velocity 174
2.8.2 Lateral Relative Wind Velocity 175
2.8.3 § Variation of Relative Wind 175
2.9 Driving and Transport Application 176
2.9.1 Mission, Road and Traffic 176
2.9.2 Driver 176
6
3.5.3 Longitudinal Motion Function Architecture 227
7
5.4.4 One-Mode Models 346
5.5 Functions for Stationary Oscillations 347
5.5.1 Ride Comfort * 347
5.5.2 Fatigue Life * 352
5.5.3 Road Grip * 353
5.5.4 Other Functions 353
5.6 Variation of Suspension Design 354
5.6.1 Varying Suspension Stiffness 355
5.6.2 Varying Suspension Damping 356
5.6.3 Varying Unsprung Mass 356
5.6.4 Varying Tyre Stiffness 357
5.6.5 § Varying Skyhook Damping 358
5.7 Two Dimensional Oscillations 358
5.7.1 Heave and Roll 358
5.7.2 Heave and Pitch 359
5.8 Three Dimensional Oscillations 362
5.9 Transient Vertical Dynamics 363
5.10 Control Functions 363
BIBLIOGRAPHY 365
8
Introduction
1 INTRODUCTION
1.1 Definition of Vehicle Dynamics
Vehicle Dynamics is an engineering subject about motion of vehicles in user-relevant operations. The
subject is applied, and applied on a certain group of products, i.e. vehicles. Vehicle Dynamics always
uses terms, theories and methods from Mechanical/Machine engineering, but often also from Con-
trol/Signal engineering and Human behavioural science.
9
Introduction
year1 year2
Pre-Series1 Pre-Series2 Pre- Start
(mule=old (new chassis) Series3 Of
chassis) (final Production
chassis)
Research
projects
Geometry package, Changing geometry, Changing bushings,
Strength of parts, Function content, Tuning SW,
Signal interface, … …
… within subsystems within subsystems
between subsystems (hi-&lo mu tests) (hi&lo-mu tuning)
Figure 1-1: An example of vehicle programme and Vehicle Dynamics related activities.
10
Introduction
few selected customers, pilot projects. Some claim that, for paradigm shifts like automated driving and
electromobility, smaller commercial pilots are more suitable than large vehicle programmes. The soft-
ware in the vehicles, can even be updated after sales, which can be seen as a continuous development,
leaving the concept of only develop towards a certain production start date.
Figure 1-2: V-process for a vehicle program. The more design loops are utilized, the more ”agile”.
1.4.1 Attributes
An attribute is a high-level aspect of how the users perceive the vehicle. Attributes which are espe-
cially relevant for Vehicle Dynamics are listed in Figure 1-3. The table is much generalised and the at-
tributes in it would typically need to be decomposed into more attributes when used in the engineer-
ing organisation of an OEM. Also, not mentioned in the table, are attributes which are less specific for
11
Introduction
vehicle dynamics, such as Affordability (low cost for user), Quality (functions sustained over vehicle
lifetime), Styling (appearance, mainly visual), etc.
A set of Attributes is a way to categorise or group functions, especially useful for an OEM organisation
and vehicle development programs. A set of Functions is a way to group requirements. Legal require-
ments are often, but not always, possible to trace back to primarily one specific attribute. Require-
ments arising from OEM-internal platform and architecture constraints are often more difficult to
trace in that way. Hence, “platform/architecture” is a “requirement container”, beside the attributes.
Attribute Description
This attribute means to maximize output from and minimize costs for transportation.
Transport output can be measured in 𝑝𝑒𝑟𝑠𝑜𝑛 𝑘 𝑡𝑜𝑛 𝑘 or 3 𝑘 . The costs are
Transport Efficiency
mainly energy costs and time, but also wear of vehicle parts influence. The attribute is most
important for commercial vehicles but becomes increasingly important also for passenger
vehicles. The attribute is mainly addressing long-term vehicle usage pattern, typically 10
min to 10 hours. There are diverse ways to define such usages, e.g. (Urban / Highway /
Mixed) driving cycles. So far, the attribute is mainly required and assessed by the vehicle
customers/users.
The attribute can also be seen to include “Environmental Efficiency”, which means low us-
age of natural resources (mainly energy) and low pollution, per performed transport task.
This is to a substantial extent required and assessed by society/legislation.
Minimizing risk of property damages, personal injuries and fatalities both in vehicle and
Safety
outside, while performing the transportation. This attribute is to a considerable extent re-
quired and assessed by society/legislation. In some markets, mainly developed countries, it
is also important for vehicle customers/users.
How the occupants (often the driver) experiences the vehicle during transport; from relaxed
transport (comfort) to active driving (sensation). This attribute contains sub-attributes as:
• Ride comfort. Ride comfort often refers to vibrations and harshness of the occu-
pants’ motion, primarily vertical but secondly longitudinal and lateral. So, V and H in
User Experience (Driver Experience)
1.4.2 Functions
In this compendium, a function is more specific than an attribute. A function should define measures of
something the (complete) vehicle does, so that one can set (quantitative) requirements on each meas-
ure, see 1.4.3. The function does not primarily stipulate any specific subsystem. However, the realisa-
tion of a function in a certain vehicle programme, normally only engages a limited subset of all
12
Introduction
subsystems. So, the function will there pose requirements on those subsystems. Hence, it is easy to mix
up whether a function origin from an attribute or a subsystem. One way to categorise functions is to let
each function belong to the subsystem which it mainly implies requirements on rather than the source
attribute. Categorizing functions by subsystems tends to lead to “carry-over” function realisations
from previous vehicle program, which can be good enough in many cases. Categorizing functions by
source attribute facilitates more novel function realisations, which can be motivated in other cases.
The word “function” has appeared very frequently lately along with development of electrically con-
trolled systems. The function “Accelerator pedal driving” in 3.5.2.1 has always been there, but when
the design of it changed from mechanical cable and cam to electronic communication and algorithms
(during 1990’s) it became much more visible as a function, sometimes referred to as “electronic throt-
tle”. The point is that the main function was there all the time, but the design was changed. The change
of design enabled, or was motivated by, improvement of some sub-functions, e.g. idle speed control
which works better in a wider range of engine and ambient temperature.
At some places, the compendium emphasizes the functions by adding an asterisk “*” in section head-
ing and a “Function definition” in the following typographic form:
Function definition: {The Function} is the {Measure} … for {Fixed Conditions} and certain {Parameter-
ized Conditions}.
The word “conditions” should be understood as a manoeuvre, operation or use case. It is often possible
and efficient to define multiple measures from one “condition”.
The {Measure} should be one unambiguously defined measure (such as time, velocity or force) of
something the vehicle does. The {Measure} is ideally a continuous, objective and scalar physical quan-
tity, subjected for setting a requirement on the vehicle. The {Fixed Conditions} should be unambigu-
ously defined and quantified conditions for the vehicle and its surroundings. The keyword “certain”
identifies the {Parameterized Conditions}, which need to be fixed to certain numerical values or proba-
bilistic distributions, before using the Function definition for requirement setting, see 1.4.3.
Since the term “Function” is defined very broadly in the compendium, these definitions become very
different. One type of Function definition can be seen in “3.2.3.1 Top Speed *”, which includes a well-
defined measure. Another type of Function definition is found in “3.5.2.3 Anti-Lock Braking System,
ABS *” and “4.3.3 Under-, Neutral- and Over-steering *”. Here, the definitions are more on free-text for-
mat, and an exact measure is not so well defined.
13
Introduction
vehicle. However, the structure below does neither reach the full level of details, nor the amount, of
functions and requirements. Also, the structure does not show any requirements broken down to sub-
systems.
• Transport Efficiency
(These measures require a certain transport operation to be defined, see 3.3.1.)
The inverse of Transport efficiency is Transport cost in [€/ ] or [€⁄(𝑡𝑜𝑛 𝑘 )]. This is the sum
of several terms:
o Energy consumption; which depends on energy [€⁄𝑁 ] or fuel price [€⁄ 3 ], see
3.3.4.1
o Transport time costs; which depends on driver salary for goods transports and on occu-
pant travel time cost [€⁄ℎ] for passenger transports, see 3.3.4.3
o Component wear; e.g. tyre wear (2.2.8), battery degradation, etc. The cost can be quan-
tified via a life quantity (e.g. [ ] tread thickness for tyres) and a DegradationQuantity
(e.g. [ /𝑠] for tyres) and a Replacement cost [€/𝑟𝑒𝑝 𝑒 𝑒𝑛𝑡].
• Safety
(These measures requires a certain manoeuvre or traffic scenario. When adding several
measures, one have to consider severity and probability of those.)
o Speed reduction before collision; in [m/𝑠]
o Avoidance speed; in [m/s] maximum entry speed to the traffic scenario, or to a cone
track, lane change or tightening curve
o Longitudinal Deceleration; in [s] for decelerating from one speed to another speed in a
straight-line or curve.
• User Experience
o Ride comfort
▪ Vertical stationary oscillations; in vertical amplitude per road displacement am-
plitude [(m/𝑠 2 )/m] for a certain irregularity of road at certain speed
▪ Vertical and longitudinal transient shock; in [m/𝑠 3 ] vertical and longitudinal
peak when driving over certain cleat on road in certain speed
o Performance
▪ Longitudinal Acceleration ([s] for accelerating from certain speed to another
certain speed)
▪ Lateral SteadyState Acceleration ([m/𝑠 2 ] when driving in certain radius)
o Driveability, Handling and Road-holding
▪ Longitudinal Acceleration margins ([m/𝑠 2 ] or subjective assessment peak or av-
erage during certain transport task)
▪ Transport time ([s] or subjective assessment around a certain handling track)
▪ Pedal response ([(m/𝑠 2 )/m] or [(m/𝑠 2 )/N]) for step pedal apply
▪ Steering wheel response ([(rad/s)/deg]) or subjective assessment for step or
oscillating steering
o Trust (requirements here are far from well established)
▪ Subjective assessment after longer use (maybe ≥a week) of the vehicle. Many
and/or well-trained test users needed. A questionnaire can be used to organise
subjective assessment in different traffic situation, low/high speed, light/dark,
dry/slippery road, longitudinal interaction with other road users, lateral inter-
action with lane edges, etc.
1.4.3 Requirements
A requirement shall be such that it is possible to verify how well a product fulfils it. A requirement on
the complete vehicle is typically formulated as:
“The vehicle shall … do something or have measure … < 𝑜𝑟 > 𝑜𝑟 ≈ … number [unit] …
… under certain conditions.”
Examples: The vehicle shall…
• … accelerate from 50 to 100 km/h in <5 s when full acceleration pedal. On level road.
14
Introduction
• … decelerate from 100 to 0 km/h in <35 m when brake pedal is fully applied, without locking
any rear wheel. On straight and level road.
• … turn with outermost edge on a diameter <11m when turning with full steering at low speed.
• … have a characteristic speed of 70 km/h (10 km/h). On level ground and high-friction road
conditions and any recommended tyres.
• … give a weighted RMS-value of vertical seat accelerations < .5 ⁄𝑠 2 when driving on road
with class B according to ISO 8608 in 100 km/h.
• … keep its body above a 0.1 m high peaky two-sided bump when passing the bump in 50 km/h.
To limit the amount of text and diagrams in the requirements it is useful to refer to ISO and OEM spe-
cific standards. Also, it is good to document the purpose and/or use cases with the requirement.
The above listed requirements stipulate the function of the vehicle, which is the main approach in this
compendium. Alternatively, a requirement can stipulate the design of the vehicle, such as “The vehicle
shall weigh <1600 kg” or “The vehicle shall have a wheel base of 2.5 m“. The first type (above listed)
can be called Performance based requirement. The latter type can be called Design based requirement or
Prescriptive requirement and such are rather “means” than “functions”, when seen in a function vs
mean hierarchy. It is typically desired that requirements are Performance based, else they would limit
the technology development in society.
15
Introduction
split,
-2 5 deg 0.2 (lo) 25
left/right=lo/hi
step up
step down
Figure 1-4: § An example of “Span Matrix” and “Experiment Matrix” for brake performance. Span
matrix could generate 4 experiments. Experiment matrix shows that only 4 of
them are chosen.
Note that variation of (vehicle) design parameters is very essential to find best product design. So, each
experiment needs to be performed for many vehicle designs. This can be documented either as adding
more columns and rows to the experiment matrix or in a vehicle design alternative matrix, separate
from the experiment matrix. The latter way is proposed with the motivation that vehicle design alter-
natives rather belongs to the design/optimization process than the requirement setting, see Figure
1-8. If separate experiment matrix with 𝑛 rows and vehicle design alternative matrix with rows, the
total number of experiments to carry out will be maximum 𝑛 .
Also note that the required numerical value of the requirement measure (e.g. braking distance) typi-
cally change between the experiments (e.g. differently long braking distance is required for different
road friction). It is even so that, in situations where we have use for an experiment matrix, it is typi-
cally more relevant with comparisons of the measure between different vehicle designs than checking
if the measure is > or < an absolute value.
16
Introduction
Reqmnt
Front axle Design
Reqmnt
suspension design Decision Reqmnt
(High level)
Vehicle activity by “front
Reqmnt Design suspension engineers”
design Decision
activity by Reqmnt
Rear axle Design
“attribute Reqmnt
engineers” suspension design Decision
activity by “rear Reqmnt
suspension engineers”
1 2 2
3
Figure 1-5: § Requirement decomposition as a design activity. A design decision on one level
generate the requirement to the next level. Here exemplified with front and rear axle. Red arrows
indicates different design loops.
Note that different experience and different computation/simulation models are useful in the first and
second design step. In each design step, the models needed has to verify how well a certain design of a
system fulfils the requirements on the same system, where the system can be ether complete vehicle,
front suspension or rear suspension. When all low-level design is done, a 3rd type of model is useful, to
verify how well certain designs of all subsystems fulfil the requirements on the complete vehicle. The
last-mentioned verification is traditionally done in prototype vehicles, but with a good virtual verifica-
tion architecture in at the vehicle manufacturer, it becomes increasingly possible to do virtually. The
following figure is a development of Figure 1-2 to show the 3 verification/design loops, which typically
requires different models and computations. Note that number 3 also requires organisation and data-
bases and tools, i.e. a virtual verification architecture.
Product planning Customers / Users
Describe Complete Use of
Vehicle Attributes Vehicle
Complete Vehicle Attributes Designed Vehicle
Quantify Requirements on Validate
Complete Vehicle Complete Vehicle
Requirements on Complete Vehicle Functions 3 Designs of all Subsystem
1 Verify Subsystem Design
Decompose for Requirements on
Requirements Complete Vehicle
Figure 1-6: § V-process where 3 types of verification is shown: 1,2 and 3. Compare with Figure 1-2.
Red arrows indicates the same design loops as in previous figure.
17
Introduction
function. One example is that the “2.4 Propulsion System”, “3.2.1 Traction Diagram” and “3.2.2 Power
and Energy Losses” are placed before “3.2.3.1 Top Speed *”. Functions only appear in Chapters 3, 4 and
5.
1.5 Engineering
Engineering (or Design Engineering, in Swedish often “Ingenjörsvetenskap”, in German “Ingenieurwis-
senschaft”), as a science has an important portion of Synthesis. As support for synthesis, it also relies
on Analysis, Inverse analysis or (Nature) Science, see Figure 1-7. Figure 1-8 distinguishes between Anal-
ysis and Synthesis, which shows that Analysis is a step in the whole design loop. The actual Design step
requires Synthesis. The overall purpose is to propose a design, e.g. numerical values of Design parame-
ters of a product. The distinction between Analysis and Inverse Analysis can be made when there is a
natural causality (“cause-effect-direction”).
In 1.5 some useful general methods and tools are described. Parts of these are probably repetition for
some of the reader’s previous education, in mechanical and control engineering. In the end of 1.5,
there is a stronger connection to the vehicle engineering.
18
Introduction
Input Output
System
Laws of Nature
Given Find Process
Input, Laws of Nature, System Output Analysis
Output, Laws of Nature, System Input Inverse Analysis
Input, Output, System Laws of Nature (Nature) Science (induction)
Input, Output, Laws of Nature System Engineering Design (Synthesis)
Dixon, J.R., (1966) Design Engineering, Inventiveness, analysis and Decision Making
Figure 1-7: Engineering Design and related activities. Picture from Stefan Edlund, Volvo Trucks.
Vehicle Design
Requirements (Required output)
Design parameters are varied to find best design Operation/Manoeuvre parameters are
(optimization) or good enough design (satisficing). varied in order to make a robust design.
Figure 1-8: How Analysis/Synthesis and Design/Operation parameters appears in Vehicle Design.
19
Introduction
Formulate
Evaluate
engineering Design/Re Real-world requirement
OK Propose
task -design testing fulfilment Design
(problem)
NOK
Figure 1-9: (Dynamic) Modelling stages. Real versus Theoretical world. The dashed boxes in the
background indicate that same design has to fulfil multiple requirements, which might require
multiple models.
1.5.1.1.2 Physical Modelling
In this stage, one should generate a physical model, which in this compendium means sketches and
text. Some call this an engineering model and some do not call it a model at all. Free-body diagrams
(1.5.2.1), data flow diagrams (1.5.1.8) and operating conditions (1.6.2) are always important ingredi-
ents in the physical model of a vehicle, but also other domains than mechanical are often represented,
such as hydraulics, electrics, driver and control algorithms. The physical model shall clarify assump-
tions/approximations/simplifications, e.g. rigid/elastic, inertial/massless, continuous/time sampled
algorithms, small angles, etc. Discrete dynamics, see 1.5.1.4, such as ideal dry friction, backlash or end
stop does not exist in real world, so such assumptions are typical examples of clear differences be-
tween real system and physical model. How to model here should be based on what phenomena is
needed to be captured and what variations (e.g. which design parameters) are intended for simula-
tions with the model. The parameters and variables are first defined in the drawig, but might need ad-
ditional text. To distinguish between whether a physical quantity is a parameter or a variable, see
1.5.1.3, is a very important modelling decision. A physical model is often not executable, but it can be.
One example of executable format is a model build in an MBS tool, see 1.5.4.4 and Figure 1-39.
1.5.1.1.3 Mathematical Modelling
In this stage, one should generate a mathematical model, which means equations. Basically, it is
about finding the right equations; equally many as unknowns. For dynamic systems, the unknowns are
unknown variables of time. In the mathematical model, the assumptions are transformed into equa-
tions. For dynamic systems, the equations form a “DAE” (Differential-Algebraic system of Equations). It
is seldom necessary to introduce derivatives with respect to other independent variables, such as posi-
tions, i.e. one does seldom need PDE (Partial Differential Equations). The general form of a DAE:
• 𝒇𝑫𝑨𝑬 (𝒛 𝒛 𝑡) 𝟎
(Here, the function 𝒇𝑫𝑨𝑬 is an operator which include only algebraic operations. An alternative nota-
tion would be 𝒇𝑫𝑨𝑬𝟐 (𝒛(𝑡) 𝑡 𝒑) 𝟎 , but then the function 𝒇𝑫𝑨𝑬𝟐 is an operator which include both al-
gebraic and differential operations.)
The 𝒛 are the (dependent) variables and 𝑡 is time (independent variable). The mathematical model is
complete only if there are equally many (independent) equations in 𝒇𝑫𝑨𝑬 as there are variables in 𝒛,
dim(𝒇𝑫𝑨𝑬 ) dim(𝒛). (Since DAE, we don’t count 𝒛 and 𝒛 as different variables.) Note that one can be
interested if treating also cases where dim(𝒇𝑫𝑨𝑬 ) < dim(𝒛). It is when dim(𝒛) dim(𝒇𝑫𝑨𝑬 ) of the var-
iables in 𝒛 can be considered as known, so called, input variables (𝑡). The notation “(𝑡)” marks that ,
generally, might have to be known DAE variables, i.e. their time derivatives also have to be known.
It is not only a question of finding suitable equations, but also to decide parameterisation, which is how
parameters are defined and related to each other. Parameterisation should reflect a “fair” comparison
20
Introduction
between different design parameters, which often requires a lot of experience of the product and the
full set of requirements on the vehicle. It is here the engineering science comes in, while the equations
is rather physics and mathematics. To underline the parameters, 𝒑, they can be included in the DAE
form as 𝒇𝑫𝑨𝑬 (𝒛(𝑡) 𝒛(𝑡) 𝑡 𝒑) 𝟎 .
Selection of output variables is important so that output variables are enough to evaluate the require-
ments on the system. Selecting more might drive unnecessary complex models.
The Mathematical model is acausal, i.e. describes relations between the variables, not how and in
which order they are computed. A Mathematical model can be equations on a piece of paper or in a
word processing format and then it is not executable. However, writing the Mathematical model in
Modelica format makes it executable using Modelica tools, see 1.5.4.5.
§ Differentiation Order
Above, the ODEs are assumed to be of first order. There is a strong tradition to in mechanical engineer-
ing to model with 2nd order differential equations, where accelerations of inertial bodies appear as 2 nd
derivative of position. However, numerical methods for solving first order differential equations are
much more mature and Vehicle Dynamics mainly aims at such. Hence, the compendium aims at first
order differential equations.
It is easy to go from few equations to higher order to many first order differential equations; variables
appearing in 2nd order derivative ̈ are replaced with 𝑛 𝑤 and one equation 𝑛 𝑤 is added.
The opposite is not generally as easy, but often possible. An example is a linear model 𝒛 𝑨 𝒛 , which
can be converted from many 1st order to fewer 2nd order becomes as follows:
𝒛 𝑨 𝑨𝟏𝟐 𝒛𝟏 differentiate eliminate
[ 𝟏 ] [ 𝟏𝟏 ] [𝒛 ] ⇒ { }⇒{ }⇒
⏟𝒛𝟐 ⏟𝑨𝟐𝟏 𝑨𝟐𝟐 ⏟ 𝟐 equation for 𝒛𝟏 𝒛𝟐 (and 𝒛𝟐 )
𝒛 𝑨 𝒛
⇒ 𝒛̈ 𝟏 (𝑨𝟏𝟏 + 𝑨𝟏𝟐 𝑨𝟐𝟐 𝑨𝟏𝟐𝟏 ) 𝒛𝟏 + 𝑨𝟏𝟐 (𝑨𝟐𝟐 𝑨𝟏𝟐𝟏 𝑨𝟏𝟏 𝑨𝟐𝟏 ) 𝒛𝟏 𝟎
Another way to go from higher order to lower order differential equations is to go via Laplace trans-
form. This is exemplified on same equations as above, but only for scalar 𝒛𝟏 and 𝒛𝟐 :
𝒛 𝐴 𝐴2 Laplace 𝐴 𝐴2
[ 𝟏] [ ] [ ] ⇒{ }⇒ 𝑠 [ ] [ ] [ ] ⇒
⏟𝒛𝟐 ⏟𝐴2 𝐴22 ⏟ 2 transform 2 𝐴2 𝐴22 2
𝒛 𝑨 𝒛
𝐴 𝐴2
⇒ ([ ] 𝑠 𝑰) [ ] 𝟎 ⇒ {eliminate 2 } ⇒
𝐴2 𝐴22 2
⇒ 𝑠2 (𝐴 + 𝐴22 ) 𝑠 + (𝐴 𝐴22 𝐴 2 𝐴2 ) 0 ⇒
inverse
⇒ { Laplace } ⇒ ̈ (𝐴 + 𝐴22 ) + (𝐴 𝐴22 𝐴 2 𝐴2 ) 0
transform
21
Introduction
because the effect from constant terms would obviously appear twice in 𝑓 (𝑢 ) + 𝑓 (𝑢2 ) but only
once in 𝑓 (𝑢 + 𝑢2 ). So, it is often motivated to look for a linear form.
If 𝑨 is non-singular we can find a form which is linear in state derivatives by offsetting 𝒙 by replacing 𝒙
with (𝒙𝟎 + 𝒙𝒐𝒇𝑨 ), where 𝒙𝟎 𝑨 𝟏 𝒌𝒙 . The result becomes:
• 𝒙𝒐𝒇𝑨 𝑨 𝒙𝒐𝒇𝑨 + 𝑩 𝒚 (𝒌𝒚 + 𝑪 𝒙𝟎 ) + 𝑪 𝒙𝒐𝒇𝑨 + 𝑫
⏟
𝒌𝒚𝒐𝒇𝑨
If 𝑨 is singular but 𝑩 is non-singular, we can find a form which is linear in state derivatives by offsetting
by replacing with ( 𝒐𝒇𝑩 + 𝟎 ), where 𝟎 𝑩 𝟏 𝒌𝒙 . The resulting form becomes:
• 𝒙 𝑨 𝒙+𝑩 𝒐𝒇𝑩 𝒚 (𝒌𝒚
⏟ 𝑫 𝑩 𝟏
𝒌𝒙 ) + 𝑪 𝒙 + 𝑫 𝒐𝒇𝑩
𝒌𝒚𝒐𝒇𝑩
It is often enough to find a form which is linear in state derivatives, as shown above. However, if both 𝑨
and 𝑫 are non-singular, we can find a Linear Form, i.e. linear in both state derivatives and outputs, i.e.
without both 𝒌𝒙 and 𝒌𝒚 . This is done by offsetting by replacing 𝒙 with (𝒙𝒐𝒇𝑨𝑫 + 𝒙𝟎 ) and with
𝒙𝟎 𝟏 𝒌
𝑨 𝑩
( 𝒐𝒇𝑨𝑫 + 𝟎 ), where [ 𝟎] [
𝑪 𝑫
] [ 𝒙]
𝒌
. The resulting form becomes:
• 𝒙𝒐𝒇𝑨𝑫 𝑨 𝒙𝒐𝒇𝑨𝑫 + 𝑩 𝒐𝒇𝑨𝑫 𝒚 𝑪 𝒙𝒐𝒇𝑨𝑫 + 𝑫 𝒐𝒇𝑨𝑫
(The used terms, affine and linear, can be confusing because the common meaning of the word linear
can include a constant term, while absence of constant term would instead be called proportional.)
If the model is affine (or linear), 𝑨 𝑩 𝑪 𝑫 can be found by identification of the coefficients for the com-
ponents of 𝒙 and . Manual manipuations or symbolic tools can be used. If the model is non-linear, one
can sometimes motivate to neglect small terms so that the model becomes affine. A more general and
automated way to find an affine model is to make a Taylor expansion (symbolic or numerical) around
(𝒙𝟎 𝟎 𝑡 ). If the model is affine and there is no explicit dependency of 𝑡, the Taylor expansion will be
exact:
𝒙 𝒇𝑶𝑫𝑬 (𝒙 ) {affine}
𝜕𝒇𝑫𝑨𝑬 𝜕𝒇𝑫𝑨𝑬
𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎) +( )|𝒙=𝒙𝟎 (𝒙 𝒙𝟎 ) + ( )|𝒙=𝒙𝟎 ( 𝟎) {affine}
𝜕𝒙 𝜕
= 𝟎 = 𝟎
𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) + 𝑨 (𝒙 𝒙𝟎 ) + 𝑩 ( 𝟎) where 𝑨 and 𝑩 independent of 𝒙𝟎 and 𝟎 ⇒
⇒ 𝒙𝒐𝒇 𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) + 𝑨 ⏟ (𝒙 𝒙𝟎 ) + 𝑩 ⏟
( 𝟎)
𝒙𝒐𝒇 𝒐𝒇
If (𝒙𝟎 𝟎 ) is a steady state condition it means same as that 𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) 0, so the constant term line-
arization gets the form linear in state derivatives 𝒙𝒐𝒇 . Similar Taylor expansion can be done with 𝒈 for
outputs , to find a linear (or at least affine) from also for outputs.
Models with explicit dependency of 𝑡 are treated in (Bruzelius, 2004).
Note that the model was assumed to be affine. This meant that 𝑨 and 𝑩 becomes independent of 𝒙𝟎
and 𝟎, as opposed to a non-linear model, see 1.5.1.1.9.1.
1.5.1.1.4 Explicit Form Modelling
In this stage, one should generate an Explicit Form model, which means equations rearranged to as-
signment statements, i.e. to an explicit form (algorithm) which outputs the state derivatives. The ex-
plicit form model is generally causal. You probably recognize this formulation as “ODE” (Ordinary Dif-
ferential Equation) or “IVP” (Initial Value Problem). So, the general form is:
• 𝒙 ← 𝒇𝑶𝑫𝑬 (𝒙 𝑡) 𝒚 ← 𝒈(𝒙 𝑡)
(Here, the function 𝒇𝑶𝑫𝑬 is an operator which include only algebraic operations.)
We have essentially the same set of equations in the Mathematical model as in the Explicit Form
model. However, seen as Explicit Form model, we count 𝑥 and 𝑥 as different variables. So far, this
means that we have too few equations. To keep equality between number of variables and equations,
we can count in also the equations 𝑥 ← ∫ 𝑥 𝑡 .
22
Introduction
The 𝒙 is the state variables, is the input variables (𝑡), and 𝒚 the output variables. In the mathemati-
cal model, there was no distinction between different dependent variables in 𝒛. However, to reach the
explicit model, each variable in z has to be identified as belonging to either of 𝒙 𝒚 or . Simply speak-
ing, states 𝒙 are the variables which appears both as and , inputs are variables that cannot be
solved for within the system of equations and outputs 𝒚 are the remaining variables. Since the state
variables are identified, the Explicit Form is sometimes called State Space Form.
A dataflow diagram (see Figure 1-16 and Figure 1-36) is a graphical representation of the explicit
form. It is drawn using blocks with input and output ports and arrows representing signals between;
integration is represented by integration blocks with 𝒙 as input and 𝒙 as output.
§ Semi-Explicit form
If the explicit form cannot be found, there are time integration methods also for solving the semi-ex-
plicit form: 𝒙 ← 𝒇 𝒆 𝒊 (𝒙 𝒚 (𝑡) 𝑡) 𝟎 𝒈 𝒆 𝒊 (𝒙 𝒚 (𝑡) 𝑡) .
A generalisation of the semi-explicit form is when we cannot even find explicit expression for all state
derivatives 𝒙. We can still formulate an Explicit form model, but then using “implicit form expres-
sions”. As extreme example, when no variables can be solved for:
• [𝒙 𝒚] ← 𝑠𝑜 𝑒(𝒇𝑶𝑫𝑬 𝒊 𝒑𝒍 (𝒙 𝒙 𝒚 (𝑡) 𝑡) 𝟎)
The operator “𝑠𝑜 𝑒” is here meant to be implemented as a (numerical) iteration during the Computa-
tion (compare 1.5.1.4.3). The only difference to the original 𝒇𝑫𝑨𝑬 is that the states 𝒙 are identified
among the variables 𝒛. But, it should also be noted that this is no general cure to avoid doing Mathe-
matical manipulations, because far from all iterations succeed; so, explicit form expressions should be
strived for as far as possible! Often, some parts of the Explicit form can be explicit form and only a few
variables appear in a 𝑠𝑜 𝑒 expression. When implicit expressions appear in the Explicit form model,
we have a situation where causality is not defined for the involved variables, although the model is on
an overall Explicit form.
§ Implicit form
If neither explicit or semi-explicit form cannot be found, there are time integration methods also for
solving the implicit form: 𝒓𝒆 𝒊𝒅 𝒂𝒍 ← 𝒇𝒊 𝒑𝒍𝒊𝒄𝒊𝒕 (𝒙 𝒙 𝒚𝟏 𝒚𝟐 𝒚𝟐 (𝑡) 𝑡) . The difference with the model
in 1.5.1.1.3 is that the states 𝒙 are identified. The 𝒓𝒆 𝒊𝒅 𝒂𝒍 𝟎 is solved numerically and iteratively
yielding {𝒙 𝒚𝟏 𝒚𝟏 𝒚𝟐 } at each model evaluation. Then 𝒙 ← ∫ 𝒙 𝑡 . Both 𝒚𝟐 and 𝒚𝟐 are found without
time integration, which means that dim(𝒚𝟏 ) extra equations are needed compared to 𝒇𝑫𝑨𝑬 in 1.5.1.1.3.
These are found through differentiation of some equations, see more 1.5.1.6.
An alternative solution method time integrates also 𝒚𝟐 ← ∫ 𝒚2 𝑡 without fulfilling the extra equa-
tions. Instead the extra equations are used to correct 𝒙 and 𝒚𝟐 to eachother between each time step.
This is used in many MBS tools, such as Adams, see 1.5.4.4.
Time Sampling and Time Events
In vehicle dynamics, the dynamics occurring in the digital control is essential. A control algorithm can
of course be modelled with continuous equations, but the effect of time sampling and signal communi-
cation delays are often essential to include. A mathematical method to model this phenomenon is
“time events”.
Time events are similar to, but not same as, discrete dynamics in 1.5.1.4. The similarity is that both
time events and state events (used for discrete dynamics) uses transition conditions, 𝒉 > 𝟎. But for
time events, the transition conditions are only a function of time 𝒉(𝑡). Time events appear since algo-
rithms are executed on digital processors or are delayed sin signal communication. Time events are
generally more established and easier to get physically consistent than the more general state events,
𝒉 𝒉(𝒙𝒄 𝒙𝒅 𝑡).
In Modelica, a continuous model with its time sampled controller can be modelled as in Figure 1-10.
The example is speed control of a vehicle in uphill. The longer sample time gives an unstable solution.
23
Introduction
24
Introduction
end when;
𝑢
-0.05
//discrete, EulerBwd: deryd = (yd-pre(yd))/SampleTime
when sample(0, SampleTime) then
udB=u; -0.10
derydB = -ydB+udB; //model
ydB=pre(ydB)+derydB*SampleTime;
end when; -0.15
end DiscreteIntegrator; 0.0 2.5 5.0 7.5
• For linear or linearized models and simple excitation one can find the solution without time inte-
gration, by using the “matrix exponential”, 𝑒 𝑴𝒂𝒕𝒓𝒊𝒙 : The solution to a linear system without input,
𝐴 , starting from a given initial state 𝒙(0) 𝒙𝒊𝒗 is found as 𝒙(𝑡) 𝑒 𝑨 𝒙𝒊𝒗 . An alterna-
tive way, using a solution ansatz, is shown in 4.5.4.1.
• Frequency domain analyses. Linear models or models linearized around an operating condition
by differentiating and 𝐴 𝑓 ⁄ 𝑥 and 𝐵 𝑘 𝑓 ⁄ 𝑢𝑘 . The matrices are very useful for eigen-
modes, eigen-frequencies, step response, stability, or use as model base in model-based control
design methods.
• Stability analysis is to study when small disturbances lead to large response and how. (In linear
models, when limited disturbances lead to unlimited solutions, typically oscillating or exponen-
tially increasing.) Stability analysis is normally done for one operating point [𝒙𝟎 𝟎 ], around
which one linearizes. Instability is detected as when any eigenvalue to 𝑨 has positive real parts.
If there are inputs (dim( ) > 0), the stability is “for the open loop system”, e.g. for the vehicle
without driver. If the equations for is included in the model, there will be no 𝑩 matrix, and the
stability is then “for the closed loop system”, e.g. for the vehicle with driver. Adding a driver can
make the vehicle with driver more or less stable.
• Optimization. Either optimizing a finite number of defined design parameters or time trajecto-
ries, e.g. 𝑢 (𝑡). There are many optimization methods, ranging from trial-and-error to mathemati-
cally/numerically advanced gradient based or evolutionary inspired methods.
§ Closed Form Solution of Model on Affine Form
It will here be shown that a closed form solution can be found for the affine form. This is an alternative
to simulation. The following example shows how to find the solution for 𝒙 𝒌 + 𝑨 𝒙 with 𝒙(0) 𝒙𝟎 .
𝑇
Make an ansatz: 𝒙 𝒙𝒄𝒐𝒏 𝒕 + 𝑽 [𝑒 𝜆1 𝑒 𝜆2 ⋯ 𝑒 𝜆dim(𝒙) ] 𝒙𝒄𝒐𝒏 𝒕 + 𝑽 𝒆𝝀 ⇒
⇒ 𝒙 𝑽 𝐃𝐢𝐚𝐠([𝜆 𝑒 𝜆1 𝜆2 𝑒 𝜆2 ⋯ 𝜆dim(𝒙) 𝑒 𝜆dim(𝒙) ]) 𝑽 𝐃𝐢𝐚𝐠(𝜆 𝜆2 ⋯ 𝜆dim(𝒙) )
𝒆𝝀 𝑽 𝑫 𝒆𝝀
Insert: 𝑽 𝑫 𝒆𝝀 𝒌 + 𝑨 (𝒙𝒄𝒐𝒏 𝒕 + 𝑽 𝒆𝝀 )
Identify constant terms: 𝟎 𝒌 + 𝑨 𝒙𝒄𝒐𝒏 𝒕 ⇒ 𝒙𝒄𝒐𝒏 𝒕 𝑨 𝒌
Identify exponential terms: 𝜆 𝑰 𝑨 ∀ ⇒ [𝑽𝒏𝒐𝒓 𝑫] eig(𝑨) 𝑛 𝑽 𝑽𝒏𝒐𝒓 𝐝𝐢𝐚𝐠( )
𝑽𝒏𝒐𝒓 is a matrix with normalized eigenvectors as columns. is a column vector with magnitudes of
eigenvectors.
The solution before fulfilling initial conditions becomes: 𝒙 𝑨 𝒌 + 𝑽𝒏𝒐𝒓 𝐃𝐢𝐚𝐠( ) 𝒆𝝀
25
Introduction
𝑥
𝑥
Rewritten to system of system of difference equations with 𝑁 + (𝑁 + ) 𝑁 + unknowns
( 2 ⋯ 𝑁 𝑥 𝑥2 ⋯ 𝑥𝑁 ⋯ 𝑁 ) and 𝑁 + (𝑁 + ) 𝑁 + equations:
𝑣𝑖+1 𝑣𝑖
+
𝒇𝒙𝑫𝑨𝑬 [ 𝑖+1
𝑖+1
𝑖
𝑖
] 𝟎 for 0 ⋯ 𝑁
𝑖+1 𝑖
𝒇𝒚𝑫𝑨𝑬 [ 𝑥] 𝟎 for 0 ⋯ 𝑁
There are many ways to solve such (algebraic) system of equation. One do not need a dynamic tool,
since the independent variable time is removed. In case of a linear dynamic model 𝒙(𝑡) 𝑨 𝒙(𝑡) , one
can assembly a large algebraic system of equations with “another 𝑨 matrix”, here called 𝑨𝒍𝒂𝒓𝒈𝒆 :
𝑨𝒍𝒂𝒓𝒈𝒆 [𝒙𝒂𝒓𝒓𝒂𝒚 𝒚𝒂𝒓𝒓𝒂𝒚 ] 𝒃
Note that 𝒙(𝑡) [𝑥 (𝑡) ⋯ 𝑥dim( ) (𝑡)] means an array of variables, while 𝒙𝒂𝒓𝒓𝒂𝒚 means a longer array
(or column vector) 𝒙𝒂𝒓𝒓𝒂𝒚 [[𝑥 (𝑡 ) ⋯ 𝑥 (𝑡𝑁 )] ⋯ [𝑥dim( ) (𝑡 ) ⋯ 𝑥dim( ) (𝑡𝑁 )]]. Our example, with
𝑁 , becomes:
26
Introduction
0 0 0 𝑡 +𝑡 0 0
[ ] [ ] [ ] [ ] [ ]
0 0 0 𝑡 + 𝑡2 0 2 0
0 0 0 0 0 0 𝑥 𝑥 + (𝑡 𝑡 )
[ ] [ ] [ ] [𝑥 ] [ ]
𝑡
𝑡2 0 0 0 0 2 0
0 0 0 0 0 0 𝑥
[0 0] [ 0] [0 0] [ ] [ 0 ]
[
⏟ ] ⏟ 2 ]
[ [
⏟ 0 ]
0 0 0 0 0
𝑨𝒍𝒂𝒓𝒈𝒆 𝒙𝒂𝒓𝒓𝒂𝒚 𝒃
However, a Modelica tool is a dynamic tool and it can be useful, because a Mathematical model in Mod-
elica can, with minor code editing, be changed to a difference equation. The “traditional” Modelica
model in Figure 1-37 can be edited to a variant with difference equations as shown in the following
figure. (Some Modelica tool have this rewriting as a build-in and selectable automatic feature, typically
used for real time simulation.) Another derivative approximation is used in the figure, to show that
also “implicit” approximations as “midpoint” can be used.
v F x model ExampleModel_InlineIntegration
4 parameter Real m=2, c=3;
Real v, x, F;
-2
𝑥 end for;
Figure 1-12: § Example from Figure 1-37 implemented as difference equation model in Modelica.
Finally, it should be reminded about that the form described above is “contaminated” by containing
the derivative approximation, so it is a mixture of “1.5.1.1.3 Mathematical Modelling” and “1.5.1.1.5
Computation”.
§ Parameter vs Trajectory Optimization, Cost vs Constraints
So far, we have assumed that the model is used to predict how the system will behave (simulation). An
alternative use of the model is optimization. As engineering task, one can differ between parameter op-
timization or trajectory optimization:
• 𝑛 𝒑 𝑡ℎ 𝑡 𝑛 𝑒𝑠 𝑓 (𝒑) 𝑠𝑢 ℎ 𝑡ℎ 𝑡 𝒈𝒍𝒊 𝒊𝒕𝑪𝒐𝒏 𝒕𝒓 (𝒑) < 𝟎
• 𝑛 𝒛 𝒛(𝑡) 𝑡ℎ 𝑡 𝑛 𝑒𝑠 𝑓 (𝒛) 𝑠𝑢 ℎ 𝑡ℎ 𝑡 𝒈𝒍𝒊 𝒊𝒕𝑪𝒐𝒏 𝒕𝒓 (𝒛) < 𝟎
Parameter optimization has (a finite number of) design parameters 𝒑, while trajectory optimization
has (a finite number of) design trajectories 𝒛(𝑡). A trajectory has an infinite number of values, which
makes a big difference. The first can be to find the best gear ratios (𝒑 [𝑟 𝑟2 … 𝑟6 ]) and the cost
function 𝑓 includes a simulation of the dynamic system, e.g. modelled by a driver model and a vehicle
model. The latter can be to find the best pedal and gear selection trajectories (𝒛
[𝑝𝑒 (𝑡) 𝑒 𝑟(𝑡)]). The cost function also includes a simulation, e.g. a simulation of a the same vehi-
cle model but without the driver model. The simulation represents a look into the future to find out
27
Introduction
what to do from present time instant and some time on, until a new decision is computed. An optimiza-
tion problem can be both types, i.e. include both design parameters 𝒑 and design trajectories 𝒛(𝑡).
The cost 𝑓 can sometimes be evaluated without simulation, 𝑓 𝑓 (𝑡 ) 𝑓 (𝒛𝟎 ). Then the optimi-
zation problem collapses to what to do at present time instant 𝑡 . We can then find a 3rd categorhy of
optimization, momentaneous variable optimization:
• 𝑛 𝒛𝟎 𝒛(𝑡 ) 𝑡ℎ 𝑡 𝑛 𝑒𝑠 𝑓 (𝒛𝟎 ) 𝑠𝑢 ℎ 𝑡ℎ 𝑡 𝒈𝒍𝒊 𝒊𝒕𝑪𝒐𝒏 𝒕𝒓 (𝒛𝟎 ) < 𝟎 ∈ ℝ
This might look very similar to the parameter optimization because both 𝒑 and 𝒛𝟎 are finite numbers of
scalars. However, 𝒑 is to be decided during product development while 𝒛𝟎 (typically requests to actua-
tors) has to be computed during product use, which is also true for 𝒛(𝑡).
The engineering task has to be reformulated to a computation task. Then, some adjustments are often
made to be able to use a particular existing mathematical/numerical method for optimization:
• Objectives can be moved from constraints to cost. Instead of demanding (𝒑) < 0, one add
terms (𝒑) to the cost function 𝑓 (𝒑). The weigth is then chosen so large so that the
original constraint is approximatively fulfilled.
• For trajectory optimization:
o The trajectories often have to be discretized, either as 𝒛 values over a grid in time or as
parameters of curve fits of 𝒛(𝑡). So, when coming to the computation task, the differ-
ence between parameter and trajectory optimization might vanish so same computa-
tional methods can be used.
o The dynamic model can be used also as constraint, which changes computation task to:
𝒈 (𝒛) < 𝟎 𝑛
𝑛 ⋯ 𝑡ℎ 𝑡 𝑛 𝑒𝑠 𝑓 (⋯ ) 𝑠𝑢 ℎ 𝑡ℎ 𝑡 { 𝒍𝒊 𝒊𝒕𝑪𝒐𝒏 𝒕𝒓
𝒈 𝒐𝒅𝒆𝒍𝑪𝒐𝒏 𝒕𝒓 (𝒛) 𝟎
This formulation is necessary if the method is of collocation type, see 1.5.1.1.5.5.
In the engineering task the distinctions cost vs constraints and cost vs dynamic model are often im-
portant. They are often mixed up in the computation task. Misunderstandings between vehicle engi-
neers and computation engineers might therefore occur.
§ Single- or Multi-Objective Optimization
One can say that optimization can only optimize a scalar cost. One can, of course, weight several costs
together to one cost: 𝑓 𝑓 + 2 𝑓2 . This can be called be called multi-objective optimization. It is
then important to understand the units. Consider 𝑓 as having the unit [€], which leads to that you
have to be able to set a number on the weigths [ 2 ⋯ ] with units [€⁄𝑢𝑛 𝑡𝑓 €⁄𝑢𝑛 𝑡𝑓2 ⋯ ]. This is
often difficult to motivate, which makes the interpretation of the found optimum difficult to use for di-
rect design decisions.
However, for parameter optimization, there is a useful variant of multi-objective optimization. It is to
not take the optimization all the way to one optimum, but stay at a relation between design parame-
ters. One can then draw a Pareto front, i.e. a relation between design parameters which forms an enve-
lope in a diagram with the scalar costs on the axes. The same relations can be drawn in a diagram with
design parameters on the axes, which reduces the dimension of the space in which one should look for
the best design parameters.
§ Shooting and Collocation Methods for Optimization
A divider among (trajectory) optimization methods is between those who iteratively find the solution
𝒙 through simulation (time integration) and iteratively update between those simulations (shooting
methods) and those who solves iteratively for both 𝒙 and at the same time (collocation methods).
Using collocation methods, one uses the mathematical model to generate equality constraint
𝒈 𝒐𝒅𝒆𝒍𝑪𝒐𝒏 𝒕𝒓 , see 1.5.1.1.5.3. One also have to discretize the model in time, which converts the differen-
tial equations into difference equations, using a certain derivative approximation, such as 𝒙𝒌+𝟏 ← 𝒙𝒌 +
Δ𝑡 ∙ 𝒙𝒌 𝒙𝒌 + Δ𝑡 ∙ 𝒇(𝒙𝒌 (𝑡𝑘 ) 𝑡𝑘 ) 𝒙𝒌 + Δ𝑡 ∙ 𝒇(𝒙𝒌 𝒌 𝑘 Δ𝑡) . From computational point of view, the
problem is then converted from trajectory optimization to an optimization of a finite number of trajec-
tory values (𝑡 ). Typically, very simple derivative approximations are used, compared to today’s inte-
gration methods for simulation.
28
Introduction
29
Introduction
A general model is not affine and consequently also not linear. They are often called non-linear (but
non-affine would be more correct, because an affine model is normally not called non-linear). There
are many reasons to why models are non-linear: examples of continuous non-linearities are terms of
the type 𝑥 2 , 𝑥 .5, 𝑥 𝑥 , 𝑥 𝑢 , sin(𝑥 ). There are also discrete non-linearities, such as time sampling or
discrete physical phenomena like dry friction, back-lash, etc. Note the difference between an affine
model resulting from physical model without additional approximations, and a linearization of a
non-linear model. The latter means that there are more approximations than what was done in the
physical model. These additional approximations normally have less physical interpretation.
In order to find the numerical solution, one typically does not gain anything on finding an affine form.
But there are reasons for such form; basically the same reasons as mentioned in 1.5.1.1.3.2 for finding
the linear form of a linear model.
If non-linear model, but only continuous non-linearities, one can linearize the models via Taylor expan-
sion and get an approximate model, a linearization on affine form. A Taylor expansion to find an ap-
proximate model in affine form, corresponding as in 1.5.1.1.3.2, would look like:
If the model is non-linear but there is no explicit dependency of 𝑡, the Taylor expansion will be exact:
𝒙 𝒇𝑶𝑫𝑬 (𝒙 ) ≈
𝜕𝒇𝑫𝑨𝑬 𝜕𝒇 non-
≈ 𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) + ( 𝜕𝒙 )|𝒙=𝒙𝟎 (𝒙 𝒙𝟎 ) + ( 𝑫𝑨𝑬 )|𝒙=𝒙𝟎 ( 𝟎) { }
=
𝜕
=
𝑛𝑒 𝑟
𝟎 𝟎
If (𝒙𝟎 𝟎 ) is a steady state condition it means same as that 𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) 0, so the linearization gets
the form linear in state derivatives 𝒙𝒐𝒇 and inputs 𝒐𝒇 , i.e. without constant term. Similar Taylor expan-
sion can be done with 𝒈 for outputs , to find a linear (or at least affine) from also for outputs.
Note that the model was assumed to be non-linear. This meant that 𝑨 and 𝑩 became dependent of 𝒙𝟎
and 𝟎. For an affine model this was not the case, see 1.5.1.1.3.2. The consequence is that the model
can be stable around one (𝒙𝟎 𝟎 ) and instable around another. Therefore, many methods for fre-
quency analysis, control engineering and signal processing is not so applicable if the system is strongly
non-linear.
§ Stability Analysis
Stability is when all components in 𝒙 stays limited (finite) with limited (finite) and over long but fi-
nite time. Stability for a linear model or linearization of a non-linear model is guaranteed if the real
part of all eigenvalues of 𝑨 are negative. For a linear model, the 𝑨 matrix does not depend on 𝒙, so the
stability is relatively easy checked. For a non-linear model one would typically linearize around the
initial value and check sign of real part of all eigenvalues. However, this is not enough, since the solu-
tion can “slip away” to a regime where the signs of real part of the eigenvalues changes. The following
picture gives some examples.
30
Introduction
21
20 all 𝑥 stable
19
18
0 1 2 3
𝑥 0. + sign 𝑥 40 𝑥 40 3/2 𝑥 40 x[4, 1] x[4, 2] x[4, 3] x[4, 4] x[4, 5] x[4, 6] x[4, 7]
12
e 𝜆 <0
11
10
9 all 𝑥 unstable
8
𝑥 0 1 2 3 𝑡
Figure 1-13: § 1st and 2nd rows: Linear models. Stability or instability same for all 𝑥 and follwos
what 𝑅𝑒(𝜆) indicates. 3rd: Non-linear model. Stability regimes are larger than 𝑅𝑒(𝜆) indicates. 4th
row: Non-linear model. Stability regimes are smaller than 𝑅𝑒(𝜆) indicates.
The influence of varying will now be treated. Stability is easiest understood for a model without .
However, if we have a , one often assumes is limited, slowly varying and independent of 𝒙, e.g. is
constant. Then it still holds to look only on eigenvalues of 𝑨 in 𝑨 𝒙 + 𝑩 . The only additional note
here is that 𝑨 can depend on even if is constant. An example is the influence from propulsion force
on lateral dynamics, 𝑓 𝑤 in 4.4.2.
However, if is depending on 𝒙, e.g. 𝑪𝑳 𝒙 we have a closed loop system. Then we have to look
on eigenvalues of 𝑨𝑪𝑳 in 𝑨 𝒙 + 𝑩 𝑨 𝒙 + 𝑩 𝑪𝑳 𝒙 (𝑨 + 𝑩 𝑪𝑳) 𝒙 𝑨𝑪𝑳 𝒙. This can change
the stability totally. A vehicle dynamics related case of this is that we can either assess lateral stability
for vehicle alone, with input steering angle limited and independent or, opposite, assess lateral stabil-
ity for vehicle with a closed loop driver, who steers as a direct function of the vehicle motion, 0. An-
other situation when the 𝑨 matrix changes is when a drvier model includes an own state variable, e.g. a
time delay. Then, both 𝒙 and 𝑨 grows with one dimension.
1.5.1.2 Approximations
As mentioned above, assumptions (or simplifications or approximations) are made during Physical
modelling. These can be directly motivated by vehicle design, such as assuming that some part is rigid
or massless. They can also be directly motivated by the manoeuvre to be studied with the model. Ex-
amples of approximations during Physical modelling are massless bodies, steady state conditions and
small angles. These have implications on the Mathematical modelling stage: some fictive forces do not
appear in the equations, and 𝑛 𝑒 replaces sin( 𝑛 𝑒) in the equations.
We can also assume that some variables or parameters are small. If a term in the final Explicit form
model gets a high order, one can approximate mathematically by removing it. For example, a term
𝑠 𝑁
where 𝑁 ≥ can be removed if the other terms have 𝑁 < . Note that such approximations
can not be done in an equation (during Physical or Mathematical modelling), but in the final expression
31
Introduction
32
Introduction
A very important special case of variables are state variables or states. State variables are given ini-
tial values and then updated through integration along the time interval studied. (Continuous) States
are updated through time integration. Which variables to use as states is not uniquely defined by the
physical (or mathematical) model. See 1.5.2.1.2.
1.5.1.3.3 Signals
Since vehicle dynamics so often requires models of the control algorithms, one often use the words
variable and signal interchangeable. It is suggested to, at least, reflect over the difference:
• A signal can represent an ODE variable with prescribed causality. So, a variable cannot be
represented by a signal before the modelling stage “1.5.1.1.4 Explicit Form Modelling”.
• Signals can appear already in mathematical model (DAE), typically as interface on models of
mechatronic subsystems or subsystems consisting purely of algorithms (software). For such
signals, the causality is normally prescribed already in the mathematical model and one have
to differ between differentiation orders, so that and are counted as two variables. A
strict way to implement this is given in the Modelica modelling format, see 1.5.4.5.
Only continuous dynamics, as opposed to discrete dynamics, were considered in 1.5.1.1.3 to 1.5.1.1.5.
Discrete dynamics modelling can be used for computational efficient models of, e.g., ideal dry friction
or ideal backlash but also to model time sampled (digital) systems. See Figure 1-14. The way the states
evolve over time differs between continuous and discrete dynamics: During the modelling, the contin-
uous states, 𝒙𝒄, should be thought of as changing continuously. In the computation there has to be a
discretization in time steps since computers are digital (not analogue). Hence the 𝒙𝒄 are updated only
between each time step but thought of as continuously changing in between. The update is based on
the state derivatives, 𝒙𝒄, and a derivative approximation, e.g. 𝒙𝒄 (𝒙𝒄 (𝑡 + Δ𝑡) 𝒙𝒄 )⁄Δ𝑡 . Discrete
states, 𝒙𝒅 , should be thought of as constant except that they stepwise change value at the time instants
when one of the transition conditions becomes true (not between these time instants!). These time in-
stants are called (state) events, and can be implemented as when event indicators, 𝒉(𝒙𝒄 𝒙𝒅 (𝑡) 𝑡) >
0, becomes true.
Solid show
𝑥 mathematically
correct solution.
Dashed show
computed solution,
Shows that 𝑥 can be here linear
reset during the event. interpolation over
time steps.
𝑥
Initial Event
states
ℎ 𝑡
All these time steps are …but this time step is adjusted to find the exact
constant or adopted to error time when the event’s transition condition
estimate in 𝑥 … becomes true, i.e. when ℎ becomes > 0.
Figure 1-14: A visualization of a solution with both continuous and discrete states.
At a first glance, the 𝑥 should not change in the events, but are two reasons why one sometimes model
them to change (stepwise) at an event: one is to reinitialize to physically obvious values (e.g. a bounc-
ing ball need to start from the surface it bounces at when the bounce happens) and the other is if one
change physical model at an event, e.g. introduce an inertia or elasticity. The latter can be powerful to
solve an engineering task, but the engineer has to be especially observant on how credible the numeri-
cal solution is, because the simulation tools can typically not estimate errors in such model. The latter
33
Introduction
way is here described as one model, but in another context it can be seen as “splicing” of several mod-
els.
A system which includes both types (hybrid dynamics) evolves as follows:
𝒙𝒄 (𝑡 + Δ𝑡) ← 𝒙𝒄 + Δ𝑡 ∙ 𝒙𝒄 𝒙𝒄 + Δ𝑡 ∙ 𝒇𝒄 (𝒙𝒄 𝒙𝒅 (𝑡) 𝑡) and
ℎ𝑒𝑛 𝒉(𝒙𝒄 𝒙𝒅 (𝑡) 𝑡) > 𝟎 𝑡ℎ𝑒𝑛 𝒙𝒅 (𝑡 + ) ← 𝒇𝒅 (𝒙𝒄 𝒙𝒅 (𝑡) 𝑡) .
Note that a “knee” (discontinuous derivative but continuous value) in an equation does generate an
event, but no discrete state, since no “memory” is needed. But a step (mathematical discontinuity in
value) or a hysteresis (mathematical function with multiple values depending on which branch is ac-
tive) generate state event with discrete state. See examples in 2.4.4.
One can see discrete dynamics as if the physics get stuck on the steep part of a step or in a branch in a
hysteresis, until when an exit condition becomes true. Examples where this occur is dry friction, one-
way clutches, back-lashes, diodes, etc. Another example appears if one changes physical models, e.g.
remove a mass or compliance, during some parts of the simulation, governed by certain events.
Discrete dynamics is not as well established in most basic engineering education as continuous dy-
namics.
1.5.1.4.1 § Event Iteration and Chattering
Several discrete state transitions can occur during same time instant, which calls for event iteration
one time instant, i.e. during zero time. An example is when a friction element switches state from posi-
tive slipping to sticking and to negative slipping during one time instant. A typical problem with these
models is that transition conditions are such that event iteration does not converge, so called chatter-
ing.
It can be needed to reinitialize continuous state variables (𝒙𝒄 (𝑡 + ) ⋯) in events, e.g. motivated by im-
pact dynamics. There is a risk that the numerical solution gets stuck in undesired high frequency
switching of discrete states which also can lead to chattering.
Chattering can be caused by coding errors of the event handling or inconsistent physical modelling.
1.5.1.4.2 § Finite State Machines
Examples of state events which typically appear in control algorithms or driver models are (finite)
state machines. These are easier in the way that typically do not have event iteration, but only one
state transition should happen in each time sample. This reduces the risk of chattering, or at least
makes the chattering to appear with the sample time, not during a time instant.
1.5.1.4.3 § Discrete Dynamics in Modelica Tools
Here is an example of how discrete states can be modelled (or implemented) in Modelica. The example
is purely mathematical and without physical interpretation.
Model example_DiscreteState
Real y, x_c (start=0);
Integer x_d (start=0);
equation
//Eqs defining continuous variables y and x_c (but can use the discrete variable):
y = x_c^2;
der(x_c) = if (x_d == 0) then -1 else +1;
//Eq defining discrete variable x_d (how it is updated at state events):
when (pre(x_d) == 1 and x_c > +1) then x_d = 0; reinit(x_c,+2);
elsewhen (pre(x_d) == 0 and x_c < -1) then x_d = 1; reinit(x_c,-2);
end when;
end example_DiscreteState;
34
Introduction
Figure 1-15: § Same model as shown in Modelica format, but here implemented in the Dataflow
Diagram Based Tool “Simulink”. The dashed arrow shows how to handle updates of a discrete
state 𝑥 .
+ 𝑓 + 0
If the guess is adjusted
𝑓 ℎ+ 𝐿⁄ 𝐿⁄ 0 until it gives same
𝑓
calculated we have
𝐿/ 𝐿/ 𝑓
∫ 𝑡 an iterative solution.
35
Introduction
2 𝑡
2 2 ∫ 2 2
1.5.1.7 Causality
Systems can be modelled with Natural causality. For mechanical systems, this is when forces on the
masses (or motion of the compliance’s ends) are prescribed as functions of time. Then the velocities of
the masses (or forces of the compliances) become state variables and have to be solved through time
integration. The opposite is called Inverse dynamics and means that velocities of masses (or forces of
compliances) are prescribed. For instance, the velocity of a mass can be prescribed and then the re-
quired forces on the mass can be calculated through time differentiation of the prescribed velocity. Cf.
Analysis and Inverse Analysis in Figure 1-7.
1.5.1.8 Drawing
Drawing is a very important tool for engineers to understand and explain. Very often, the drawing con-
tains free body diagrams, FBD, see 1.5.2.1, but also other diagrams are useful. On top of normal draw-
ing conventions for engineering drawings, it is also important to draw motion and forces. The notation
for this is proposed in Figure 1-18.
It is often necessary to include more than just speeds and forces in the drawings. In vehicle dynamics,
these could be: power flow and signal or data flows. These can preferably be drawn as arrows, but of
another kind than the motion and force arrows.
When connecting components with signal flow, the resulting diagram is a data flow diagram or block
diagrams. Physical components and physical connections can be included in such diagram, and if ar-
rows between them it would represent data flow or causality. Physical components can also be con-
nected by “Physical connections”, which does not have a direction, see 1.5.1.8.1. It should be noted that
a (computation) flow charts and (discrete) state diagram represent something quite different from data
flow diagrams, even if they may look similar; in state diagrams, an arrow between two blocks repre-
sents a transition from one (discrete) state or operation mode, to another.
36
Introduction
(“axis”)
- 𝑓 (vector)
𝐿3 (< 0) 𝐿2 (< 0) 𝐿 0
𝐿 (axis)
𝐿2
𝐿2
Figure 1-19: Different ways to define physical quantities. The length 𝐿2 is defined in 3 ways. Note
also different types of arrows: “axis”, “vector” and other, where other are ”meassurements”.
Defining forces and moments in a drawing is special in the sense that they always appear as action and
reaction, see 1.5.2.1.1.
37
Introduction
denotes multiplication between scalars and matrices equally, using “ ”. Multiplication between geo-
metric vectors are denoted: Cross multiplication, denoted × ⃗ and scalar multiplication, denoted •
⃗.
Parentheses shall be used to avoid ambiguity, e.g. ( ⁄ ) ∙ or ⁄( ∙ ), and not ⁄ ∙ .
An interval has a notation with double dots. Example: Interval between and is denoted . . .
An explanation, between two consecutively following steps in a derivation of equations is written
within {} brackets. Example: 𝑥 + 𝑦 { + + } 𝑦+𝑥.
An inverse function is denoted with superscript 𝑛 , 𝑦 𝑓(𝑥) ⇔ 𝑥 𝑓 𝑛𝑣 (𝑦). The use of superscript
can cause an ambiguity whether 𝑓 means inverse function or inverted function value ( ⁄𝑓).
When a name of the inverse function is available, it can be used, e.g. arctan(⋯ ) instead of taninv(⋯ ).
Fourier and Laplace transform of function 𝑓 (𝑡) are denoted ℱ(𝑓(𝑡)) and ℒ(𝑓(𝑡)), respectively:
+∞
𝜔
ℱ(𝑓(𝑡)) ∫ 𝑓 (𝑡) 𝑒 𝑡 ℎ𝑒𝑟𝑒 ∈ 𝑅𝑒
∞
∞
ℒ(𝑓(𝑡)) ∫ 𝑓(𝑡) 𝑒 𝑡 ℎ𝑒𝑟𝑒 𝑠 𝜎+𝑗 ℎ𝑒𝑟𝑒 𝜎 𝑛 ∈ 𝑅𝑒
There are many practical rules for manipulation transformed differential equations, such as
ℱ (𝑓(𝑡)) 𝑗 ℱ(𝑓(𝑡)) and ℒ (𝑓(𝑡)) 𝑠 ℒ(𝑓(𝑡)) 𝑓(0) . For Laplace transform, more examples
are given in Figure 1-20.
Function Function in time
Function in Laplace domain, ℒ(𝑓(𝑡))
description domain, 𝑓(𝑡)
Unit impulse (𝑡) ℒ( (𝑡))
Unit step 𝜎(𝑡) ℒ(𝜎(𝑡)) ⁄𝑠
Unit ramp 𝑡 𝜎(𝑡) ℒ(𝑡 𝜎(𝑡)) ⁄𝑠 2
Exponential 𝑒 𝛼 ℒ(𝑒 𝛼 ) ⁄(𝑠 + )
Sine sin( 𝑡) ℒ (sin( 𝑡)) ⁄(𝑠 2 + 2 )
Cosine cos( 𝑡) ℒ (cos( 𝑡)) 𝑠⁄(𝑠 2 + 2 )
Figure 1-20: Examples of function in time domain and the corresponding in Laplace domain.
An example: The dynamic model 𝑦 (𝑦(𝑡)) 𝑛 𝑢(𝑡) looks as follows, if written Fourier trans-
formed or Laplace transformed: 𝑦( ) 𝑛 𝑢( )⁄( 𝑗 ) or 𝑦(𝑠) 𝑛 𝑢(𝑠)⁄𝑠 , respectively.
The order within subscripts can be debated, e.g. should a longitudinal (𝑥) force ( ) on rear axle (𝑟)
should be denoted or . The intention in this compendium is to order subscript with the physical
vehicle part (here rear axle) as 1st subscript and the specification (longitudinal, x) as 2 nd, leading to .
If there are further detailed specifications, such as coordinate system, e.g. wheel coordinate ( ), it will
be the 3rd subscript, leading to 𝑤 . Further additional specifications, such as case, e.g. without (0) or
with (1) a certain technical solution, will be the 4 th subscript, leading to 𝑤 and 𝑤 . If subscripts
have >1 token, it can sometimes be good to use comma, e.g. 𝑤 𝑙𝑤 .
Naming of variables (signals), parameters and function components (blocks) is often important in
large products as vehicles. Some naming needs to be readable for many engineers, such as CAN signals.
Hence, companies developing their own naming standards. But also, the standardization organisation
AUTOSAR has released a naming standard, with intention to be accepted by both vehicle manufactur-
ers and system suppliers.
38
Introduction
39
Introduction
40
Introduction
𝑥 𝑥
𝑥
𝑥 𝑝
𝑥 𝑥 𝑥 𝑥
+ + 𝑝 +
⇒
Single state ODE: Single state ODE:
̈+ + + ⇒ ⇒ 𝑝+
⇒ 𝑥̈ + 𝑥+ 𝑥 ̈+ 𝑝+ + + +
Single state ODE:
𝑥 + 𝑥 + 𝑜𝑛𝑠𝑡 𝑛𝑡 𝑝
̈+ + 𝑝 + + + +
Figure 1-21: § Examples of damped mass and spring systems, as 1st order ODEs and single state
ODEs.
Equilibrium gives relations between forces (including moments). For a static system, we can use the
(static) equilibria: Sum of forces in any direction is zero: ∑ 𝑭 𝟎 and Sum of moments around any
axis is zero: ∑ 𝑴 𝟎
In a dynamic system, there is typically inertia effects. Inertia effects can be seen as “fictive forces” and
“fictive moments” (or “d’Alembert forces or moments”). Typically, the fictive forces are 𝑓
(where translational acceleration) and the fictive moments are 𝑀𝑓 . These are counter-
directed to and , respectively. We can find the equations either as “Dynamic equilibria”:
• Sum of forces, including 𝑓 , in any direction is zero: ∑𝒊𝒏𝒄𝒍.𝒇𝒊𝒄𝒕 𝑭 𝟎
• Sum of moments around any axis, including both 𝑀𝑓 and “ 𝑓 𝑒 𝑒𝑟”, is zero:
∑𝒊𝒏𝒄𝒍.𝒇𝒊𝒄𝒕 𝑴 𝟎
or as “Equations of motion” or “Newton’s 2nd law”:
• Sum without fictive in direction have to be equal to : ∑𝒆𝒙𝒄𝒍.𝒇𝒊𝒄𝒕 𝑭 𝒂
• Sum without fictive around axis through CoG in direction of have to be equal to :
∑𝑪𝒐𝑮 𝒆𝒙𝒄𝒍.𝒇𝒊𝒄𝒕 𝑴 𝑱𝑪𝒐𝑮 𝝎 + 𝝎 × (𝑱𝑪𝒐𝑮 𝝎)
The is the mass moment of inertia matrix in CoG along the same coordinate axes as is expressed
in. In this compendium, the alternative with Dynamic equilibria is mainly used. The fictive forces are
introduced in the free-body diagrams with dashed arrows, see Figure 1-22 and Figure 1-23. Ad-
vantages with Dynamic equilibria, as opposed to Newton’s 2 nd law, are easiness to form moment equi-
libria since it can be taken around any axis (example in Figure 1-23). It is also easier to include motion
effects within a part of the system (such as Jf ωf and in Figure 1-22). The general form for the
fictive forces is:
41
Introduction
𝑓 (𝑡𝑟 𝑛𝑠 𝑡 𝑜𝑛 𝑜 𝑒𝑛𝑡𝑢 ) ( ) ( 𝑟)
𝑡 𝑡 𝑡 𝑡 𝑡
𝒗 ⏟ 𝝎×𝒗 ⏟ 𝝎 × (𝝎 × 𝒓) ⏟ 𝝎×𝒓
𝑙 𝑓 𝑛 𝑓 𝑙𝑓 𝐸 𝑙 𝑓
The 𝑟 is the position vector. The Coriolis and Euler forces can often be assumed as zero in examples in
this compendium due to no motion within the vehicle and rigidity of rotating bodies. The general form
for the fictive moments is:
⃗⃗ 𝑓
𝑀 (𝑟𝑜𝑡 𝑡 𝑜𝑛 𝑜 𝑒𝑛𝑡𝑢 ) (𝑱 ⃗ ) 𝑱𝑪𝒐𝑮 𝝎 + 𝝎 × (𝑱𝑪𝒐𝑮 𝝎)
𝑡 𝑡 𝑪𝒐𝑮
The term 𝝎 × (𝑱𝑪𝒐𝑮 𝝎) can often be assumed as zero in examples in this compendium due to symme-
tries and rotation in one plane at the time.
Figure 1-22 and Figure 1-23 show examples of free-body diagrams (FBDs) with all forces, including
fictive forces. With such FBDs, the equilibrium equations are implicitly defined.
In Figure 1-22, the mass of the wheels can be included in the FBD of the vehicle body and rear axle, but
not the FBD of the front axle. The moment of inertia is marked as , but in practice we can often
neglect these terms in the FBD of the whole vehicle (but not for the FBD of the wheel, where they are
often essential in the equilibrium 𝑇 𝑊ℎ𝑒𝑒 𝑅 𝑢𝑠 ), see more in 3.3.5.2.
acceleration,
𝑓 𝑓
𝑓 𝑓 𝑇𝑓 𝑓
Selecting:
Selecting: • vehicle body together 𝑓 𝑓
𝑓 • whole vehicle body with rear axle 𝑓
𝑓
with both axles • front axle alone 𝑓
Figure 1-22: Free body diagram. The dashed arrows marks “fictive” force or moment. The star is a
way to mark around which point(s) moment equilibrium is taken in the later stage “Mathematical
model”. The fictive moments 𝑓 𝑓 and are often neglectable in equilibrium for whole
vehicle, but significant in equilibrium for individual wheels.
In Figure 1-23, the prescribed motion actuation (steering angle (𝑡)) and good road grip result in 4
states, e.g. [ 𝑀 ] or [ 𝑀 ]. If unsprung parts have only a neglectable portion of
the total vehicle mass, it can be easier to use 0 and replace with the total vehicle mass. How-
ever, should not be changed, because it is still only the sprung body that rolls.
42
Introduction
Free body diagram of Equilibrium for “sprung body”. Method with with fictive
“sprung body”: forces, e.g.:
Vehicle seen • Lateral: 0
from rear: • Vertical: + 0
• Roll around any point, e.g. the star ( ):
𝑀 + Δℎ + Δℎ 0
Alternative method, without fictive forces. Equally correct, but it
𝑪𝒐𝑮 ( ) only works with roll equilibrium around 𝑜 :
• Lateral in direction:
( ) 𝑀 • Vertical in direction:
• Roll around 𝑜 : 𝑀 + Δℎ + Δℎ
Δℎ
Equilibrium for “unsprung parts”:
• Lateral: + 0
ℎ
RC
• (Vertical and Roll could have been included. However, here, we
( ) 𝑀 do not aim at solving for and 𝑀 .)
ℎ
Compatibility:
ℎ
• ≈0 Δℎ +
ℎ ℎ +
Free body diagram of 𝑀 Constitution for suspension spring:
“unsprung parts” (axles): • 𝑀
Constitution for tyre-to-ground ( is steering angle):
•
Eqs (or dependencies) between parameters: Actuation (steering):
+ • 𝑓 𝑡 (𝑓 =known function of time 𝑡 )
ℎ ℎ + ℎ ℎ 8 eqs, 8 unknown variables ( 𝑀)
Figure 1-23: Free body diagrams for combined translation and rotation. The figure also shows the
two ways of setting up equilibria, with and without fictive forces. Angle is assumed small. The
is written as a general damping coefficient, while, in 2.2.4, it can be found that ⁄
§ A Modelica implementation of the model in Figure 1-23 is here shown:
//Constitution: -6E4
0 2 4 6 8 10
-6E4
0 2 4 6 8 10
der(M_s)=-c*w_sx; phi_sx
phi_x phi_sx
phi_x
F_y=-d*(v_uy-v_x*delta); 5 5
end RollDynamics_Sprung_Unpsrung; 0 0
-5 -5
0 2 4 6 8 10 0 2 4 6 8 10
Figure 1-24: § A Modelica implementation of the model in Figure 1-23 and two simulation results.
§ Centrifugal Forces in General 3D Motion
Centrifugal force terms appear as where ≠ 𝑗, i.e. for rotation perpendicular to translational
velocity:
0 + 0 +
[ ] [ ]+[ 0 + ] [ ] [ ] + [+ 0 ] [ ]
+ 0 + 0
43
Introduction
+ – 𝑛 𝑜𝑠𝑡 𝑟𝑒 𝑒 𝑛𝑡 0
[ + ] ≈ { 𝑒ℎ 𝑒 𝑜𝑝𝑒𝑟 𝑡 𝑜𝑛𝑠: } ≈ [ ] + [+ ]
+ ≫ 𝑛 ≫
In most vehicle operations, the most important centripetal acceleration terms are: ∙ (see Eq
[4.45]) and ∙ (see Eq [3.24]).
§ Force Equivalence and Coordinate Transformation
Special cases of equilibrium equations are coordinate transformation between force components, e.g.
Eq [1.3]. Another is force equivalence. Note that it is not the same as force equilibrium. Force equiva-
lence is used e.g. in Figure 2-43 and Figure 2-97. A set of forces are equivalent with some other set if
they can be replacing each other in the free body diagram. It is recommended to not draw the two set
of forces in the same free body diagram, since they are alternative to each other.
𝑓 𝑤
Axle forces in 𝑓 𝑤
Wheel directions,
𝑭𝒇𝒙𝒘 𝑭𝒇𝒚𝒘 𝑭𝒓𝒙 𝑭𝒓𝒚
𝑓
𝑤
𝑀𝑤
𝑀 𝑛
Summed Vehicle forces including
Inertial forces, 𝑭𝒙𝑰𝒏 𝑭𝒚𝑰𝒏 𝑴𝒛𝑰𝒏 𝑛
44
Introduction
• For a dry friction contact: 𝑜𝑟 𝑒 𝑜𝑛𝑠𝑡 𝑛𝑡 𝑠 𝑛(𝑆 𝑛 𝑆𝑝𝑒𝑒 ) . This is the most common
friction model in mechanical engineering, explained by adhesion between molecules and hys-
teresis when material is deforming over micro level asperities. So, the proportionality constant
depends on both cooperating bodies material and surface roughness. Note: When a friction
contact sticks, the equation switches to a compatibility equation (𝑆 𝑛 𝑆𝑝𝑒𝑒 0 ).
• For a more general model component: 𝑜𝑟 𝑒 𝑓𝑢𝑛 𝑡 𝑜𝑛( 𝑜𝑠 𝑡 𝑜𝑛 𝑆𝑝𝑒𝑒 ) . Even more gen-
eral would include models of actuators: 𝑜𝑟 𝑒 𝑓𝑢𝑛 𝑡 𝑜𝑛( 𝑜𝑠 𝑡 𝑜𝑛 𝑆𝑝𝑒𝑒 𝑆 𝑛 ) , where
𝑆 𝑛 often is a request signal, e.g. 𝑜𝑟 𝑒𝑅𝑒𝑞𝑢𝑒𝑠𝑡.
1.5.2.3.4 Algorithms and Other Equation Types
The listing of equation types in 1.5.2.3.1..1.5.2.3.2 is a help to model but it is not claimed to be com-
plete. There are many other equation types that can appear, e.g. from electrical and chemical science.
Among these others there are two sub-types (control algorithms and driver models) that are espe-
cially important for vehicle dynamics, so they will be discussed here in 1.5.2.3.4.
Via sensors and actuators, control algorithms can operate with the mechanics, mechatronics. The con-
trol algorithms with their interface to sensors and actuators is here included in the equation type “al-
gorithms”. We also include models of how the human driver controls and experience the mechanical
quantities. This equation type cannot be sorted into the traditional 1.5.2.3.1..1.5.2.3.1.2. Conceptually,
any quantities that can be sensed or actuated in, and outside of, the subject vehicle can occur in these
equations. (Finite) State machines are often useful when modelling (and designing) algorithm-based
functions but also the driver, see discrete state machine in 1.5.1.4.
A model of control algorithm can often be the same artefact as the design of it, especially if using a
modelling tool that allows automatic generation of real-time code, like Simulink. However, note that
the algorithms in real system is implemented in a time discrete digital computation platform and digi-
tal communication, so using a time continuous version as model is an approximation in itself. For best
fidelity, the models need to be formulated as time discrete dynamics. Then one can properly represent
the influence from the design parameter sample time on the vehicle functions.
1.5.2.3.5 § Lagrange Mechanics
An alternative to Newton mechanics is Lagrange mechanics, which is sometimes an easier method to
find the same equations. Lagrange mechanics is basically to express kinetic energy (𝑇) and potential
energy (𝑉) in (position) coordinates, Cartesian or Generlized. With Lagrange mechanics, the equilib-
rium is not formulated in forces (or moments) but in partial derivatives of energy (kinetic energy sur-
plus, called the Lagrangian 𝐿 𝑇 𝑉) with respect to these coordinates and their derivatives. One
can see this as “equilibria of projected energy”. Lagrange mechanics does not use the factorisation of
power into , so it undermines the categorization in 1.5.2.3.1..1.5.2.3.1.2; the compatibility and
constitution will be mixed in from start. So, the modularity corresponding to physical components in
the system often gets lost. Also, energy generating and dissipating components has to be treated with
special care, especially dry friction. Next figure shows an example of how the Lagrange method can be
applied on a simple system. The system has no external forces, i.e. it has no non-constraint forces. It
neither has energy exchange nor energy dissipation inside itself. Hence, the energy in the system is
constant. The example is the same as the one which Newton method is applied on in Figure 1-38.
Δ𝑥
Δ𝑥
𝑥 𝑥2
2
𝑥 𝑥2
Lagrange 1st kind Lagrange 2nd kind
45
Introduction
𝑛 ∑ ( 𝑠 ) 𝒗𝑻 ⏟
diag([𝑠 𝑠2 ⋯ 𝑠𝑁 ]) 𝑭 0 ⇒
= ⋯𝑁 =diag([ 𝒗 𝑭])
⇒ [𝒗𝒗 𝒗𝑻𝑭 ] [( 𝒗 𝑭𝒗 )𝑻
𝑻 ( 𝑭 𝑭𝑭 )𝑻 ]𝑻 0
⇒ [𝒗𝑻𝒗 (𝑹𝒗 𝒗𝒗 )𝑻 ] [( 𝒗 𝑹𝑭 𝑭𝑭 )𝑻 ( 𝑭 𝑭𝑭 )𝑻 ]𝑻 0
⇒ 𝒗𝑻𝒗 ( 𝒗 𝑹𝑭 + 𝑹𝑻𝒗 𝑭 ) 𝑭𝑭 0 ⇒ 𝒗 𝑹𝑭 𝑹𝑻𝒗 𝑭
𝑭 or
𝑻 𝑇
⇒ 𝑹𝑭 𝒗 𝑹𝒗 𝑹𝒗 𝑭 𝑹𝑭 𝒗
46
Introduction
For a transmission, such as a planetary gearbox with many shafts, it can often be most natural to de-
rive the 𝑹𝒗 from compatibility and 𝑹𝑭 from 𝑹𝑭 𝑻
𝒗 𝑹𝒗 𝑭. For a suspension linkage, it can often
be most natural to derive the 𝑹𝑭 from equilibrium and 𝑹𝒗 from 𝑹𝒗 𝑭 𝑹𝑭
𝑇
𝒗 . Note that, for a
suspension, the proportional relations expressed with 𝑹𝒗 and 𝑹𝑭 , are only valid close around a certain
position or pose.
If the system studied has energy storage or power dissipation included, they can often be defined as
outside the mechanism, only leading to more interfaces. However, if the energy dissipation is of dry
friction type, this is not easily possible. Instead, the geometry can be adjusted via a geometry interpre-
tation of dry friction, using a friction angle which defines a corrected geometry, see (Mägi, Jacobson, &
Resev, 1998).
Vehicle
Vehicle Level Algorithms
requests on requests on
Environment Driver Vehicle
vehicle, 𝒚𝒓𝒆𝒒 actuators, 𝒓𝒆𝒒 Actuated and
(road and Driver Interp- Motion
traffic) reter
Arbitra- Controller Sensing Vehicle
Virtual tor information,
𝒂𝒄𝒕 𝒂𝒍 𝒄𝒂𝒑𝒂𝒃𝒊𝒍𝒊𝒕𝒚
Driver
Estimator
and Sensor
information Fusion
(Open loop) Vehicle Motion (Closed loop) Vehicle Motion Controller (of gain type)
Controller
𝒚𝒓𝒆𝒒 + request for request,
Algorithm
𝒚𝒓𝒆𝒒 Algorithm 𝒓𝒆𝒒 - increase, 𝒚𝒓𝒆𝒒 Δ𝑡 𝒚𝒓𝒆𝒒 𝒓𝒆𝒒
information information
except 𝒚𝒂𝒄𝒕 𝒂𝒍
𝒚𝒂𝒄𝒕 𝒂𝒍 except 𝒚𝒂𝒄𝒕 𝒂𝒍
Figure 1-27: Context for Vehicle Level Algorithms and Vehicle (Motion) Controller. “Virtual Driver”
takes vehicle driver into account and closes the loop in vehicle speed and lateral position.
Control algorithms can be designed without utilizing knowledge about the controlled system, such as
neural networks, tuned only on observations on how the system responds. However, in this section we
only consider Model based controllers. For those, the input and output signals, as well as parameters
inside, has a clear physical interpretation in vehicle quantities, as well as units, see Figure 1-28. For
instance, the requests on vehicle, 𝒚𝒓𝒆𝒒 , is typically forces or acceleration, [ [𝑁] [𝑁] 𝑀 [𝑁 ]] or
[ [ ⁄𝑠 2 ] [ ⁄𝑠 2 ] [𝑟 ⁄𝑠 2 ]] or velocities [ [ ⁄𝑠] [ ⁄𝑠] [𝑟 ⁄𝑠]]. The model base is
helpful when tuning the controller parameters. Moreover, the model base helps to find a consistency
47
Introduction
between derivatives order in the controller and the controlled system, such as if P, I or D gains should
be used in a PID-controller.
The 𝒚𝑹𝒆𝒒 contains typically vehicle motion variables (positions, velocities, accelerations, forces) but
other vehicle variables, such as energy or power variables can be added, e.g. 𝒚𝑹𝒆𝒒 [ 𝑀 𝑓𝑓 ],
where 𝑓𝑓 is energy level (or state of charge 𝑆𝑜 ) in an energy buffer.
⋯
Vehicle
Motion
𝑣 𝑣
Control e.g. ←
e.g. ← sin
𝑓𝑖
𝑇 𝑝 +𝑇 𝑘
Actuator Coordinator 2
𝑇𝑙 𝑘
𝑤
𝑙 𝑟𝑓 𝑛 𝑙 𝑇 𝑝
𝑤 𝑅𝑤
𝑇 𝑝 𝑇𝑙 𝑘
𝑇 𝑇 𝑘
𝑘
𝑒
Actuators +
𝑙
Prop Brk
Figure 1-28: Request flow in Vehicle Motion Control. Each block has its own Physical model, which
explains the its interface quantities (variables and parameters).
⋯ 2 2
Motion Model
2 2 2
Vehicle
Motion
2 2
Control
2 Force Model 1
2 2 2 2
2 2
2
Motion quantities inherited from Motion Model
but not drawn. Not drawn forces and moments:
𝑀 𝑀 From air on bodies: 𝑀 𝑀 2
2
From ground on bodies: 2
From tyre on each unit : 𝑀 𝑀 2
Force Model 2
Axel Group Coordinator For both units, but only drawn for unit 2.
Force and
moment
𝑀 equivalence
𝑓
⋯ 2
⋯ 𝑀 𝑀 𝑀
2 2𝑓 2
Actuation Models
Actuator Actuator Actuator
Not drawn, but they can typically
Coordinator Coordinator Coordinator
be drawn per axle group.
Figure 1-29: § Similar to Figure 1-28, but for a more complex vehicle.
48
Introduction
49
Introduction
50
Introduction
matrix. Optimizing in Least Square sense (a.k.a. using quasi-inverse) gives 𝒓𝒆𝒒
((𝑾𝒗 𝑩)𝑇 𝑾𝒗 𝑩) (𝑾𝒗 𝑩)𝑇 𝑾𝒗 𝒗𝒓𝒆𝒒 (𝑾𝒗 𝑩)" " 𝑾𝒗 𝒗𝒓𝒆𝒒 .
• If dim( 𝒓𝒆𝒒 ) > rank(𝑩), the vehicle is (effectively) Over-actuated ( 𝒓𝒆𝒒 is under-determined).
Conceptually, there is no way to convert the problem to a Neutral actuated case (without disre-
garding actuators or adding requests), but one can find and approximate solution through opti-
mization. Trying an optimization as for the under-actuated vehicle will not work well, since
there is generally no unique solutions, i.e. ((𝑾𝒗 𝑩)𝑇 𝑾𝒗 𝑩) is singular). One way to handle
it would be to use some random way to find one 𝒓𝒆𝒒 among all non-unique optima. This can
lead to jumping 𝒓𝒆𝒒 in time, which is undesirable. A second way is to convert the problem to a
Neutral actuated case by simply prescribing suitably many components in 𝒓𝒆𝒒 : 𝑢 𝑢 ,
where subscript 𝑒𝑠 means desired. A third way is to to convert the problem into a Neutral- or
Under-actuated case by adding equations for each and all components in 𝒓𝒆𝒒 , typically: 𝒓𝒆𝒒
𝟎. To use zero is difficult to motivate directly from vehicle motion/actuation perspective but it
can be thought of as a way to guaranteeing unique mathematical solutions, which is good.
Combining second and third way gives an optimization problem with cost function as follows:
𝑓 ‖𝑾 𝒐𝒕𝒊𝒐𝒏 (𝒗𝒓𝒆𝒒 𝑩 𝒓𝒆𝒒 )‖ + ∑‖𝑊 (𝑢 𝑢 )‖ + ‖𝑾 𝒐𝒐𝒕𝒉 𝒓𝒆𝒒 ‖ .
2 2 2
The 𝑾 matrices are diagonal matrices with weight factors. For solving such mathematical
problems there are well established subroutines for Control allocation, CA, which can also
take constraints into account.
A visualisation of these cases and solution is given in Figure 1-30.
Under-actuated vehicle: 𝑩
𝑒 𝑓
dim=3 dim=3 𝒓𝒆𝒒 𝑩" " 𝒗𝒓𝒆𝒒 or dim=2
𝒗𝒓𝒆𝒒 𝒚𝒓𝒆𝒒 Actuated vehicle:
𝒚𝒓𝒆𝒒 𝒗𝒓𝒆𝒒 𝒓𝒆𝒒
𝒈 𝑩 𝟎 𝑢𝒓𝒆𝒒 arg min 𝑜𝑠𝑡 𝒓𝒆𝒒
𝒇𝑫𝑨𝑬 𝒙 𝒓𝒆𝒒 𝒚𝑡 𝟎
𝒓𝒆𝒒
Over-actuated vehicle: 𝑩
𝑒 𝑓
" " 𝒗𝒓𝒆𝒒
𝑩
dim=2 dim=2 dim=3
𝒗𝒓𝒆𝒒 𝒚𝒓𝒆𝒒 𝒓𝒆𝒒 𝑢 0 or Actuated vehicle:
𝒚𝒓𝒆𝒒 𝒗𝒓𝒆𝒒 𝑰 𝟎 𝒓𝒆𝒒
𝒈 𝑩 𝟎 𝒇𝑫𝑨𝑬 𝒙 𝒓𝒆𝒒 𝒚𝑡 𝟎
𝒓𝒆𝒒 arg min 𝑜𝑠𝑡 𝒓𝒆𝒒
𝒓𝒆𝒒
Figure 1-30: § Visualisation of possible conceptual controllers for vehicles with different degree of
over-actuation. Superscript ⋯" " denotes “quasi-inverse”. Weigthing is not shown.
1.5.3.4.1 Momentaneous Optimization
Note that the weigthing is not shown in Figure 1-30. The weighting is probably the most difficult, in
some sense, one have to compare costs for different motion components, so if simple weigth matrix, its
components should typically have the units [€⁄𝑁] in longitudinal and [€⁄𝑁] in lateral which is very
difficult to set and might vary with such as distance to lead vehicle and distance to road edge. Also, if
𝑦 includes energy related requests, the units can be of type [€⁄ ] depending on the fuel or electric
energy price. So, the approach described above is very mathematical and a bit naive when we see the
original engineering problem, see 1.5.1.1.5.3.
Instead of a naive weigthing “outside” the model, we can include the cost as a variable already in our
model. The cost will then appear more intuitive and less artificial. Typically, one see two types of cost
variables: 𝑜𝑠𝑡 in [€] and 𝑜𝑠𝑡𝑅 𝑡𝑒 in [€/𝑠]. For transport costs (fuel, driver salary, etc) these are often
51
Introduction
very natural. For motion they are still very difficult to grasp, which suggest to look for ways to see mo-
tion requests as constraints (to fulfil exactly) and transport costs as costs (to be minimized).
Looking only at transport costs, one realizes that the overall goal is to minimize the 𝑜𝑠𝑡 [€], but that
has to be done predictive, i.e. by integrating 𝑜𝑠𝑡𝑅 𝑡𝑒 [€/𝑠] over a longer time, see 1.5.3.2 and
1.5.3.4.2. However, if the motion constraints should be fulfilled as constraints over the time, it is prob-
ably a good idea to minimize 𝑜𝑠𝑡𝑅 𝑡𝑒 [€/𝑠] momentanously (at each time instant) such that the mo-
mentaneous motion constraints are fulfilled.
An example of such optimization is given here: A vehicle with two propulsion motors, subsrcitps 1 and
2, is modelled as:
𝑟 𝑡 𝑜 (𝑇 + 𝑇2 )
2 𝑟 𝑡𝑜
𝑓𝑢𝑒 𝑅 𝑡𝑒 𝑓 (𝑇 ) + 𝑓2 (𝑇2 2 )
𝑜𝑠𝑡𝑅 𝑡𝑒 𝑓𝑢𝑒 𝑅 𝑡𝑒 𝑓𝑢𝑒 𝑟 𝑒
The 𝒗𝑹𝒆𝒒 [ ] and 𝑹𝒆𝒒 [𝑇 𝑇2 ].
Minimizing 𝑜𝑠𝑡𝑅 𝑡𝑒 under constraint of fulfilling motion request :
{𝑇 𝑇2 } arg (min( 𝑜𝑠𝑡𝑅 𝑡𝑒(𝑇 𝑇2 )) 𝑠𝑢 𝑗𝑒 𝑡 𝑡𝑜 𝑟 𝑡 𝑜 (𝑇 + 𝑇2 ) )
𝑇1 𝑇2
where 𝑜𝑠𝑡𝑅 𝑡𝑒(𝑇 𝑇2 ) (𝑓 (𝑇 𝑟 𝑡 𝑜 ) + 𝑓2 (𝑇2 𝑟 𝑡 𝑜 )) 𝑓𝑢𝑒 𝑟 𝑒
There are methods to find the optimum for certain forms of the 𝑜𝑠𝑡𝑅 𝑡𝑒 and the constraint. One ex-
ample is quadratic programming (e.g. Matlab command quadprog), which assumes that 𝑜𝑠𝑡𝑅 𝑡𝑒 can
be formulated on a quadratic form:
𝑻
𝑜𝑠𝑡𝑅 𝑡𝑒 𝑹𝒆𝒒 𝑸 𝑹𝒆𝒒 + 𝒇𝑻 𝑹𝒆𝒒 𝑠𝑢 𝑗𝑒 𝑡 𝑡𝑜 𝑩 𝑹𝒆𝒒 𝒗𝑹𝒆𝒒
This form includes both linear and quadratic terms, so coefficient can often be found which makes a
rather good approximation.
1.5.3.4.2 Momentaneous and Predictive Optimization
The above in the 1.5.3.4 could be applied also for predictive control if we think of 𝒓𝒆𝒒 and 𝒚𝒓𝒆𝒒 as hav-
ing 𝑁 times more elements, where 𝑁 is the number of time instants considered in the prediction. The
model can be applied 𝑁 times, and if it contains state variable, a time integration method is needed to
make the sequences of each variable consistent. If model is linear and optimization is to be applied,
there are cooking-book methods, such as Model Predictive Control (MPC), see e.g. Ref (Ross,2015),
describes MPC. Simply adding more time instants adds equally many equations as unknowns ( 𝒓𝒆𝒒 ), so
it does not change the degree of over-actuation. However, if the variation of 𝒓𝒆𝒒 (𝑡) is parameterised
(e.g. 𝒓𝒆𝒒 (𝑡) 𝒑 or 𝒓𝒆𝒒 (𝑡) 𝒑𝟏 + 𝒑𝟐 𝑡 ), the problems becomes as under-actuated, which is more
well-conditioned and can be solved in a least square sense.
Over actuated passenger car can appear if each wheel has controllable actuation, see (Jonasson, 2009).
52
Introduction
(Whether a physical quantity is a variable or a parameter is cannot be uniquely decided in the real
world.) Generally speaking, variable estimation is more typically on-line filtering while parameter esti-
mation can have more of off-line character. Off-line can log several data samples and then find (the
constant) parameter value, best fitted for the whole sample period, e.g. in least square sense. Parame-
ter estimation can also be recursive, which means that each new sample is weighted with previous pa-
rameter estimate; so the estimate is a memory, i.e. a state.
Figure shows an example of an on-line filter is a first order filter. Figure also shows a simple differenti-
ator. Figure shows both continuous and discrete implementation in a block diagram format. Imple-
mentation of discrete integration using when statements and the pre() operator in Modelica was
shown in 1.5.1.4.
Continuous:
≈ ∫ 𝑡
+ 𝑡 𝑓𝑓 𝑡 𝑓𝑓
∫
𝑡 𝑓𝑓
Smaller time delay 𝑡 𝑓𝑓 gives more
correct differentiation and less
(Time) discrete: delayed . But, it also requires more
of the continuous time integration
+ 𝑡 𝑚𝑝𝑙 (smaller steps or higher order), or
𝑡 𝑓𝑓 smaller sample time 𝑡 𝑚𝑝𝑙 in the
discrete time integration.
Figure 1-31: § Example of differentiator (output ) and on-line filter (output y) as block diagram.
Upper: Continuous. Lower: (Time) Discrete.
In dynamic models for simulation, one should avoid differentiation because it introduces delays which
can lead to numerically unstable or wrong solutions. Off-line differentiation is much less troublesome.
However, if the derivative is not fed back, the problems become isolated with less risk for troubles.
Physical-model-based variable estimators (or State Observers) are often basically designed as simula-
tion of a model driven by the sensed inputs 𝒆𝒏 .
Physical model: Mathematical model: 𝑇
𝑇 ←
0 𝑇 𝑅 𝑅
𝑇
Explicit form model (𝒇 𝒈) :
𝑇 𝑛 + ∫
← 𝑡 𝑓𝑓
𝑅
(𝑛𝑜 𝑠𝑡 𝑡𝑒 𝑝𝑢𝑟𝑒 𝑦 𝑒 𝑟 )
Differentiator
To align both physical quantities (𝑇 and ) to same real time
instant, it is recommended to filter 𝑇 with same delay 𝑡 𝑓𝑓 .
53
Introduction
Treqveh vehicle_OneWheel_Act.w fxEst_Diff_Cont.w _est fxEst
model FxEst_Diff_Cont 70
parameter Real DifferentiationTime=0.1; 0E0 step in brake torque request
60
input Real w; -1E4
Real w_est, T_est, Fx_est; 0 1 2
50
input Real T, R, J; vehicle_OneWheel_Act.mustick vehicle_OneWheel_Act.muslip
equation 2 40
der(w_est) = (w - w_est)/DifferentiationTime; step up in friction
der(T_est) = (T - T_est)/DifferentiationTime; 30 wheel speed
0 • real
0 = T_est - Fx_est*R - J*der(w_est); 0 1 2
end FxEst_Diff_Cont; vehicle_OneWheel_Act.vx Rw
20 • continuously estimated
• discretely estimated
20 10
model FxEst_Diff_Discr vehicle and wheel speed
parameter Real SampleTime=0.05; 0
18
parameter Real DifferentiationTime=0.1; 0 1 2
input Real w;
16
Real w_est, T_est, Fx_est; vehicle_OneWheel_Act.T fxEst_Diff_Cont.T_est fxEst_D
input Real T, R, J; 0
14
Real derw_est, derT_est; wheel speed
equation • real
12 -500
when sample(-1000, SampleTime) then 0 1 2 • continuously estimated
derw_est = (w-pre(w_est))/ DifferentiationTime; vehicle_OneWheel_Act.sx • discretely estimated
w_est=pre(w_est)+ -1000
((pre(derw_est)+derw_est)/2)*SampleTime; 0.00
wheel slip
derT_est = (T-pre(T_est))/DifferentiationTime; -0.05 -1500
T_est=pre(T_est)+
((pre(derT_est)+derT_est)/2)*SampleTime; -0.10 -2000
end when;
0 = T_est - Fx_est*R - J*derw_est;
end FxEst_Diff_Discr; -0.15 -2500
0 1 2 0 1 2
Figure 1-33: § Continuous and discrete implementation of estimator in Figure 1-32. Tested with a
one-wheel vehicle model with ideal, but sampled, sensing of wheel speed and wheel brake torque.
Implicit midpoint or Trapezoidal rule (or Tustin approximation) used as derivative approximation
for the discrete integration.
1.5.4 Tools
This section presents some tools and methods for modelling and computation.
54
Introduction
F = (a*w)/(s^2 + w^2)
f = a*sin(t*w)
>> partfrac(f,x) %is the partial fraction decomposition of the expression f
>> clear all; syms f x; f=(7*x+3)/(5*x^2-3*x+0.5); partfrac(f,x)
ans = (14*x + 6)/(10*x^2 - 6*x + 1)
>> clear all; syms f x; f=(7*x+3)/(5*x^2-3*x+0.5); partfrac(f,x,'Factor-
mode','complex')
ans = (0.7 - 5.1i)/(x - 0.3 - 0.1i) + (0.7 + 5.1i)/(x - 0.3 + 0.1i)
>> ABCD=ss(A,B,C,D); %creates an object ABCD representing the linear state
space model “A,B,C,D”, see 1.5.1.1.3.2
>> step(ABCD) %computes step response of the model ABCD, see Figure 1-34.
Of special interest for dynamic systems is that Matlab has a built-in function for ”matrix exponen-
tial”, mentioned in 1.5.1.1.5. E.g., if 𝒙 𝑨 𝒙 with 𝒙(0) 𝒙𝒊𝒗 the solution is 𝒙(𝑡) 𝑒𝑨 𝒙𝒊𝒗 which can
simply be computed as:
>> x=expm(A*t)*x_iv; %with A as (square) matrix
m=1500;
f_heave=1; c=round(m*(2*pi*f_heave)^2,-3)
damping=0.3; d=round(damping*(2*sqrt(c*m)),-2)
figure(1), clf
%derx=A*x; x=[v_z;F_s;z];
%x=expm(A*t)*x_iv;
A=inv(diag([m,1,1]))*[-d 1 0; -c 0 0; 1 0 0];
x_iv=[0;1000;0];
for N=[7,100]
t_vec=[linspace(0,2,N)];
x_mat=[];
for i=1:length(t_vec)
t=t_vec(i);
x_mat(:,i)=expm(A*t)*x_iv;
end
for j=1:length(x_iv)
subplot(length(x_iv),1,j), plot(t_vec,x_mat(j,:)),
hold on, grid on
end
end
ylabel('t /[s]')
subplot(length(x_iv),1,1), ylabel('v_z /[m/s]')
subplot(length(x_iv),1,2), ylabel('F_s-m*g /[N]')
subplot(length(x_iv),1,3), ylabel('z /[m]')
Figure 1-35: § An example in Matlab code, where expm is used to find the solution.
1.5.4.1.2 § Simulation by Programming Integration
In Figure 1-36, the Matlab function f shows an implementation of an explicit form model:
function [derv,derx, F] = f(t, v,x)
m=2; c=3;
F=c*x;
55
Introduction
derv=-F/m;
derx=v;
end
The following Matlab script simulates it (using simplest possible derivative approximation, Newton
forward) and plots the result.
>> clear all; dt=0.001; t_vec=[0:dt:5]; v_vec=0; x_vec=1; F_vec=[];
for i=1:length(t_vec)
[derv,derx,F_vec(i)]=f(t_vec(i),v_vec(i),x_vec(i));
v_vec(i+1)=v_vec(i)+derv*dt;
x_vec(i+1)=x_vec(i)+derx*dt;
end
figure(1), clf
plot(t_vec',[v_vec(1:end-1);F_vec;x_vec(1:end-1)]')
grid on, legend('v','F','x')
The states x,v has to be given initial values. The output variable F is not a state and has to be treated
slightly different.
1.5.4.1.3 § Simulation of a model on explicit form using built-in Integration
Methods
There are many other built-in integration methods in Matlab, e.g. euler, ode23, ode45 and
ode23s. The same example as above can also be simulated with such, here exemplified with ode23:
>> clear all; [t_vec,States]=ode23('f2',[0 5],[0 1]);
figure(1), clf, plot(t_vec,States), grid on, legend('v','x')
A wrapper function f2 has to be defined to meet the interface format for ode23:
function [derStates] = f2(t, States)
[derv,derx, F] = f(t, States(1),States(2));
derStates=[derv;derx];
end
In these tools, the Explicit form model (or an ODE) is built with a graphical representation, around “in-
tegrator blocks”, often marked “1/s” or “∫ ”. An example using Simulink is shown in Figure 1-36. Sim-
ulink is designed for designing/modelling signal processing and control design. It can also be used for
modelling the physics of the controlled systems. There are no dedicated vehicle dynamics tools/librar-
ies from Mathworks (but there are in-house developed specific libraries in automotive companies).
From this type of tools, it is often possible to automatically generate real time code, which is more and
more used instead of typing algorithms. It can be used for rapid prototyping of control functions, or
even for generation of executable code for production ECUs.
56
Introduction
x
𝑥
v
Mathematical model:
(3 DAE equations, 3
unknown variables)
Figure 1-36: Graphical modelling using Simulink for Explicit form model.
These tools are frequently used in professional vehicle development. They contain purpose-built and
relatively advanced models of vehicles, drivers and environments (ground, road, traffic, …). They are
well prepared for parameter changes. However, they are generally less prepared for modelling con-
ceptually new vehicle designs, which can make these tools less useful for vehicle manufacturers and
advanced research. For this reason, many of these tools offer also an interface to Simulink or FMU, so
that the user can add in their own vehicle models or export parts to other tools.
Modelica is not a tool but a globally standardized format for lumped dynamic models on DAE form (or
Mathematical form, see 1.5.1.1.3). There are several tools which supports the format. Specification of
Modelica is found at https://www.modelica.org/. When learning Modelica, http://www.xogeny.com/
is helpful. The model format is acausal and all variables and parameters are declared. An example of
model is given in Figure 1-37. Declaration of which are parameters and variables is a necessary part
for a DAE model, see “parameter Real” and “Real”, respectively.
The model format is also object oriented, which means that libraries of model components are facili-
tated. These are often handled with graphical representation, on top of the model code. There are
some open-source libraries for various physical domains, such as hydraulic, mechanics, thermodynam-
ics and control. There are also commercial libraries, where we find vehicle dynamics relevant compo-
nents: Vehicle Dynamics Library and Powertrain Library. Some simple Modelica code will be used to
describe models in this compendium.
57
Introduction
Physical model:
Simulation in Dymola:
𝑥 (Mathematical) Model
in Modelica:
model ExampleModel
parameter Real m=2;
parameter Real m=2; parameter Real c=3;
parameter Real c=3; parameter Real F0=4;
Real v; Real v;
Real F(start=3); Real F;
equation Real x(start=-1/3); The tool manipulates the 3 equations to
m*der(v) = -F; equation
der(F) = c*v;
this explicit form:
m*der(v) = -F;
v = der(x); 𝑒𝑟( ) /
Corresponding explicit form: 𝑆𝑡 𝑡𝑒𝐷𝑒𝑟 𝑡 𝑒𝑠
F = F0+c*x; 𝑒𝑟(𝑥)
𝑒𝑟( ) / end ExampleModel; 𝑢𝑡𝑝𝑢𝑡𝑠 𝑥
𝑒𝑟( )
Figure 1-37: Example of model in Modelica format (using the tool Dymola). Two alternative
models are given, leading to either [x,v] or [v,F] as states.
Mathematical modelling is more efficient than Explicit form modelling, since the engineer does not
need to spend time on symbolic/algebraic manipulation of the equations. This is especially true when
a model is reused in another context which changes the causality or for so called “higher index prob-
lems”, see Figure 1-38. In this compendium, many models are only driven to Mathematical model,
since it is enough if assuming there are modern tools as Modelica tools available. One the other hand,
an explicit form model has the value of capturing the causality, i.e. the cause-to-effect chain. The cau-
sality can sometimes facilitate the understanding and in that way help the engineer, which is why at
least one and rather complete model is shown as explicit form model, see 4.5.3.2.
One can also declare variables with prescribed causality, i.e. signals, in Modelica. Declaration of input
signal: “input Real z;”. Modelica can also handle sampled signals and discrete states.
2 2
2 2
2
2
(Mathematical) Model
in Modelica:
model ExampleModel_HigherIndex
parameter Real m1=0.5;
parameter Real m2=1.5;
parameter Real c=3;
Real v1;
Real v2; The tool first realizes need for differentiating, to get
Real F1;
another algebraic equation:
Real F2(start=3);
equation 2 ⇒ 𝑒𝑟( ) 𝑒𝑟( 2)
m1*der(v1)
F1 = -F1;
+ m1*der(v1) = 0; The tool then selects state variables and manipulates the
m2*der(v2)
F2-F1 = F1-F2;= 0;
+ m2*der(v2) equations to this explicit form:
v1=v2; 𝑒𝑟( ) 2 /( + 2)
der(F2) = c*v2; 𝑆𝑡 𝑡𝑒𝐷𝑒𝑟 𝑡 𝑒𝑠
𝑒𝑟( 2)
end ExampleModel_HigherIndex; 2
𝑢𝑡𝑝𝑢𝑡𝑠 /( +
2 2)
58
Introduction
-1
-2
-3
-4
automatic, 0 2 4 6 8 10
by the tool
same, in text editor:
model ExampleModel_ModelicaLibraries
Modelica.Mechanics.Translational.Components.Fixed fixed;
Modelica.Mechanics.Translational.Components.Spring spring(s_rel0=1, c=3);
Modelica.Mechanics.Translational.Components.Mass mass1(m=0.5);
Modelica.Mechanics.Translational.Components.Mass mass2(m=1.5);
equation
connect(mass1.flange_b, mass2.flange_a); model Mass "Sliding mass with inertia"
connect(mass2.flange_b, spring.flange_a); parameter SI.Mass m(min=0, start=1);
connect(spring.flange_b, fixed.flange); parameter StateSelect stateSelect=StateSelect.default;
end ExampleModel_ModelicaLibraries; extends Translational.Interfaces.PartialRigid;
SI.Velocity v(start=0, stateSelect=stateSelect);
SI.Acceleration a(start=0);
equation
v = der(s);
a = der(v);
m*a = flange_a.f + flange_b.f;
end Mass;
Figure 1-39: Example of model in Modelica format, using Modelica libraries of component.
1.5.4.5.1 § OMwebbook
The cloud computation service OMwebbook for the example in Figure 1-37 is shown in figure below.
Figure 1-40: Modelica format, using OMwebbook, for same example model as in Figure 1-37
Contribution from Pinar Boyraz, Vehicle Safety at Chalmers
FMI (Functional Mock-up Interface) is not a tool but a globally standardized format for dynamic mod-
els on explicit form (see https://fmi-standard.org/). There are several tools which supports this for-
mat. FMI enables model export/import between tools. It also allows to hide Intellectual Property (IP)
by using “black-box format”, i.e. models compiled (non-human readable) for certain processors, which
59
Introduction
is important in relation between OEMs and suppliers. An FMU (Functional Mock-up Unit) is a model on
FMI format.
1.5.4.6.1 § Model Exchange and Co-Simulation FMUs
FMUs can be either of Model Exchange type (requires external solver) or Co-Simulation type (includes
an own solver for the own states).
Figure 1-41: § FMI standard models. Left: FMU for Model Exchange. Right: FMU for Co-Simulation.
There are also add-on standards for combining several FMUs in an hierarchy (System Structure and
Parameterization SSP) and for co-simulation with FMUs (Distributed Co-Simulation Protocol DCP).
60
Introduction
Co-Simulation and Real-time simulation very often they appear at the same time. Many challenges ap-
pear, such as keeping integration error under control, computation fast enough for real time, synchro-
nising variables, etc.
1.5.4.7.2 § Co-Modelling
An alternative way where models of sub-systems are connected on Mathematical form, as opposed to
Explicit form. It could be called “co-modelling” (or “acausal co-simulation”) and make it easier to or-
ganize libraries of models since one would not need to keep different versions of part models for dif-
ferent causality on interface variables. However, co-modelling would require a globally accepted and
used standard of Mathematical modelling format, which does not exist yet; Modelica is the closest to
reach such status. Modelica has a proposal for encryption of models, which would be important for
model exchange with kept confidentiality.
ground, road
or horizontal ISO 8855
plane z
yaw y
transversal
plane
x
symmetry
plane
Figure 1-42: Left: Vehicle (motion) degrees of freedom and important planes. Right: ISO
coordinate system
61
Introduction
Velocity of wheel
=v hub i,
lateral=y x
yaw rate= =wz y
𝑤
𝑜𝑟
𝑜𝑟
Δ 5 or 1,3,1 3 or 1,2,1 1 or 1,1,1
Δ
6 or 1,3,2 4 or 1,2,2 2 or 1,1,2
2 𝑜𝑟 2
If number of units, 𝑛 > , there Δ Δ Dpz=
are >1 couplings. Then use, e.g., =articulation angle of coupling 2 𝑜𝑟 2
𝑗 for ⋯ 𝑛𝑢 and 𝑗 𝑓 𝑟 or
axle 3 axle 2 axle 1
… for ⋯ 𝑛𝑢 .
or 1,3 or 1,2 or 1,1
unit 2 unit 1
Figure 1-44: Proposed numbering of units, axles, wheels and articulation angle. Example shows a
rigid truck with trailer.
If multiple units: 𝑓 , so the rule is “articulation (yaw) angle in front cou-
pling on a unit is the yaw angle of the unit ahead, relative to the unit ”. This can be extended to
62
Introduction
1.6.1.1 Couplings
Coupling forces can be denoted using “single cut force notation” or “multiple cut forces notation”, see
1.5.2.1.1. Coupling, in 1.6.1.1, means moment-free 2D connection between bodies. Similar reasoning is
made for transmission coupling in 2.4.1.
1.6.1.1.1 Single cut force notation
front end of unit + : rear end of unit :
Figure 1-45: Coupling force notation, using “single cut force notation”, for the coupling named
“ 𝑅” (for unit , coupling Rear).
For single cut force notation, it can be an alternative to replace subscript “ 𝑅” with a sequential num-
bered identifier of the coupling within the vehicle, e.g. ⋯ (𝑛 ), where 𝑛 are the number of
units.
1.6.1.1.2 Multiple cut forces notation
free body diagram of coupling:
compatibility of coupling:
+
+ +
Figure 1-46: Coupling force notation, using “multiple cut forces notation”, for the coupling
between unit and + . Subscripts: “ 𝑅” as in Figure 1-45 and “ + ” after unit + ,
coupling Front.
For multiple cut forces notation, introduces more notations than needed to describe the physical real-
world situation. Hence one need extra transformation equations in Eq [1.1] and Eq [1.2]. Also, extra
equations are needed to set the front coupling forces of the first unit and the rear coupling forces of
the last unit to zero.
Compatibility:
63
Introduction
Force equilibrium of coupling (or Force equivalence between forces on the two units):
From unit to unit + :
+ cos( ) sin( )
[ ] [ ]∙[ ]
+ sin( ) cos( )
[1.2]
Same equation, but expressed as from unit + to unit :
cos( ) sin( ) +
[ ] [ ]∙[ ]
sin( ) cos( ) +
𝑣 𝑤
𝑤
𝑣
64
Introduction
A sign convention, aligned with articulation angles (between units) and steering angles, would be: Rel-
ative roll angle of an axle= , so the rule is “axle orientation of axle 𝑗 is relative to the ve-
hicle body”. This can be extended to translatory position and orientation in yaw and pitch.
2 2 2 2
√ 𝑣 + 𝑣 √ 𝑤 + 𝑤
𝑣 cos( ) sin( ) 𝑤
[ ] [ ]∙[ ] or [1.4]
𝑣 sin( ) cos( ) 𝑤 arctan(
⏟ 𝑣⁄ 𝑣) + arctan(
⏟ 𝑤⁄ 𝑤)
{ 𝛽 𝛼
If approximating with small angles, it is easier to approximate to a linear vehicle model, see 4.3.2.4 and
4.4.2.2. For reversing, the angles are close to ±𝜋 instead of close to 0.
For ≈0 ≈ ≈0⇒ 𝑣 >0 𝑤 >0 For ≈0 ≈ ≈ ±𝜋 ⇒ 𝑣 <0 𝑤 <0
𝑤 𝑤
𝑠 𝑤 𝑠 𝑤
𝑤 𝑤
𝑥
𝑦
vehicle body
+𝑥 𝑣 +𝑥
arctan ≈
𝑣 𝑣 arctan 𝑣 𝑣 ≈𝜋 𝜋+
𝑣 𝑣
𝑦 𝑣 𝑦
𝑣 𝑣
arctan 𝑠 ≈𝑠 𝜋 arctan 𝑠 𝑤 ≈𝜋 𝑠 𝑤
𝑤 𝑤
𝑜𝑡ℎ +𝑥
+ ⇒ ⇒ ≈ +𝑠 𝑤
𝑝𝑝𝑟𝑜𝑥. 𝑦
Figure 1-48: Approximate compatibility relation for small side slip 𝑤 . Left: For forward driving.
Right: For reversing. Both lead to same final equation.
65
Introduction
The vehicle also has a varying orientation, (𝑥) or (𝑠), which often is often relevant, but the term
“path” does necessarily include this. In those cases, it might be good to use an expression “path with
orientation” instead.
If the vehicle is articulated, it has not only the orientation as additional variable over the path. It also
has the articulation angle in each articulation point. The set of articulation angles 𝜽 describes the
“pose”, so the total description would be [𝑥 𝑦 𝜽] and called “path with orientation and pose”.
A (time) trajectory is a more general term than a path and it brings in the dependence of time, 𝑡. One
typical understanding is that trajectories can be [𝑥(𝑡) 𝑦(𝑡) (𝑡) 𝜽(𝑡)]. But also, other quantities,
such as steer angle or vehicle propulsion force can be called trajectory: (𝑡) and (𝑡), respectively.
The word “trace” is sometimes used interchangeably with trajectory.
1.6.1.6.1 Path with Orientation
The path and path with orientation was y
introduced in 1.5. The path, in global co-
ordinate system, is related to vehicle
speeds, in vehicle fix coordinates, as given
in Figure 1-49 and Equation [1.5].
Knowing ( (𝑡) (𝑡) (𝑡)), we can de-
termine “path with orientation”
(𝑥(𝑡) 𝑦(𝑡) (𝑡)), by time integration of x
the right-hand side of the equation.
Hence, the positions are typically “state Figure 1-49: Model connecting “path with orientation”
variables” in lateral dynamics models. to velocities in vehicle coordinate system.
𝑥 cos( ) sin( )
[ ] [ ]∙[ ]
𝑦 sin( ) cos( ) [1.5]
It should be noted that in some problems, typically manoeuvring at low speed, the real time scale is of
less interest. Then, the problem can be treated as time independent, e.g. by introducing a coordinate, s,
along the path, as in Equation [1.6].
𝑥′ ∙ cos( ) ∙ sin( )
𝑠 𝑠
𝑦′ ∙ cos( )+ ∙ sin( )
𝑠 𝑠 [1.6]
′
𝑠
where prime notes differentation with respect to s
Here, 𝑠 can be thought of like an arbitrary time scale, with which all speeds are scaled. One can typi-
cally choose 𝑠 [ ⁄𝑠]. However, in this compendium we will keep notation t and the dot notation
for derivative.
66
Introduction
2
𝑊
𝑊
𝑊𝑓
𝑊
𝑓
3 2 0
𝐿 𝐿
Figure 1-50: Two examples of dimension definitions and notation.
≈ ( ) ∑ 𝑓𝑜𝑟 𝑒𝑠
𝐈n-road-plane
( + )≈ ( + ) ∑ 𝑓𝑜𝑟 𝑒𝑠
(irp) equilibria:
{ ∑ 𝑜 𝑒𝑛𝑡𝑠
[1.7]
( )≈ ( + ) ∑ 𝑓𝑜𝑟 𝑒𝑠
Out-of-road-plane
∑ 𝑜 𝑒𝑛𝑡𝑠
(oorp) equilibria:
{ ∑ 𝑜 𝑒𝑛𝑡𝑠
Chapter 3 and 4 requires complete vehicle models that can measure complete vehicle functions ex-
pressed in irp motion.
67
Introduction
Low speed condition means that vehicle moves with low speed irp: forward, reverse or stand-still. Also,
accelerations are low. No fictive forces ( 𝑠𝑠 𝑒 𝑒𝑟 𝑡 𝑜𝑛) is modelled irp. All terms on left side in
Eq [1.7] are neglected.
Steady State
Steady State condition means that time history is irrelevant for the quantities studied. Seen as a ma-
noeuvre over time, the studied quantities are constant. Explained for a certain model, it means that the
influence of the derivatives of the corresponding variables, that else would generate state variables, is
neglected. In physical model of a mechanical system this often means that “mass acceleration ” or
“force derivative⁄stiffness ” is neglected (in mathematical model: equation “ 0 ” is added, in explicit
form model: “ “ is replaced with “0” ). If only some quantities are treated in this way, one might call
the conditions steady state with respect to these quantities. Only the terms (centrifugal
forces) on left side in in Eq [1.7] are kept.
Transient
Transient (or Transient State, as opposed to Steady State) condition means that time history is rele-
vant; i.e. there are delays, represented by “state variables” when simulated. All terms on the left side in
in Eq [1.7]are kept. Note that also cases with low velocities but large accelerations belongs here, and
not in 1.6.2.1.1.2.
Stationary Oscillating
Stationary Oscillating condition is a special case of transient, where cyclic variations continue over
long time with a repeated pattern. Long time means to that all none-cyclic components of the variation
is damped out. The pattern is often modelled as harmonic (sinus and cosine variations in time) with
constant amplitudes and phases. Example is sinusoidal steering with small enough steering amplitude,
see 4.4, but also driving with over an undulated road surface, see Chapter 5. All terms on left side in in
Eq [1.7] are kept, as for transient operation.
Quasi-Steady State
Quasi-Steady State condition have a more diffuse meaning. It can refer to steady state with respect to
some quantities, i.e. some terms on left side in in Eq [1.7] are neglected but not all. Alternatively, it can
refer to that the quantities are prescribed to an explicit function of time, e.g. 𝑓 (𝑡) meaning that
also is known. The latter is sometimes also known as “inverse dynamics”.
1.6.2.1.2 Other Characterizations
Let us also briefly list some other possible categorizations for Chapter 3 and 4:
• Categorization referring to small angles (sin( 𝑛 𝑒) cos( 𝑛 𝑒) 0 ) or not, applied to
steering/articulation angles and/or tyre/body side-slip angles.
• Categorization referring to tyre models are further explained in 2.2.6.
• One can also categorize referring to subsystem models.
o The suspension can add states per wheel or axle , at least the vertical spring force .
o The propulsion and brake system can be actively controlled. They give wheel torques 𝑻(𝑡)
[ 2 ⋯ ]. Such models often add one state per wheel or axle , the rotational speed .
o Chapter 3 uses different models of propulsion, brake and suspension. E.g., the propulsion sys-
tem can add more states: engine delay, gear shifting, torsional shaft, control algorithms.
o Chapter 4 adds the steering subsystem, with its control algorithms. Also, more about model
categorisation is found in 4.1.1.
Chapter 5 is very different since it treats vehicle functions expressed in out-of-road-plane motion. So,
most models in Chapter 5 uses only the oorp equations, i.e. not the irp equations, above.
To make the overall model simulate-able one needs some form of driver model and environment
model. A rather complete example model is described in 4.5.3.2.
1.6.2.2 1D Models
In some cases, it can be enough to model motion in 1 dimension. Some examples are:
• Longitudinal 1D in 3.2 and 3.3, it can be enough to model : ∑ 𝑓𝑜𝑟 𝑒𝑠 .
68
Introduction
𝑅 𝑅
𝑓
𝑓
𝑅2 2
𝑅2
2
2
2
22 𝑅
𝑅
2 2
22 3
2 2
2
𝑓 3
𝑓 2
Figure 1-51: Different longitudinal 1D models.
69
Introduction
Figure 1-52: Different in-road-plane vehicle model concepts. The forces are from tyre contacts,
acting on vehicle body. Fictive forces ( 𝑒 𝑒𝑟 𝑡 𝑜𝑛) and body forces (gravity and
aerodynamic) are not drawn.
1.6.2.3.2 One-Track and Two-Track Models
The models can assume that each wheel on the axles have their own ground contact (two-track mod-
els) or if there is only one “virtual tyre” modelled per axle (one-track model or single-track model or
“bicycle model”). See Figure 1-52 and Figure 1-53. For multi-axle vehicles, one can even simplify one
step further and model only one “virtual tyre” per axle group, see Figure 1-52. It should be noted that
the simplifications have limitations, they are not suitable when the forces of the different tyres are sig-
nificantly different, e.g. differences in actuated wheel torques or different wheel side slips.
𝑓
𝑓
𝑓
Common
intersection of all
wheels’ axes of
rotation 𝑅
Figure 1-53: Collapsing a two-track vehicle to a one-track model.
Lateral dynamics phenomena which one-track models do not capture well are, e.g.:
• Deviations from Ackerman geometry within an axle, see 0.
• Roll motion
• Lateral load shift (4.3.7.2) and combined tyre slip (4.3.7.6).
• Added yaw moment due to left/right-asymmetric wheel torque, such as ESC interventions.
The in-road-plane (irp) models does not capture the out-of-road-plane motion, . However,
they can still capture the transfer of loads (vertical forces on wheels).
70
Introduction
For longitudinal dynamics, one-track models are often enough. For longitudinal functions in “3.2
Steady State Functions” and “3.3 Functions Over (Long) Cycles”, even particle models are often
enough. But, it is important to understand the validity limit of the model!
71
Introduction
A test with real vehicle, carried out with a steering-robot (and/or pedal robot) can also be called
closed-loop if the robot is controlled with a control algorithm which acts differently depending on the
vehicle states, i.e. if the algorithm is a driver model.
With increasing level of automation, there is sometimes a need for distinguishing between
closed/open loop with respect to human driver or automated driver.
1.6.3.2.1 § Closing Loop via Human and/or Virtual Driver
The concepts of open- and closed-loop control are demonstrated in figure below. The important differ-
ence to note is that in open-loop control there is no feedback to the driver. The figure also shows a
“Virtual driver” which represents today’s automated driving functions based on environment sensing,
such as adaptive cruise control (section 3.5.2.2) and lane keeping aid (section 4.6.2.1). Closed loop con-
trol incorporates the different types of feedback to the driver. Drivers automatically adapt to the dif-
ferent feedback. Understanding driver response to different feedback, and expressing it in mathemati-
cal descriptions, is an active area of research, particularly for driver support functions.
Open-loop system
Environment sensing
(radar, camera, GPS, V2X, …)
“Virtual driver”
Requests (signals)
Visual
Steering
Suspension,
Surroundings / system,
Environment Human Propulsion Vehicle
Linkages,
(air, road, other Driver Steering system, Road wheel Forces, body
road users, …) wheel, Brake angles, Wheel Moments
torques, … Wheels
Pedals, … system
Absolute
Position, Warnings, Info
HMI (lamps,
Speed sounds, …)
Sensed quantities
Noise,
vibrations
Steering wheel torque Inertial
forces
Figure 1-54: § Open and closed loop representations of vehicle dynamic systems
72
Introduction
for roll-over instability. Also, the articulation angle can be unstable when reversing in low speed with a
trailer. Stability analysis in a stricter physics/mathematical meaning is touch upon in 0.
It is useful to know about this confusion of words. An alternative expression for the wider meaning is
“loss of control” or “loss of tracking” or “directional unstable”, which can include that vehicle goes
straight ahead, but road bends.
73
Introduction
1.6.4.1 Subsystems
The architectures are dependent on the business model for how to purchase and integrate subsystems
to a vehicle. Hence, it is relevant to define the subsystems. For vehicle motion functionality, the rele-
vant subsystems (or “motion support devices”) are typically:
• Propulsion system
• Brake system
• Wheel suspension
• Wheels and tyres
• Steering system
• Environment sensing system
Each of these can typically be purchased as one subsystem. Each typically have mechanical and signal
interface to the remaining vehicle. Different vehicle manufacturers can have different strategies for
signal interface and how much the subsystems are allowed to be dependent of each other.
Figure 1-55: Concepts of a vehicle motion (and energy) function architecture. Arbitrators,
Coordinators and Actuators are the most important architectural objects for vehicle dynamics
(vehicle motion).
74
Introduction
Vehicle Transport Mission, Route, Route Segment Layer Predictive Energy Route HMI Human
Environment Management Machine
Sensing Longitudinal Speed, Energy Split Interface
Traffic Situation Layer Longitudinal
ACC
Traffic
LaneSteering Driver ACCbtn
Interpretation
V2V / V2I Collision Interpretation Dist btn
Maps Avoidaice Arbitration
…
FwdSens Longitudinal Acceleration,
Vehicle Capability Actual Power Split SteWhl
AnSens
Vehicle Motion Layer
Driver APed
Vehicle Interpretation
Vehicle Control Arbitration
State BPed
Estimator Momentary irp Accelerations,
Energy Power Split
Management Accel to Force
Vehicle Stability
Capability Control Coordination
Estimator
requests
actual values
capabilities
physical connections
Figure 1-56: One example of reference architecture of vehicle motion functionality. Red arrows:
Requests, Blue arrow: Information, Black lines with dot-ends: physical connection.
In order to be able to formulate design rules in reference architecture of functionality the following are
relevant questions:
• Which physical quantities should be represented on the interface between Devices (Sensors and
Actuators) and Vehicle Level Functionality?
• Partitioning within a reference architecture of vehicle motion functionality could be realised as
shown in Figure 1-56. Different Layers/Domains are defined:
• Human Machine Interface Domain: This includes the sensors/buttons which the driver
uses to request services from the vehicle’s embedded motion functionality.
• Vehicle Environment Domain: Includes surrounding sensors mounted on vehicle but also
communication with other vehicles (V2V) and infrastructure (V2I) and map information.
• Route Layer: Planning the whole transport mission, horizon 10..1000 km. These functions
exist mainly for commercial traffic and might be done outside the subject vehicle.
• Traffic Situation Layer: Interpret the immediate surrounding traffic which the vehicle is
in, road/lanes and other road users, horizon ≈100..300 m. Example of functions: adaptive
cruise control, collision avoidance, and lane steer support.
• Vehicle Motion Layer: This includes the Energy management, powertrain coordination,
brake distribution, and vehicle stability such as ESC, ABS. Horizon ≈10..30 m This layer
also estimates the vehicle states e.g. . In addition, this layer would be able to give
vehicle level capability of max/min acceleration and their derivative.
• Motion Support Device Layer: This includes the devices/actuators that can generate ve-
hicle motion. This layer is also consisting sensors which could include the capability and
status of each device e.g. max/min wheel torque.
• Formalisation of different types of:
• Blocks, e.g. Controller, Information Fusion, Arbitrator and Coordinator.
75
Introduction
• Both arbitrators and coordinators have inputs and outputs as requests, typically ex-
pressed in same physical quantity. An arbitrator has more requests in that out and a
coordinator has more requests out than in.
• Signals, e.g. Request, Actual (or Status) and Capability.
• Parameters used in Functional blocks. One can differ between Physical parameters (or
Model parameters) and Tuning parameters. Some parameters can be common across the
whole vehicle, which enables a kind of communication between blocks without normal
signals, but instead exchanging values during start-up of the system.
Each signal should have a definition of which physical quantity and unit it refers to. For signals of Ac-
tual type, there also needs a concept to handle how accurate they are; e.g. as a tolerance or sending an
upper and lower value between the physical quantity shall be. The definition of physical quantity is
very important, not the least for Request signals. An example is, if using tyre longitudinal slip as quan-
tity, sender and receiver of signal have to agree on slip definition. Since there are many slip definitions
around, it might suggest using wheel rotational speed instead, which is less ambiguous.
The functionality is then allocated to ECUs, and signals allocated to network communication. The ref-
erence architecture can be used for reasoning where the allocation should be done. Which functional-
ity is sensitive for e.g. time delay and thus should be allocated in the same ECU? Detailed control algo-
rithm design is not stipulated by a reference architecture. Instead the reference architecture assists
how detailed control algorithms be managed in the complete problem of controlling the vehicle mo-
tion. Whether solutions of Functional Safety (ISO 26262, 2011-2012) is represented in a reference ar-
chitecture of functionality can vary.
Vehicles consisting of several units add special challenges, especially if the units are actuated. A heavy
truck trailer is always actuated with at least brakes.
It becomes increasingly important for vehicle manufacturers to take responsibility of their products.
The main reason is that numbers and complexity of sensors, actuators and control functions increases.
The risks are often connected to how vehicle moves, i.e. to vehicle dynamics/vehicle motion. The num-
ber of risks increase both because the systems becomes more complex and because more information
is available on-board. The availability of information can mean that it is ethically wrong to not act on
it, e.g. to not automatically brake if camera detects a possible pedestrian ahead. An acceptable excuse
to not brake could however, e.g., be that braking would mean a more severe risk due to causing a colli-
sion from rear-coming traffic.
Failure Mode Effect Analysis (FMEA) has been around for a long time and it is applicable on any design
in any product, and not only safety aspects. Basically, one identifies things that can go wrong with a
certain conceptual design early in development and decide how to follow it up during the develop-
ment. The follow-up normally includes to formulate requirements on the design. ISO26262 (ISO
26262, 2011-2012) describes a similar way of working, but stricter and more specialized on traffic
safety hazards and vehicles as products.
A very brief description of (ISO 26262, 2011-2012) follows here, see also figure below. A certain Item
(part, subsystem or function) is identified for Hazard Analysis early in product development. Hazards
are defined and assessed to a certain Automotive Safety Integrity Level (ASIL) in a Risk Assessment. The
ASIL levels, listed with increasing severity, are QM, ASIL-A, ASIL-B, ASIL-C, ASIL-D. The levels ASIL-A to
76
Introduction
ASIL-D require Safety Requirements managed under an ISO26262 process during the continued devel-
opment, while the QM does not. A Safety Requirement is a requirement on the design which reduces
the ASIL to a tolerable level.
Conceptually, one can calculate ASIL for a certain hazard: 𝐴𝑆𝐼𝐿 𝑅 𝑠𝑘 𝑆𝑒 𝑒𝑟 𝑡𝑦 ( 𝑥𝑝𝑜𝑠𝑢𝑟𝑒
𝑜𝑛𝑡𝑟𝑜 𝑡𝑦) . For instance, loss of acceleration has generally lower Severity than loss of decelera-
tion. For instance, the Exposure becomes lower if the hazards only appears in sharp turns or if there is
redundancy in the solution. For instance, a hazard during automated driving increases the ASIL via the
Controllability factor, since human driver is then less alert and less likely to control the situation.
ISO26262 describes how the work should be documented. The documentation can be asked for in a
legal court case.
Figure 1-57: § Overview of the safety life cycle, from (ISO 26262, 2011-2012).
77
Introduction
78
Introduction
Requirements
ulator can utilize HIL, SIL
Repeatable and Parallelized testing + + -
and MIL, from 0, for some Safe testing + + -
part of the vehicle model. Representative integration in vehicle - +! ++
Driving simulator can be w.r.t. to vehicle - - +!
Represen-
compared to other verifi- tative
w.r.t. driver - +! +
cation methods, see Figure behaviour w.r.t. to surroundings -- -
(road & traffic) (env. sens) (env. sens) +!
1-59. Subjective scale from
Figure 1-58 can be used. Figure 1-59: Comparison of verification methods for complete vehicle
functions.
79
Introduction
1.7.3 Definitions
In Figure 1-60, we can find the following units:
• Towing units: Tractor or Rigid (Truck)
• Towed units: Full trailer (FT), Semi-trailer (ST), Centre-axle trailer (CAT), (Converter) Dolly
• The couplings between the units can of 2 types:
o Fifth-wheel coupling (e.g. between Tractor and Semi-trailer). Designed to take force in all
3 directions. Furthermore, it is designed to be roll-rigid, but free in pitch and yaw.
o A Turntable is similar to a fifth wheel but has a rolling bearing instead of a pitch-hinged
greased surface, which leads to less yaw friction and no pitch degree of freedom. In con-
verter dollies, one often sees both, fifth wheel on top of turntable.
o Hitch coupling or Drawbar coupling (e.g. between Rigid and Full trailer). Designed to
take forces in longitudinal and lateral directions, but only minor in vertical direction. Also,
it is rotationally free in all 3 directions. A full-trailer has pitch-moment-free rear end of
drawbar. A converter dolly or centre-axle-trailers have pitch-rigid rear end of drawbar.
• There is often a need for several axles close to each other, to manage a high vertical force. Such
group of axles is called “axle group”, “axle arrangement” or “running gear”.
80
Introduction
A-double FT
Tractor-SemiTrailer- ≈30 m, ST Tra
FullSemiTrailer 80 t
Truck Truck-CentreAxleTrailer- 27.5 m, CAT CAT Tru
DuoCAT CentreAxleTrailer 66 t
AB-double Truck-Dolly-LinkTrailer- 34 m, ST Link Tru
or Truck B- Dy
double SemiTrailer 90 t
81
Introduction
• Transport and user type: Short travels (typically urban, 5-10 km, 50 km/h) or long travels (typi-
cally inter-urban, 10-30 km, 100 km/h).
• Chassis concept:
o Narrow (e.g. normal bicycles and motorcycles) or wide (at least one axle with 2 wheels, re-
sulting in 3-4 wheels on the vehicle). Note that UPVs in both categories are typically still less
wide than passenger cars.
o Roll moment during cornering carried by suspension roll stiffness or roll moment during cor-
nering avoided by vehicle cambering. The first concept can be called “Roll-stiff vehicle”. The
second concept can be called “Cambering vehicle” or “Leaning vehicle”. One-tracked vehicles
are always Cambering vehicles. two-tracked are normally Roll-stiff, but there are examples of
Cambering such (see upper right in Figure 1-61). See Figure 1-62.
o (This compendium does only consider vehicles which are “Pitch-stiff”, i.e. such that can take
the pitch moment during acceleration and braking. Examples of vehicles not considered,
“Pitching vehicles”, are: one-wheeled vehicles as used at circuses and two-wheeled vehicles
with one axle, such as Segways.)
• Note that also Roll-stiff Vehicles camber while cornering, but outwards in curve and only
slightly, while Cambering Vehicles cambers inwards in curve and with a significant angle.
Cambering Vehicles is more intricate to understand when it comes to how wheel steering is
used. In a Roll-stiff Vehicle, the wheel suspension takes the roll-moment (maintains the roll
equilibrium), which means that driver can use wheel steering solely for making the vehicle
steer (follow an intended path). In Cambering Vehicles, the driver has to use the wheel steering
for both maintaining the roll equilibrium and following the intended path. A model is given in
4.5.2.3. Figure 1-62 shows some possible conceptual design of a cambering vehicle. It is not
possible to do a partly roll-stiff and partly cambering vehicle, unless one uses suspension
springs with negative stiffness, i.e. the suspension would need to consume energy to tilt the
vehicle inward in curve.
3 wheeled cambering Piaggio MP3
concept vehicle from BMW
Toyota
iRoad
Twizy from Renault
82
Introduction
Translate Lean sprung body Lean whole vehicle, Lean whole vehicle,
2 wheel axle single wheel axle
Figure 1-62: How Cambering vehicles avoid taking the roll moment while cornering. The vehicles
are viewed from rear and turns to the left. The dash-dotted arrows mark the resulting forces
which are equal between the 5 concepts.
83
Introduction
Notation
Categorization Recommended Alternatives
Subject for notation in code by hand Unit Description / Note
Basic (Shaft) Torque T T M Nm
Basic Coefficient of friction mu μ mue, muh 1=N/N
N/(m/s) or
Basic Damping coefficient d d c,k,D
Nm/(rad/s)
Basic Density roh ρ
Ratio between
Basic Efficiency eta hη 1=W/W useful/output power
and used/input power
Basic Energy E E W Nm=J
Basic Force F F f N
9.80665 average on
Basic Gravity g g m/s2
Earth
Basic Height h h H
Basic Imaginary unit j j i - j*j=-1
Basic Mass m m M kg
Subscript can be given
Basic Mass moment of inertia J J I kg*m2 to denote around
which axis
Basic Moment (of forces) M M T Nm Not for shaft torque
Basic Power P P W=Nm/s=J/s
Ratio between input
Nm/(Nm) or and output rotational
u,
Basic Ratio r r (rad/s)/(rad/s speed, or output and
i (ISO8855)
) input torque, in a
transmission
Basic Rotational speed w ω w rad/s
Basic Shear stress tau τ N/m2
1. General
N/m or
Basic Stiffness coefficient c c C, k, K
Nm/rad
Basic Strain eps ε m/m
Basic Translational speed v v V m/s
Basic Wave length lambda λ m
Can be track width,
Basic Width W W w m vehicle width, tyre
width, …
Dynamics (Time) Frequency f f 1/s=periods/s
Dynamics Angular (time) frequency w ω w rad/s
Dynamics Angular spatial frequency W Ω rad/m
Dependent variables in a dynamic Both state variables
Dynamics z z <various>
system and output variables
Input variables in a dynamic
Dynamics u u <various>
system
Dynamics Mean Square value MS MS
Output variables in a dynamic
Dynamics y y <various>
system
Dynamics Power spectral density G Φ PSD
Dynamics Root Mean Square value RMS RMS
Spacial frequency as radians per
Dynamics W Ω rad/m
travelled distance
1/m=periods/
Dynamics Spatial frequency fs fs
m
States variables in a dynamic
Dynamics x x <various>
system
The independent
Dynamics Time t t t, T s variable in a dynamic
system
Operators Transfer function H H
1. General Air resistance coefficient cd cd Cd 1
2. Vehicle
84
Introduction
Notation
Categorization Recommended Alternatives
Subject for notation in code by hand Unit Description / Note
1. General Mass m m M kg
1. General Rolling resistance coefficient f f 1=N/N
1. General Track width W W m
1. General Tyre stiffness C C c N/1 dF/ds at s=0
Kus, kus, N/(N/rad) =
1. General Understeer gradient Ku Ku
U (ISO8855) rad/(m/s2)
If no subscript, undefined
1. General Vehicle side slip angle b β b rad
or CoG
1. General Wheel base L L l, lf+lr, WB m
1. General Wheel radius R R r, Rw m
2. Road Curvature kappa κ roh, r rad/m=1/m road or path
2. Road Curve radius R R r, roh, r m road or path
Positive when right side
2. Road Road bank angle pxr φ xr of ground is lower than
left side ground
2. Road Road inclination angle pyr φ yr rad Positive downhill
2. Road Vertical position of road zr zr z, Z m
Vehicle acceleration, in interial ax, ay, decomposed in vehicle
3. Motion a x, a y, a z m/(s2)
system az coordinates direction
dervx, der(vx),
Time derivatives of each 2 decomposed in vehicle
3. Motion dervy, der(vy), m/(s )
component v x , v y , v z coordinates direction
dervz der(vz)
r=[rx,ry,rz] often position of Center
3. Motion Vehicle position x,y,z x, y, z m
or [X,Y,Z] of Gravity, CoG
decomposed in vehicle
3. Motion Vehicle velocity, in interial system vx,vy,vz v x, v y, v z u,v,w m/s
coordinates direction
From ground, air, towed
Fx, Fy, units, colliding objects,
F x, F y, F z, [N, N, N, Nm,
4. Forces Forces and moments on vehicle Fz, Mx, etc. May appear also for
M x, M y, M z Nm, Nm]
2. Vehicle
Forces on one wheel, axle or side Fixw,Fiyw F ixw , F iyw , in wheel coordinate
4. Forces lowercase f N
from ground ,Fizw F izw system
Sign convention positive
when either of "rotated
5. Angles Camber angle g γ g rad as wheel roll angle" or
"top leaning outward vs
vehicle body".
The angle required for a
Low speed or Ackermann speed
5. Angles dswA δswA rad certain vehicle path
steering wheel angle
curvature at low speeds
5. Angles Euler rotations ψ, θ, φ rad Order: yaw, pitch, roll
px, py, Angles, not rotations. Roll
5. Angles Roll, pitch, yaw angle φx, φy, φz j, q, y rad
pz and Pitch normally small
decomposed in vehicle
Vehicle angular velocity, in interial wx, wy,
5. Angles ωx , ω y , ω z rad/s coordinates direction
system wz
(roll, pitch, yaw)
May refer to steering
5. Angles Steering angle d δ delta rad
wheel, road wheel or axle
Normally an average of
5. Angles Steering angle of road wheels drw δrw RWA rad
angles on front axle
85
Introduction
Notation
Categorization Recommended Alternatives
Subject for notation in code by hand Unit Description / Note
5. Angles Steering angle of steering wheel dsw δsw SWA rad
6. Slip Tyre (lateral) slip angle a α alpha rad a=arctan(sy)
6. Slip Tyre lateral slip sy sy sy=tan(a)
1=
6. Slip Tyre longitudinal slip sx sx k, -k
(m/s)/(m/s)
6. Slip Tyre slip s s
used e.g. as Ca=cornering
7. Subscript axle a a
stiffness for any axle
COG, cog,
7. Subscript centre of gravity CoG CoG
CG, cg
Often used as double
7. Subscript front f f F, 1 subscript, e.g. fr=front
right
Often means with respect
7. Subscript inner i i
to curve
Often used as double
7. Subscript left l l L
subscript, e.g. fl=front left
2. Vehicle
86
Introduction
Figure 1-66: Typical values of parameters, common for typical passenger cars and heavy trucks.
87
Introduction
overall height =
overall height
3.7 m
=4 m
Figure 1-68: Typical data for a heavy vehicle, exemplified with “Tractor with Semi-trailer”
88
Vehicle Interactions and Subsystems
Freight
Vehicle Road surface construction, Covered
Weight with water/snow, Topography,
Payload, Curvature, Unevenness, …
Occupants, … and Location Tyre Forces
kg
Ground, Road
Environment
Figure 2-1: How the vehicle interacts with driver and environment.
The chapter also contains descriptions of the subsystems which are most relevant for Vehicle Dynam-
ics, with focus on Wheels and Tyres, see Figure 2-2.
The Environment Sensing System is described as a subsystem, see 2.7. There are also a lot of other
sensors in the vehicle, sensing the conditions in the subject vehicle, see 3.5, 4.6 and 5.10. These sen-
sors are often more established and therefore often treated as included in a certain subsystem, such as
wheel (rotational) speed sensors are often a part of the brake system.
89
Vehicle Interactions and Subsystems
(vehicle)
Propulsion body
Suspension
brake disk
Propulsion, Brake and Steering are
knuckle, bearing Brake “active” in the meaning that they are
“controllable through algorithms run
Wheel & Tyre on an electronic control unit ECU”.
Figure 2-2: Vehicle with the 6 Vehicle Dynamics relevant subsystems: Wheels and Tyres,
Suspension, Propulsion, Brake, Steering, and Environment Sensing. Towed units, with similar
subsystems, are not shown.
𝑇 𝑓 Wheel
shaft bolted axle end (non-driven) 𝑝
onto here
𝑤 𝑙 Non-driven: shaft ends
Wheel here, 𝑇 𝑓 0.
from http://19526-presscdn-0- bolted onto Driven: shaft continue to 𝑤 𝑙 axle end (driven) 𝑇
66.pagely.netdna-cdn.com/wp- propulsion system, via
𝑓
brake disk
content/uploads/2014/12/wheel-bearings.jpg drive-joint if steered.
⃗
Figure 2-3: Concepts of wheels are typically integrated. All vectors are 3D, while moment 𝑁 is 2D
in the wheels xz-plane and moment 𝑇 𝑓 is perpendicular to the same plane.
90
Vehicle Interactions and Subsystems
The structure of the carcass can be different: Bias-ply, Bias-ply Belted, and Radial-ply, see Figure 2-4.
Bias-ply tyres were the first types of pneumatic tyres to be used on motor vehicles. Radial ply tyres fol-
lowed 1946 and became the standard for passenger cars and is today also dominating also on trucks.
Note how the bias-ply constructions have textile structures oriented at an angle to the tyre centreline
along the x-z plane. This angle is referred to as the crown angle and is further illustrated in Figure 2-4.
Note the textile orientation for the bias-ply and radial tyres. Also note the difference in crown angles
between the two tyre constructions. This difference plays an important part in the rolling resistance
characteristics of the tyre which is 2.2.1.6.
The tyre components in Figure 2-4 have been constructed to provide the best tyre performance for dif-
ferent loading directions. Trade-offs are necessary between handling performance and comfort, be-
tween acceleration and wear, as well as between rolling resistance and desired friction for generating
forces in ground plane. The rubber components and patterns incorporated in the tread are critical to
the friction developed between the tyre and road under all road conditions (wet, dry, snow, etc.). Fric-
tion is most relevant in longitudinal and lateral vehicle dynamics. The belts define the circumferential
strength of the tyre and thus braking and acceleration performance. The sidewall and plies define the
lateral strength of the tyre and thus influence the lateral (cornering) performance of the vehicle. The
sidewall as well as the inflation pressure are also significant contributors to the vertical stiffness prop-
erties of the tyre and affect how the tyre transmits road irregularities to the remainder of the vehicle.
A tyre that has strong sidewalls will support vertical load well, but at the cost of vertical compliance.
bias-ply:
belt
sidewall
Figure 2-4: Left: Carcass Construction, (Wong, 2001). Left top: Bias-ply construction. Left bottom:
Radial construction. Right: Radial Tyre Structure, (Cooper Tire & Rubber Co., 2007).
From https://putneys.ca/winter-tires-time-
take-off-now/
Figure 2-5: Left: Tyre marking (radial tyre). Right: Typical summer and winter tread patterns.
91
Vehicle Interactions and Subsystems
𝑀
𝑀
𝑀
Figure 2-6: Tyre coordinate system. Forces and moments on tyre from ground. From (ISO 8855).
2.2.1.3.1 Steer Angle
A wheel on a vehicle has a Steer angle, . The Steer angle is the angle from vehicle longitudinal axis to
the wheel plane about the vehicle vertical axis (ISO 8855). Assuming vehicle 𝑥𝑦-plane is parallell to
road plane, this angle is same as angle from vehicle longitudinal axis to the intersection between wheel
plane and road plane.
2.2.1.3.2 Camber Angle
A wheel on a vehicle on a road surface has a Camber angle. The Camber angle is the angle from vehicle
longitudinal axis to the wheel plane about the vehicle longitudinal axis (ISO 8855). Assuming vehicle
𝑥𝑦-plane is parallell to road plane, this angle is same as the deviation from right angle between wheel
plane and road plane. A symmetry within an axle means that camber angle typically has opposite signs
on left and right wheel. However, for two-track vehicles, one can often see the sign convention that a
wheel top leaning outward from vehicle body is negative, regardless of on left or right side.
2.2.1.3.3 Steer and Camber
If a wheel has both significant Steer angle and significant Camber angle at the same time, these angles
alone does not define the orientation, since steering-before-cambering gives another orientation than
cambering-before- steering. One way is to express the orientation in a rotational angle around the
steering axis, see 2.2.1.3.5. An alternative is to use the concept of Euler rotations.
2.2.1.3.4 Castor Angle
A wheel, suspended to be steered, has a Castor angle (or Caster angle). This is the angle between the
vehicle vertical axis and the projection of the steering axis (2.2.1.3.5) on the vehicle 𝑥 -plane (ISO
8855). Castor angle is positive when top of steering axis is inclined rearward. Castor angle provides an
additional aligning torque, see 2.2.4.7 and 0, and changes the Camber angle when the wheel is steered.
2.2.1.3.5 Steering Axis Inclination
A wheel, suspended to be steered, has a Steering axis inclination (or Kingpin inclination).
92
Vehicle Interactions and Subsystems
This is the angle between the vehicle vertical axis and the projection of the steering axis (2.2.1.3.5) on
the vehicle 𝑦 -plane (ISO 8855). Steering axis inclination is positive when top of steering axis is in-
clined inward. See also 0.
2.2.1.3.6 Static Toe Angle
A wheel on an axle has a Static toe angle (a.k.a. Toe-in). The Static toe angle is the angle between vehi-
cle longitudinal axis and the wheel plane about the vehicle vertical axis, with the vehicle at rest and
steering in the straight-ahead position (ISO 8855). Positive Static toe angle (a.k.a. Toe-in or “ Toe-
out”) is when forward portion of the wheel is closer to the vehicle centreline than the wheel centre.
Typically, left and right wheel on same axle has same Static toe angle. Hence it measures an axle prop-
erty rather than a wheel property.
One could define an “instantaneous toe angle”, for an axle during arbitrary vehicle operation (non-rest
and non-zero steering), as ( + ( 𝑙 𝑓 ))⁄ . This will have a different value than Static toe angle
due to suspension linkage geometry and elasticity in suspension bushings.
Toe (regardless of toe-in or toe-out) generally generates opposing lateral forces on left and right side
leading to propulsion energy loss and tyre wear. Toe-out on front axle and toe-in on rear axle makes
the vehicle more yaw stable (less over-steered). Tone-in on front axle makes vehicle more yaw agile
but it also improves on-centre steering feel. Normal design choice for a passenger car is positive Static
toe angle (Toe-in) on both axles, and more on front axle.
Upper Sidewall
Air, inflated
Belt
Shaft
Lower Sidewall
(Front and Rear Sidewall are neglected here, but
could have been modelled as shear springs.)
increased
Upper
pull
𝑭𝒛
pull
decreased pull
Sidewall
but same
push
pull
Belt,
Lower
pull
pull
Figure 2-7: A pneumatic tyre is pre-tensioned by inflation pressure. This figure explains how the
wheel and tyre takes (vertical) load without (significant) increase of inflation pressure.
93
Vehicle Interactions and Subsystems
role of variables (i.e. they are varied) in the optimization, but not in time domain. This section follows
approximately the STI tyre model interface standard (Besselink, 2011), but not in all details.
Operational parameters
Driver
(Wheel and
Tyre) Design Behavioural Physical
Parameters Rest of
Tyre response Wheel response
variables variables vehicle Environment
Figure 2-8: How wheel and tyre models can come into a model context.
2.2.1.5.1 (Wheel and Tyre) Design Parameters
We limit ourselves to today’s traditional pneumatic tyres. Then the design is captured by:
• Carcass/Material: Rubber quality and plies arrangement.
• Tread/Grooves: Groove pattern, Groove depth, Tread depth, Spikes pattern (if spikes)
• Main dimensions: Outer radius, Width, Aspect ratio
• Installation parameters: Inflation pressure
2.2.1.5.2 Operational Parameters
These are operating conditions which vary slowly, and in this description assumed to be constant dur-
ing one manoeuvre/driving cycle. These are:
• Road surface (dry/wet, asphalt/gravel/snow/ice, …)
• Road compliance and damping (hard/soft, …)
• Wear state of tyre
• Age of tyre
• Temperature
2.2.1.5.3 Operational Variables
These are operating conditions which vary quickly, and in this description assumed to be variable in
time during one manoeuvre/driving cycle. These are:
• Tyre velocities ( ) or Tyre slip (𝑠 and 𝑠 (or ) and sign( ))
• Vertical tyre force, ( )
• Wheel torque (𝑇 𝐴 𝑛 𝑇 𝑝 𝑙 𝑛 +𝑇 𝑘 )
• Camber angle
The wheel’s rotational velocity, , can be contained in Wheel and Tyre model instead of inside Model of
rest of the vehicle, which would mean that will not be part of the Operational variables interface.
Tyre forces in road plane ( ) can be given instead of tyre slip. Another alternative is to give the
corresponding actuation, e.g. wheel shaft torque and wheel steer angle relative to wheel course angle.
With those setups, the response variables in 2.2.1.5.4.
Vertical tyre force can be modelled as arbitrarily varying in time or as an offset amplitude for different
frequencies where offset is from a mean value. The mean value would then be a parameter instead of a
variable. The latter alternative can be more efficient if simulating a longer driving cycle, where follow-
ing each road wave would be very computational inefficient.
2.2.1.5.4 Response Variables
Response refers to response to Operational variables changes.
Physical Response Variables
The variables from tyre model to model of rest of the vehicle is essential the forces and moments, see
Figure 2-6:
94
Vehicle Interactions and Subsystems
vehicle, wheel 𝑭𝒉 𝒃𝒛
including torque source, excluding wheel
𝑻
𝝎
𝑭𝒉 𝒃𝒙
𝒗𝒙
𝑻 𝒗𝒙 𝑭𝒉 𝒃𝒙
𝑭𝒉 𝒃𝒛
𝑒
𝑅𝑙
𝑛
𝑛
ground
95
Vehicle Interactions and Subsystems
w w
𝑅𝑙2
𝑅𝑙
𝑅 inflated unloaded radius 𝑅 free radius when loaded 𝑅2 free radius when loaded and propelled
𝑅 𝑅𝑙 lever when loaded 𝑅𝑙2 lever when loaded and propelled
𝑅 > 𝑅 and 𝑅2 < 𝑅 and
≈ 𝑅 ≈ 𝑅𝑙 ≠ 𝑅2 ≠ 𝑅𝑙2
Figure 2-10: Radius and speed relations of a tyre. 𝑅 and 𝑅𝑙 are not the same across a), b) and c).
There is a relative speed between tyre and the road surface. The ratio between this relative speed and
a reference speed is defined as the “tyre slip”. The reference speed can be the circumferential speed or
the translational speed of the tyre, depending on the application. For a driven wheel, the slip definition
in Eq [2.1] is often used and for braked wheels Eq [2.2] is often used. This is to avoid division by small
numbers in take-off and brake tests, respectively. The physical model in 2.2.3.1 reveals that Eq [2.1] is
the physically most motivated.
𝑅∙ 𝑅∙
𝑠 [2.1] 𝑠 [2.2]
|𝑅 ∙ | | |
With the definition in Equation [2.1] and a model for how tyre longitudinal force varies with 𝑅 and
(Eq [2.16] taken as example) we get Figure 2-11, which shows cuts in Figure 2-31.
+
𝑠 /
-40 40 +
/𝑁
-40 40 +
𝑅 / /𝑠
Figure 2-11: Slip 𝑠 and force as function of 𝑅 for constant . 50 [𝑘𝑁⁄ ] and
5000 [𝑁]. Semi-inverted scales around 𝑅 | | 0 and 𝑠| | . Dashed shows where definition
of 𝑠 is not relevant, since is not dependent of 𝑠 , since < 0.
It is not obvious which 𝑅 to use in Equations [2.1]..[2.2], e.g. 𝑅 𝑅 𝑅2 𝑅𝑙 or 𝑅𝑙2 . However, this com-
pendium recommends the free radius (𝑅 or 𝑅2 ), rather than the loaded radii (𝑅𝑙 or 𝑅𝑙2 ) because the
free radii are better average value of the radius around the tyre and the tyre’s circumference is tangen-
tially stiff, so speed has to be same around the circumference.
96
Vehicle Interactions and Subsystems
Sometimes one defines the Rolling radius 𝑅 ( ⁄ )| 𝑇= , i.e. a speed ratio with dimension length,
between translational and rotational speeds, measured when the wheel is undriven (𝑇 0). This ra-
dius can be used for relating vehicle longitudinal speed to wheel rotational speed sensors, e.g. for
speedometer or as reference speed for ABS and ESC algorithms.
Yet another approach is to use the radius ( ⁄ )| = , i.e. the ratio when the wheel is pure rolling. Us-
ing 𝑅 ( ⁄ )| 𝑇= or 𝑅 ( ⁄ )| = in the slip definitions shifts (𝑠 ) curve and 𝑇 (𝑠 ) curve, see
more in 2.2.3.5.
The variable 𝑠 is the longitudinal slip value, sometimes also denoted as 𝜅 or 𝜆. When studying brak-
ing, one sometimes uses the opposite sign definition, so that the numerical values of slip becomes posi-
tive.
𝛾 𝑅𝑧 0.15 0.00
0 4000 8000
0.10
𝑅𝑓 tyre.Fz [N]
𝑅𝑓
0.05
𝑭𝒛 𝐿/ 20
tyre.gamma
𝑅 𝑧
0.00
0
-0.05
0 2000 4000 6000 8000 0 4000 8000
tyre.Fz [N] tyre.Fz [N]
𝛾 6000 0.2
ground 0.0
level 4000
0 4000 8000
tyre.Fz [N]
𝑭𝒛 2000
𝐿
tyre.Rfree tyre.RFz
0.3004
0
0.3000
Figure 2-12: Left: Physical model. Right: Example result of model defined by Eqs [2.3] and [2.5].
A typical usage of this model is to know variable and parameters 𝑅𝑓 and . This leaves 4 un-
known variables (Δ 𝛾 𝐿 𝑅 𝑧 ) and equally many equations. An explicit solution is easy for Δ , but diffi-
cult for the remaining 3 variables [𝛾 𝐿 𝑅 𝑧 ]. For a single operation conditions, one can solve by itera-
tion. For whole simulations, it is more computational efficient to pre-process numerically in a table for
how [Δ 𝐿 𝑅 𝑧 ] varies with 𝛾 (for certain 𝑅𝑓 , which typically not varies during a simulation). For
each time instant during simulation, interpolation in this table gives [𝛾 𝐿 𝑅 𝑧 ] from known Δ .
The stiffness can be found for an existing tyre with a certain inflation pressure, but it is far from a
design parameter. Also, the linear model is often too simplified. To improve this, alternative vertical
deformation models can be used. If studying only single operation conditions, Finite Element models
are suitable. But for simulations, a more computational efficient solution is almost a necessity. The
97
Vehicle Interactions and Subsystems
model in Eq [2.5] can replace Eq [2.4]. This is simple but indicates the mechanisms for how sidewall
stiffness and inflation pressure come into play: Assume that the vertical load is carried by sidewalls
( 𝑤 ) and the belt area between the sidewalls ( 𝑙 ). Assume belt has no bending stiffness and side-
walls follow Hertz contact theory:
𝑤 + 𝑙
𝑓 𝑇⁄
≈{ }≈ | | This definition includes both force and speed losses.
≈0 𝑧
𝑇 𝐿 𝑇 𝑇⁄
• Torque (or Force) definition: 𝑅𝑅 sign( ) sign( ) This def-
𝑧 𝑧 𝑧
inition only includes force losses relative to a nominal force 𝑇/𝑅; not velocity or speed
losses.
The 𝑅𝑅 (or 𝑓 ), is Rolling Resistance Coefficient.
Energy definition 𝑅𝑅 𝐸 has the advantage that it does not require the radius 𝑅. But it is not useful in
most vehicle dynamics models, since it does not resolve into force and velocity. And, the 𝑅𝑅 𝐸 gener-
ally varies more than 𝑅𝑅 . So, the compendium uses the force definition. It should be mentioned that
(ISO28580, 2018) uses the Energy definition.
§ The 𝑅𝑅 is dimensionless. With the energy definition, we can understand it as a normalized energy
loss and with force definition we can understand it as a normalized torque loss. In both cases
[𝑁 ⁄(𝑁 )], but [𝑁 ] is energy and torque, respectively.
The speed losses are governed by the following equation, which is explained more in 2.2.3:
𝑛𝑜 𝑡𝑒- 𝑅
• {
𝑟 𝑠 𝑝} 𝑥( 𝑥 ) { 𝑥 ≪ } 𝑥 𝑠𝑥 { ≈ 𝑁𝑜 } 𝑥 𝑠𝑥 |𝑅 |
𝑥
98
Vehicle Interactions and Subsystems
The overall explanation of rolling resistance for pneumatic tyres on hard flat surfaces is that the pres-
sure distribution is offset in rolling direction. In Figure 2-13, radial damping and friction is the cause
for the offset. Figure 2-13 also shows another off-set effect; that the whole contact patch is moved for-
ward due to longitudinal force and shearing of the sidewalls.
The rolling resistance coefficient is almost the same for very low speeds; even when the wheel starts
rolling from zero speed, after gradually increasing the torque up to the rolling resistance moment. The
radial friction in Figure 2-13 can explain that, but there is also another explanation, see Figure 2-14. It
explains an additional reason for why the contact patch is moved ahead of wheel hub. The belt is cir-
cumferentially stiff and takes a short-cut along the chord, through the contact. This builds up shear
stresses in the sidewall, 𝜏 𝑤 . The belt is flexible for bending, so belt radius is proportional to belt ten-
sion force, . This is because same effect as for tension in rope around a cylinder: 𝑓𝑜𝑟 𝑒 𝑝𝑟𝑒𝑠𝑠𝑢𝑟𝑒
𝑟 𝑢𝑠 𝑡ℎ . In our case, the pressure corresponds to the summed effects of inflation pressure 𝑝 𝑛𝑓𝑙
and radial stress in sidewalls 𝜎 𝑤 . So, the radius becomes smaller than original radius in inlet and
larger in outlet. Assumption of constant contact length and geometric constraints from tensional rigid
belt requires that contact patch is offset towards the inlet, so 𝑒 sign( ) |𝑒|;.
Pressure distribution components Resulting pressure and Equivalent forces
𝑇 𝑇
Rl Rl
Pressure from elastic
radial deformation e
damping Pressure from radial
deformation speed
friction
pressure
Contact patch offset, due to shear of
tyre sidewalls (drawn for > 0) equivalent
Figure 2-13: Normal force distribution on a tyre. The measure e is the force offset. In steady state,
the forces in hub and contact patch are the same: and .
99
Vehicle Interactions and Subsystems
undeformed
Rl deformed 𝑤
𝜏 𝑤
𝑟 𝑢𝑠 2 > 𝑟 𝑢𝑠 < 𝑤+ 𝑤+
+
𝑤
Figure 2-14: 1. Wheel at low speed ( small and negative). 2. Free body diagrams with belt
separate. 3. How belt tension force changes before and after contact. 4. Influence of radii variation
on contact patch position.
denotes the longitudinal force on the wheel, 𝑇 denotes the applied torque and 𝑅 denotes the tyre
radius. For a free rolling tyre, where 𝑇 0, 𝑓 becomes simply ⁄ . One often sees definitions of
𝑅𝑅 which assumes free rolling tyre; but Eq [2.6] is also valid when 𝑇 ≠ 0, which is useful.
A free body diagram of the forces on the wheel can be used to explain the rolling resistance. Consider
Figure 2-13 which represents a free rolling wheel under steady state conditions. The inertia of the
wheel is neglected.
Longitudinal and vertical force equilibria are already satisfied, due to assumptions above. However,
moment equilibrium around wheel hub requires:
𝑇 𝑒
𝑇 ∙ 𝑅𝑙 ∙𝑒 0 ⇒ [2.7]
𝑅𝑙 𝑅𝑙
This result suggests that the force , which pushes the vehicle body forward, is the term 𝑇⁄𝑅𝑙 (arising
from the applied torque T) minus the term ∙ 𝑒⁄𝑅𝑙 . The term can be seen as a force 𝑙𝑙 and referred
to as the rolling resistance force. We seldom know neither 𝑅𝑙 or 𝑒, but they are rather constant and the
form of Eq [2.7] is same as Eq [2.6] if:
𝑅𝑅 |𝑒⁄𝑅| [2.8]
Eq [2.8] is a definition of rolling resistance coefficient based on physical mechanisms internally in the
tyre with road contact, while Eq [2.6] is based on quantities which are measurable externally, see Eq
[2.6]. Sometimes one sees 𝑅𝑅 ⁄ as a definition, but that is not suitable since it assumes ab-
sence of torque.
100
Vehicle Interactions and Subsystems
It is not obvious which radius to use. 𝑅𝑙 is can be motivated because 𝑅𝑙 is the lever for . However, in
2.2.1.7, we found arguments for using other radii. If same radius is used, for slip and rolling resistance,
we can fully see the tyre as a transmission with the nominal ratio 𝑅 and zero energy loss if 𝑅𝑅 0
and 𝑠 0.
It is important to refer to this phenomenon as rolling resistance as opposed to rolling friction. It is
not friction in the basic sense of friction, because ≠ 𝑅𝑅 ∙ except for the special case when un-
driven wheel (𝑇 0). Figure 2-15 shows an un-driven and pure rolling.
Rolling resistance is a torque loss. Other torque losses, which can be included or not in tyre rolling re-
sistance, are:
• losses associated with friction in gear meshes,
• drag losses from oil in the transmission,
• wheel bearing (and bearing sealings) torque losses,
• drag from brake discs,
• drag losses from aerodynamic around the wheel, and
• uneven road in combination with suspension damping that dissipates energy.
These should, as rolling resistance, be subtracted from propulsion/brake torque. However, sometimes
they are included as part of the tyres rolling resistance coefficient, which can be misleading. The wheel
bearing torque loss have two torque terms: one is proportional to vertical load on the wheel (adds typ-
ically 0.000 to rolling resistance coefficient), and the other is of constant magnitude but counter-di-
rected to rotation speed. The former term can be included in rolling resistance coefficient. The aerody-
namic losses due to wheel rotation are special since they vary with wheel rotational speed, meaning
that they (for constant vertical load) are relevant to include when studying the variation of rolling re-
sistance coefficient with vehicle speed. A summarizing comment is that one has to be careful with
where to include different torque losses, so that they are included once and only once.
101
Vehicle Interactions and Subsystems
e e
e
Figure 2-15: Driven wheel with rolling resistance. Special cases “Free-rolling” and “Pure rolling”.
Operational parameters, see Figure 2-8:
• Elevated temperatures give low rolling resistance, via increased inflation pressure. Tyres need
to roll approximately 30 km before the rolling resistance drop to their lowest values.
• Road/ground, sometimes covered: Clean asphalt, Asphalt with water/leaves/sand/…,
Loose/hard gravel, Snow/ice. Soft ground or covered hard ground increases rolling resistance.
• Wear state. Worn tyres have lower rolling resistance than new ones (less rubber to deform).
Operational variables, see Figure 2-8:
• Vertical force.
• Speed. Rolling resistance increases with vehicle speed due to rubber hysteresis and air drag.
• Tyre loads (propulsion/braking and lateral forces)
Approximately for modern tyres (written year 2020): Passenger car tyres have 𝑅𝑅 0.005. .0.0 ,
Truck tyres have 𝑅𝑅 0.00 . .0.005. The lower values are “eco-tyres” and can have some trade-off
with vertical comfort and road grip.
2.2.2.4.1 Variation of Tyre Type
Trucks tyres have a much lower rolling resistance coefficient than passenger vehicle tyres, approxi-
mately half. Tyres have developed in that way for trucks, because their fuel economy is so critical.
2.2.2.4.2 Variation of Vertical Force
In a first approximation, the rolling resistance force is proportional to vertical force, i.e. RRC is con-
stant. But, typically, the RRC decreases slightly with vertical force. This, and parasitic bearing losses,
𝑡𝑜𝑟𝑞𝑢𝑒 sign( ) 𝑜𝑛𝑠𝑡 𝑛𝑡, explains why commercial vehicles lift axles when driven with low pay-
load.
2.2.2.4.3 Variation of Speed
As an example, left part of Figure 2-16 shows the influence of speed and tyre construction on rolling
resistance. The sudden increase in rolling resistance at high speed is important to note since this can
lead to catastrophic failure in tyres. The source of this increase in rolling resistance is a high energy
standing wave that forms at the trailing edge of the tyre/road contact.
102
Vehicle Interactions and Subsystems
There are some empirical relationships derived for the tyre's rolling resistance. It is advisable to refer
to the tyre manufacturer's technical specifications when exact information is required. This type of in-
formation is usually very confidential and not readily available. Some general relationships have been
developed, from (Wong, 2001):
Radial-ply passenger car tyres: 𝑅𝑅 0.0 6 + 0.04 ∙ 0 6 ∙ 2
Bias-ply passenger car tyres: 𝑅𝑅 0.0 69 + 0. 9 ∙ 0 6 ∙ 2
Radial-ply truck tyres: 𝑅𝑅 0.006 + 0. ∙ 0 6 ∙ 2
Bias-ply truck tyres: 𝑅𝑅 0.007 + 0.45 ∙ 0 6 ∙ 2
As seen in Figure 2-16, a rule of thumb is that rolling resistance coefficient is constant up to around
100 km/h.
.05
0.4
.04
Sand
.03
0.2
103
Vehicle Interactions and Subsystems
𝑇 𝑣 𝑇 𝑣 𝑇
horizontal=
=vehicle-
𝑣 𝑣
longitudinal
𝑣
𝑒
𝑣
Figure 2-17 : Rolling resistance explanation for hard and soft ground. Subscripts: c=contact,
v=vehicle.
2.2.2.4.5 Variation of Longitudinal Force, Propulsion and Braking
Figure 2-13 shows an offset of contact patch due to longitudinal force . This makes 𝑅𝑅 dependent
of . The phenomena is not well studied, since 𝑅𝑅 is often given for free-rolling tyres, but according
to Figure 2-18 RRC increases with positive and decreases with negative . The wheel radius also
decreases with magnitude of force. For negative forces, these two effects have opposite influence, so
the change is less for negative force, as seen in Figure 2-18.
Rolling resistance coefficient,
𝑇⁄
𝑅𝑅
𝑧
0.08
Normalised
0.04
Longitudinal
Force,
𝑧
-0.4 -0.2 0 +0.2 +0.4
Figure 2-18 : Rolling resistance dependency of longitudinal tyre force. Inspired by (Wong, 2001)
2.2.2.4.6 § Separation of Actuated Wheel Torque and Rolling Resistance
Contribution from Toheed Ghandriz, Vehicle Dynamics
The force play in FBD numbered 3 in figure below is primarily proposed for understanding. But there
are sometimes reasons to use FBD numbered 4, where the torques 𝑇 and 𝑛 𝑒 from Figure 2-9
are treated separately and converted to one force each. Such reason is that 𝑇 𝑇𝐴 is possible to con-
trol (actuatable) while 𝑛 𝑒 𝑇 is not.
1 𝑇𝐴 2 3 4
𝑇 𝑝 +𝑇 𝑘 𝑇𝐴 𝑇𝐴 𝑇 𝑇𝐴
𝑇 𝑅
𝑅𝑅 𝑅 𝑇 𝑅 𝑅
𝑒 𝑅𝑅 𝑅 𝐴 𝑇𝐴 /𝑅
Figure 2-19: § Force equivalent FBDs with both rolling resistance (RR) and actuation (Act). Wheel
is modelled inertia free in both translation and rotation. 𝐴 and 𝑅 is a virtual decomposition
of the real force .
The decomposition of in 𝐴 and 𝑅 means a deviations from the advice in 1.5.2.1 to not draw
two forces acting at the same point of part and having same direction. Then note:
104
Vehicle Interactions and Subsystems
• The tyre models where force depends on slip (2.2.3..2.2.6) applies to , not to 𝐴 .
• The 𝑅 is not a separate physical force acting on the vehicle. It is 𝐴 𝑅 that
physically acts on the vehicle and should appear in vehicle equilibrium.
So, 𝑅 is very different from aero dynamic resistance force and grade resistance forces, which are
physical forces acting on the vehicle.
+ 𝑙
global horizontal
Figure 2-20: § Explanation of rolling resistance as wheel rolling in a “local uphill”. The “global
uphill” or “road” is marked to distinguish between and .
2.2.2.5.1 Uneven and Rigid Ground
Force Losses
Moment equilibrium of the wheel gives:
0 𝑇 𝑅 𝑒
If we model the “local”, or contact-local, rolling resistance as on rigid ground, 𝑒 𝑅𝑅 𝑅, we get:
𝑓𝑜𝑟 𝑒
0 𝑇 𝑅 𝑅𝑅 𝑅 ⇒{ }⇒
𝑒𝑞𝑢 𝑒𝑛 𝑒
⇒ 0 𝑇 ⁄𝑅 ( 𝑣 cos( ) + 𝑣 sin( )) ( 𝑣 cos( ) 𝑣 sin( )) 𝑅𝑅 ⇒
⇒ 0 𝑇 ⁄𝑅 + 𝑣 ( cos( ) + sin( ) 𝑅𝑅 ) + 𝑣 ( sin( ) cos( ) 𝑅𝑅 ) ⇒
𝑇𝑖 ⁄ 𝑖𝑧𝑣 ( in( 𝜑𝑖 𝑦𝑟 )+co ( 𝜑𝑖 𝑦𝑟 ) ) 𝑠 𝑇𝑖 ⁄ 𝑖𝑧𝑣 ( 𝜑𝑖 𝑦𝑟 + )
⇒ 𝑣 co ( 𝜑𝑖 𝑦𝑟 ) in( 𝜑𝑖 𝑦𝑟 )
≈ {| |} ≈ +𝜑𝑖 𝑦𝑟
≈
𝑠 || 𝑛
{ } ≈ 𝑇 ⁄𝑅 𝑣 ( + 𝑅𝑅 )
𝑠 |𝑅𝑅 |
One can define a rolling resistance in the vehicle direction, 𝑅𝑅 𝑣:
105
Vehicle Interactions and Subsystems
𝐿 𝑇𝑖 ⁄ 𝑖 𝑣
𝑇𝑖 ⁄ (𝑇𝑖 ⁄ 𝑖𝑧𝑣 ( 𝜑𝑖 𝑦𝑟 + ))
𝑅𝑅 𝑣 ≈ ≈ 𝑅𝑅 +( )
𝑉 𝑙 𝑖𝑧𝑣 𝑖𝑧𝑣
Note that negative is a “local downhill”, so 𝑅𝑅 𝑣 is the sum of two positive values if “local uphill”,
which should be intuitive correct. If the ground is uneven and rigid, there are equally much local
downhill as local uphill on average; also if the road has a “global grade”. Then, on average, there is no
influence of , so 𝑅𝑅 𝑣 ≈ 𝑅𝑅 .
Speed Losses
It should be intuitive to use the longitudinal slip model, see 2.2.3, in the contact directions:
𝑠 𝑠 . Inserting the slip definition 𝑠 (𝑅 )⁄|𝑅 | gives:
𝑅 𝑓𝑜𝑟 𝑒
⇒{ }⇒
|𝑅 | 𝑒𝑞𝑢 𝑒𝑛 𝑒
⇒( 𝑣 cos( )+ 𝑣 sin( ))
𝑅 ( cos( ))
( 𝑣 cos( ) 𝑣 sin( )) ⇒
|𝑅 |
𝑠 𝑅 ( )
⇒ {| |} ⇒ ( 𝑣 + 𝑣 ( )) ≈ ( 𝑣 𝑣 ( )) ⇒
|𝑅 |
⇒ 𝑣 𝑣 ≈ ( 𝑣+ 𝑣 ) 𝑠 ⇒
⇒ 𝑣 𝑣 𝑠 ≈ 𝑣 𝑠 + 𝑣 ⇒
𝑠 + 𝑠 | 𝑠 |
⇒ 𝑣 ≈ 𝑣 { } 𝑣 ( 𝑠 + )
𝑠 𝑛 𝑠 | |
Note that negative is a “local downhill”, so helps in creating 𝑣 which reduces the re-
quired 𝑣 and, hence the required 𝑇 . If the ground is uneven and rigid, there are equally much local
downhill as local uphill on average; also if the road has a “global grade”. Then, on average, there is no
influence of , so 𝑣 ≈ 𝑣 𝑠 .
The same modelling concept can be used with saturated sign(𝑠 ) min( |𝑠 | ) and
then one can study traction problems due to uneven ground; i.e. when the vehicle gets stuck.
2.2.2.5.2 Losses in Suspension Damping
Above concluded that uneven and rigid does not cause energy losses on average. However, unevenness
can still cause energy loss due to increased motion in vehicle suspension dampers.
2.2.2.5.3 Deformable Ground
If the ground is deformable (soft) it deforms due to the force from the tyre. The ground is generally
more plastic than elastic, so a simplification is that deformation is completely plastic.
106
Vehicle Interactions and Subsystems
𝑁 𝑁
𝑁 𝑁 (𝑁 )
𝑁 𝑝 𝑘 𝑁
slide 𝑁
𝑛 𝑁 𝑛
stick
𝑠 𝑝 ⁄ 𝑓
𝑅 ⁄ 𝑓
slide
• A stick mode where Δ 0. • No stick mode (if rolling, i.e. > 0).
• In slide mode, the friction force, , • The friction force, , is depending on
is only depending on sign Δ . relative sliding speed, ⁄ 𝑓 .
w*R
vx
bristle or block
Figure 2-22: Tyre ground contact for braked tyre. Origin to the “Brush model”. Picture from
Michelin.
2.2.3.1.1 Uniform Pressure Distribution and Known Contact Length
A first simple variant of the brush model, uses the following assumptions:
o Sliding and shear stress only in longitudinal direction (as opposed to combined longitudinal
and lateral)
o Uniform and known pressure distribution over a constant and known contact length (as op-
posed to using a contact mechanics-based approach, which can calculate pressure distribution
and contact length 𝐿.)
107
Vehicle Interactions and Subsystems
𝜉
pressure 𝑅
stick zone stick
𝜉
slip
slip zone
“mode”
𝜉
shear stress 𝜏
(or shear deformation angle, 𝛾)
Figure 2-23: Physical model for simple brush model for longitudinal slip. Drawn for propelled tyre.
The bristles represent the rubber tread, not including elasticity in sidewall. Drawn for 𝑅 ∙ > .
If we assume no sliding:
𝜏 ∙𝛾 ℎ𝑒𝑟𝑒 𝜏 𝑠ℎ𝑒 𝑟 𝑠𝑡𝑟𝑒𝑠𝑠 𝑠ℎ𝑒 𝑟 𝑜 𝑢 𝑢𝑠 𝛾 𝑠ℎ𝑒 𝑟 𝑒𝑓𝑜𝑟 𝑡 𝑜𝑛 𝑛 𝑒
When a bristle enters the contact patch, it lands un-deformed, i.e. with 𝛾 0. The further into contact,
along coordinate 𝜉, we follow the element, the more sheared will it become. Since the ground end of
the element sticks to ground, the increase becomes proportional to the speed difference and the
transport time 𝑡 𝑛 𝑝 , which is the time it takes for a bristle to reach the coordinate 𝜉:
= 𝑟 𝑝
∫= ∙ 𝑡 Δ 𝑅∙ 𝑟 𝑝 𝜉 𝑛 𝑝 𝑡 𝑛𝑝
𝛾 { } ∫ 𝑡 { }
𝐻 𝑅∙ 𝐻 = 𝑛 𝑝 |𝑅 ∙ |
𝜉 ⁄| 𝜔| 𝜉 ⁄| 𝜔|
𝑅∙ 𝑅∙ 𝑅∙ 𝜉 𝑅∙
∫ 𝑡 ∫ 𝑡 ∙ ∙𝜉
𝐻 = 𝐻 = 𝐻 |𝑅 | 𝐻 ⏟|𝑅 ∙ |
108
Vehicle Interactions and Subsystems
𝐿 𝐿 𝐿
𝑊
∫𝜏∙ 𝐴 𝑊∙ ∫ 𝛾∙ 𝜉 𝑊∙G ∫ ∙𝑠 ∙𝜉∙ 𝜉 ∙ 𝑠 ∙ ∫𝜉 ∙ 𝜉
𝐻 𝐻
𝑊𝐿 ξ= ξ=
𝑊 𝐿2
∙𝑠 ∙𝑠 (𝑁𝑜𝑡𝑒: 𝑇ℎ 𝑠 𝑠𝑠𝑢 𝑒𝑠 𝒏𝒐 𝒍𝒊𝒅𝒊𝒏𝒈 𝑛 𝑜𝑛𝑡 𝑡. )
⏟ 𝐻
Note that this model simplifies from a function of 2 variables ( and ) to 1 variable (𝑠 ). With this
definition of slip, 𝑠 , we automatically handle and both positive and and both negative. The
case when and have different signs ( < 0) is special and will be handled below.
(It is possible to reach Eq [2.9] via integration over 𝜉 instead. Then, one uses 𝑛 𝑝 | |. After a
more complex derivation, the resulting slip definition 𝑠 (𝑅 ∙ )⁄|𝑅 ∙ | falls out in the final ex-
pression for , i.e. Eq [2.9]. Despite this, one often sees 𝑠 (𝑅 ∙ )⁄| |.)
As long as friction limit is not reached (|𝜏| < 𝑝) within the whole contact, this model is valid. The
rsik for sliding is highest at 𝜉 𝐿 so the validity can be formulated |𝜏(𝐿)| < 𝑝 𝑊𝐿
𝑧
⇒ 𝐻∙
|𝑠 | ∙ 𝐿 < 𝑧
⇒ |𝑠 | < 𝑧
⇔ | |< 𝑧
.
𝑊𝐿 2 2
So, if that condition is not fulfilled, the friction limit is reached within the contact, i.e. at a break-away
point at 𝜉 𝜉 < 𝐿. Then, we have to split the integral in two. The point 𝜉 is defined by 𝜏(𝜉 ) 𝐻 𝑠
2 𝑝 𝑊 𝐿2 𝐿
𝜉 𝑊 𝐿2
𝑠 𝜉 𝑝 ⇒ 𝜉 2 2
𝑧
. For 𝜉 > 𝜉 , the rubber element will slide with a
constant 𝜏 ∙ 𝑝.
𝐿 𝜉 𝐿
The force terms, 𝑘 and 𝑙 𝑝 , from each of stick and slip regions are identified. These two terms
are separately shown plotted in Figure 2-25.
The case when and have different signs leads to that 𝜉 0 , since the bristles will deform in the
opposite direction; slip over the whole contact or macro slip. So, the whole contact has 𝜏 ∙ 𝑝, which
leads to . Hence, the total expression for becomes as in Eq [2.9]. We also add subscript 𝑥
on and 𝐻, to prepare for a corresponding model for lateral forces, in 2.2.4.
sign(𝑅 )∙ ∙ sign(𝑠 ) ∙ ∙ 𝑓 <0
∙ ∙
∙𝑠 𝑒 𝑠𝑒 𝑓 |𝑠 | ≤ ⇔| |≤
∙
∙ [2.9]
sign(𝑠 ) ∙ ∙ ∙( ∙ ) 𝑒 𝑠𝑒
{ 4∙ |𝑠 |
∙ 𝑊 ∙ 𝐿2 𝑅
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒] 𝑛 𝑠
∙𝐻 |𝑅 |
Eq [2.9] is plotted in Figure 2-25. Eq [2.9] is also called Dugoff tyre model. It exists also as lateral slip
and combined slip tyre model, see (Dugoff, Fancher, & Segel, 1969).
It is important to reflect over which of the physical quantities that reasonably has to be modelled as
varying. This will of course depend on the driving manoeuvre studied, but here is a typical situation:
The slip 𝑠 and normal load are typical varying and defined by the vehicle model. The quantities
109
Vehicle Interactions and Subsystems
𝑊 𝐻 are often reasonably constant, so they can be parameters. However, it is often not reason-
able to assume that the contact length 𝐿 is constant. 𝐿 is rather a function of : 𝐿( ) .
The case when < 0 is unusual and can only occur when |𝑠 | > . An example is when vehicle
moves rearward with /𝑠, would be that the wheel spins forward, e.g. with 𝑅 /𝑠.
Then 𝑠 + and ∙ . Also, if increase to 𝑅 + /𝑠, we get same ∙ , but
with 𝑠 + . One can also note that there is another 𝑅 , for same , which gives 𝑠 + . This is
𝑅 ⁄ /𝑠 and then ∙ ∙( ∙ ⁄(8 ∙ )), which is < . So, is uniquely de-
fined for any ( ), but has double solutions for some 𝑠 , when |𝑠 | > .
If we instead hold a certain forward vehicle speed, e.g. /𝑠, and study how varies with , we
can identify 4 specific levels of :
• Full rearward traction: 𝑅 ⁄𝑠. This gives 𝑠 , with different signs on and ,
and
• Locked wheel: 𝑅 0 ⁄𝑠. This also gives 𝑠 and
• Pure rolling wheel: 𝑅 ⁄𝑠. This gives 𝑠 0 and 0;
• Full forward traction: 𝑅 + ⁄𝑠. This gives 𝑠 + and ∙ ∙(
∙ ⁄(4 ∙ )) ≈ {𝑡𝑦𝑝 𝑦} ≈ (0.95. .0.98) ∙ , which is <
The first case is achievable, with an electric motor braking but not with friction brakes, where and
have different signs. The last case shows that we cannot reach in the direction the vehicle
moves, since there will always be a small part on the inlet side of the contact where the shear stress
has not reached 𝑝. In practice, we can see it as ≈ , but when using the model mathematically, it
can be good to note such small phenomena.
2.2.3.1.2 Longitudinal Tyre Slip Stiffness
In summary for many models, the following is a good approximation (compared to tests) for small lon-
gitudinal slip, and certain normal load and certain friction coefficient:
∙𝑠 [2.10]
For the brush model, or any other model which describes (𝑠 … ), one can define the “Longitu-
dinal tyre (slip) stiffness” , which have the unit 𝑁 𝑁⁄ 𝑁/ (( /𝑠)⁄( /𝑠)). It is the derivative of
force with respect to slip when at 𝑠 𝑠 0:
𝜕
(𝜕 (𝑠 … ))| ( …) [2.11]
= 𝑦=
Note that is not a stiffness in the conventional sense, force/deformation. The tyre also has such a
deformation stiffness. Often, it is obvious which stiffness is relevant, but to be unambiguous one can
use the wording: “slip stiffness” [N/1=N/((m/s)/(m/s))] and “deformation stiffness” [N/m].
With the brush models with both pressure distributions, Eq [2.9] and Eq [2.16], we get the
∙ 𝑊 ∙ 𝐿2 ⁄( 𝐻). With 0.5 [𝑀 𝑁⁄ 2 ] (typical shear modulus in rubber), 𝑊 𝐿 0. . .0. [ ]
(typical sizes of contact patch for passenger car) and 𝐻 0.0 [ ] (approximate tyre tread depth) one
gets ≈ 5[𝑘𝑁] < ∙ 𝑊 ∙ 𝐿2 ⁄( 𝐻) <≈ 40[𝑘𝑁/ ]. Empirically, we can measure for passenger
car tyres around 5. .50 [𝑘𝑁/ ]. This indicates that the brush model models the essential physical phe-
nomena and that the sheared part (the bristles) is rather only the tread than the whole elastic part
sidewall and tread together.
2.2.3.1.3 Influence of Vertical Load and Friction in Brush Model
The vertical load on the tyre affects the force generation, . For large slip our dry friction model moti-
vates that is proportional to . This is also a good approximation for small slip, via . Then,
one can define the Longitudinal Slip Stiffness Coefficient, :
𝑠 𝑠 [2.12]
110
Vehicle Interactions and Subsystems
The contact length 𝐿 will reasonably vary according to some deformation model. Hertz’s contact the-
ory for line contact motivates that 𝐿 2√ . This is implemented as 𝐿2 /𝑘, where 𝑘 is a material
modulus with dimension force/area, gives:
∙ 𝑊 ∙ 𝐿2 ∙ 𝑊 ∙ /𝑘 ∙𝑊
( )
∙𝐻 ∙𝐻 ⏟∙ 𝐻 𝑘
So, the assumption 𝐿 2√ explains why ≈ . This can be verified with experiments,
see Figure 2-55. A small tendency for degressive increase (𝜕 2 ⁄𝜕 2 < 0) can be found. A definition of
from a general ( 𝑠 ) will include , which is the vertical force around which the model
should be valid:
( ( ))| ( ) [2.13]
𝑧 = 𝑧0
The result is summarised in Figure 2-24. Increasing only increases the saturation level, while leaving
the slope at lower 𝑠 constant, or possibly slightly increased. Variation of involves the contact length
model. The assumption “𝐿 2√ ” simply scales the curve in force direction. Overall, it can be con-
cluded that there are arguments for two conceptual ways how the tyre characteristics changes, varied
vertical force and varied friction coefficient, see Figure 2-24.
Air
Asphalt temperature:
-15 to -10 C
=
⁄
Normalized force,
Snow
=
Reference 𝑠
Doubled vertical force (simplified) Ice
Doubled vertical force (tendency)
Doubled friction coefficient (simplified)
Doubled friction coefficient (tendency)
slip, 𝑠
Figure 2-24: How tyre characteristics typically vary due to varying vertical force and road friction.
Left: Theory. Right: Measurements with varying friction, i.e. varying surfaces, from PhD course by
Ari Tuononen, Aalto university, Finland, Saariselkä, Finland 2014-03-15..22.
The influence of vertical force on (and on lateral slip stiffness ) is further discussed in 2.2.4.3 and
2.2.5.4.
2.2.3.1.4 Influence of Different Static and Dynamic Friction
A common model for friction is that coefficient of friction to remain sticking, 𝑘 , is higher than the
coefficient of friction when slipping has started, 𝑙 𝑝 . This is sometimes called “stiction”, 𝑘 𝑛
𝑘 ⁄ 𝑙 𝑝 . If this is implemented in the model, the brush model can explain why the overall (𝑠 )
often has a peak, as indicated already in Figure 2-21. In the derivation in Eq Error! Reference source
𝐿
not found., the differing between 𝑘 and 𝑙 𝑝 , affects like this: 𝑙𝑝 𝑊 ∙ ∫𝜉 𝑙 𝑝 ∙ 𝑝 ∙ 𝜉 and
𝐿
𝜉 2
𝑖 𝑘 𝑧
.
111
Vehicle Interactions and Subsystems
⇔| |≤ ∙
{ ∙ 𝑒 𝑠𝑒
2
∙𝑊∙𝐿 𝑅
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒] 𝑛 𝑠
∙𝐻 |𝑅 |
The shape of this curve becomes as shown in Figure 2-25. The uniform pressure distribution model is
drawn as reference. Note that the parabolic pressure distribution does not give any linear part, but the
derivative at 𝑠 0 is same, ∙ 𝑊 ∙ 𝐿2 ⁄( 𝐻).
112
Vehicle Interactions and Subsystems
Figure 2-25: Shape of force-to-slip relation derived with brush model with different pressure
distributions. Also, the force terms from stick- and slip-regions are shown. The arctan and tanhyp
curves are examples of curve-fitting to same / 𝑠 in 𝑠 0.
Figure 2-25 indicates that tanhyp gives a better curve fit than arctan. However, that conclusion is
drawn without introducing different static and dynamic friction coefficients, see 2.2.3.1.6. Figure 2-26
shows that the curve shape before peak is influenced by how different 𝑙 𝑝 and 𝑘 are.
∙ |𝑠 | ( ) +( )
𝑘 ∙ 𝑘∙ 𝑘 7∙( 𝑘 )2
{ 𝑒 𝑠𝑒
∙ 𝑊 ∙ 𝐿2 𝑅
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒] 𝑛 𝑠
∙𝐻 |𝑅 |
Typical values of friction coefficients that give good resemblance with measurements are 𝑘 𝑛
. . . .0, see experiment with tyre tread sample in (Ludwig & Kim, 2017). Note that the peak of the
113
Vehicle Interactions and Subsystems
𝑝 𝑘
The model and analysis can be transferred to lateral slip, except that there is often a tendency that
⁄ ≠ , but rather that tyre_smallStiction.Fx
tyre_noStiction.Fx
⁄ decreases with . Hence,tyre2_noStiction.Fx
tyre_criticalStiction.Fx
𝑠 𝑝 𝑘 increases
tyre_largeStiction.Fx
tyre2_noStiction.Fx
with .
tyre2_smallStiction.Fx tyre2_criticalStiction.Fx
tyre2_smallStiction.Fx tyre2_criticalStiction.Fx tyre2_largeStiction.Fx
tyre2_largeStiction.Fx tyre2_noS
9000 9000
9000
𝑙𝑝 𝑘 .65 06 (tuned in)
8000 𝑙 𝑝 8000
8000
𝑘
7000 7000
7000
𝑘 𝑘
6000 6000
6000
𝑘 .5
𝑘
5000 5000
5000
𝑘
4000 4000
4000
𝑘 .5
3000 𝑘 3000
3000
2000 2000
2000
1000 1000
1000
0 0
0.0 0.2 0.4 0.6 0.8 0.01.0
0.0 0.2
0.2 0.4
0.4 0.6
0.6 0.80.8 1.01.0
sx sx [rad-1] sx
Figure 2-26: Brush model with uniform (left) and parabolic (right) pressure. Varying 𝑘/ 𝑙 𝑝.
114
Vehicle Interactions and Subsystems
2.2.3.2 Measurements
Some examples of measurements are given Figure 2-28. It is important to understand that there is a
big spread between different tyres (e.g. studded or not, as shown in the figure) and that the physical
phenomena we try to measure and model is not only the tyre, but the contact between one certain tyre
and one certain ground surface (e.g. ice, as shown in figure). Additionally, there is one certain wheel
suspension which can cause different high frequency oscillations which affect the averaged measured
signals. The figure shows both longitudinal and lateral grip, so both relevant for 2.2.3 and 2.2.4, but not
the combined situation in 2.2.5.
Straight-line braking Steering without wheel torque
𝑥/
/
utilized friction
utilized friction
0.05 Studded tyre
Studded tyre
Nordic non-studded winter tyre Nordic non-studded winter tyre
European non-studded winter tyre European non-studded winter tyre
115
Vehicle Interactions and Subsystems
Figure 2-30: Left: TMsimple (Lex, 2015). Right: TM-Easy Tyre Model, (Hirschberg, Rill, &
Weinfurter, 2007).
2.2.3.4.3 More Advanced Models
There are numerous of more advanced tyre models, such as Swift and FTire. They mix physical and
curve fit parameters. FTire is almost a finite element model and demands very many parameters.
2.2.3.4.4 Very Simple Tyre Models
There are many more models with different degree of curve fitting to experimental data. However, one
can often have use for very simple curve fits, such as:
116
Vehicle Interactions and Subsystems
• Linearized: ∙𝑠
• Linearized and saturated: sign(𝑠 ) ∙ 𝑛( ∙ |𝑠 | ∙ ). This means approximating the
curve _𝑥 𝑠 with 3 spliced lines with the 2 borders 𝑠
( ) ± ∙ ⁄ . One can also splice in
more lines, so that the model stays piecewise linear.
• Stiff: 𝑠 0 ⇔𝑅 (as if linear with → )
• Stiff and saturated: 𝑓 𝑠𝑡 𝑘 𝑠 0 𝑒 𝑠𝑒 Discrete state switching: ℎ𝑒𝑛 𝑠 <
0 𝑡ℎ𝑒𝑛 𝑠𝑡 𝑘 ← 𝑡𝑟𝑢𝑒 𝑒 𝑠𝑒 ℎ𝑒𝑛 > 𝑡ℎ𝑒𝑛 𝑠𝑡 𝑘 ← 𝑓 𝑠𝑒 (Approximately described.)
If wheel locked (e.g. by If vehicle stand-still (e.g. held If vehicle has constant non-zero velocity (e.g.
friction brake), i.e. 0: by other wheels), i.e. 0: controlled by other wheels), i.e. 𝑜𝑛𝑠𝑡 𝑛𝑡 ≠ 0:
>0 𝑅
𝑅
( < < 𝑅
𝑙𝑝 𝑝 𝑘 𝑘)
<0
117
Vehicle Interactions and Subsystems
118
Vehicle Interactions and Subsystems
Wheel rotational
elasticity of sidewalls Massless
ring/belt
Rim &
Rim &
wheel shaft Same dynamic wheel shaft
behaviour between
rim and road
Rigid longitudinal
and vertical supports
𝑓 𝑤
Road
𝑅 𝑙 𝑅 𝑚
Figure 2-32: Tyre model including the rotational elasticity of tyre sidewalls.
Damping in parallel with the elasticity is often motivated also:
+
𝑤 ∙ (𝑅 𝑚 𝑅 𝑙 )
𝑤 ∙ (𝑅 𝑚 𝑅 𝑙 ) [2.24]
𝑅 𝑙
𝑓 (𝑠 …) ℎ𝑒𝑟𝑒 𝑠
|𝑅 𝑙 |
If used in a system where 𝑚 and are input variables to tyre, the force will become a state vari-
able. It is then not a problem that slip is undefined for 0, because the explicit form of equations
will become as follows. Note that we simplify by only considering the case when 𝑙 > 0. And the
model validity is limited to … such that uniquely defines 𝑠 .
𝑠 ← ( …)
119
Vehicle Interactions and Subsystems
makes | | stays < . But, if decreases stepwise, one might end up there anyway for short time
intervals. In that case, it often gives physically acceptable solutions on vehicle level, to simply saturate
𝑠 so that | | is saturated at a certain level, e.g. 0.95 . For the brush model with uniform pres-
sure distribution, Eq [2.9], the inverted function becomes as follows, including such saturation:
∙
𝑓𝑜𝑟 | | ≤
∙ sign( )
( ) ∙ 𝑓𝑜𝑟 | | < 0.95 ∙
𝑠 4∙ | |
∙ [2.27]
∙ sign( )
∙ 𝑒 𝑠𝑒
{ 4∙ 0.95
∙ 𝑊 ∙ 𝐿2
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒]
∙𝐻
2.2.3.5.3 Relation between the two Transients Models
The models in 2.2.3.5.1 and 2.2.3.5.1 describe two different phenomena which both are present in the
real world. Since it is difficult to distinguish between the two delay-creating parameters, 𝑤 and
⁄𝐿 , one often models only one phenomenon. But then, one adjusts the numerical value of the used
parameter so that the model captures about the same delay as a real-world test. If 𝑤 ⁄𝐿 , the
two models coincide approximately.
This compendium does not give any recommendation of which of the two models is best.
Note that, with unsaturated in Eq [2.20], one has to limit the integration of to | | ≤ ∙ .
2.2.3.5.4 § Expansion of Brush Model to Partial Differential Equation
Contribution from Luigi Romano, Vehicle Dynamics at Chalmers
The brush model derived in 2.2.3.1 was derived for steady state deformation pattern, i.e. the defor-
mation was only 𝛾(𝜉 ), not 𝛾 (𝜉 𝑡). Some call such tyre models “(single-)point contact models”. The
models in 2.2.3.5.1 and 2.2.3.5.2 uses a point contact approach, but add in time derivatives. These
models fit well as parts in larger dynamic models of the whole vehicle and one can use ordinary simu-
lation methods for dynamic system, 1.5.1.1.5.
If assuming 𝛾(𝜉 𝑡), or “line contact model”, one actually models the transients in a physical way. The
Mathematical model then becomes a partial differential equation, PDE, as opposed to ODE. To imple-
ment model in a typical dynamic vehicle model (ODE) we can discretize the PDE to ODE. One can also
discretize already in the Physical model, see Figure below.
120
Vehicle Interactions and Subsystems
𝜉 𝐿 𝑅 𝜉 𝜉 0 𝜉 𝐿 𝑅 𝜉 (𝑡+ 𝑡) 𝜉 0
𝜉 𝜉
𝑡 𝑒 𝑡 + 𝜕𝑡
bristle at 𝜉
(not same bristle over time) 𝜕𝛾 𝛾 (𝑡 + 𝑡)
𝛾 𝜉 𝑡+𝜕𝑡 𝛾+ 𝜕𝑡 bristle at time 𝑡 + 𝑡
𝜕𝑡
𝜕𝛾 𝜕𝛾 𝑅
𝐻 𝑅 𝐻 𝜉 𝑅 𝛾
𝜕𝑡 𝜕𝜉 𝐻
𝛾 0𝑡 0 𝑛 𝛾 𝜉0 𝛾 𝜉 𝛾 0 ℎ𝑒𝑛 𝜉 > 0 𝑛 𝛾 𝛾 𝑓𝑜𝑟 ..𝑁
𝜕 𝜕 𝜉𝑖+1 𝜉𝑖 1
𝐻 𝛾 ≈𝑅 𝐻 ≈
𝜕𝜉 𝜕𝜉 2 𝐿/𝑁
𝛾 0 𝑛 𝛾 𝛾 𝑓𝑜𝑟 ..𝑁
Figure 2-33: § Two views of how bristle deformation varies in space and time. Left: Using a
“control volume” in which new bristles comes in and moves out. Right: Following a certain bristle.
Note that the figure only shows model and equation for the stick zone. With the model derived from
“following each bristle”, it is straight-forward to implement also stick, slip and transitions for each
bristle. With the model derived from the control volume, it is less easy to do this, but for many driving
cases one can assume that there is exactly one break-away point 𝜉 as in the steady state model. Refer-
ence (Romano, Bruzelius, & Jacobson, 2020) does this for a start from stand-still and brake to stand-
still, showing that the model in 2.2.3.5.2 can be motivated starting from the PDE from the “control vol-
ume”.
121
Vehicle Interactions and Subsystems
force
+ + 𝑅𝑅
+ 𝑇/𝑅𝑙 (if > 0)
+ 𝑅𝑅
𝑇/𝑅𝑙 (if < 0)
𝑅𝑅
𝑅
𝑠
𝑅
Operating point when free- 𝑅𝑅
rolling ( 0) forward ( > 0)
𝑅𝑅
𝑅𝑅
Figure 2-34: Longitudinal tyre force ( ) and normalized wheel torque (𝑇/𝑅𝑙 ). Slip 𝑠 defined so
that curve (𝑠 ) passes through diagram origin, which means that 𝑇(𝑠 )⁄𝑅𝑙 does not.
𝑤
𝑀
𝑀
𝑀
𝑡𝑝
𝑀 𝑀
≈
𝑀 𝑀 ≈𝑀
Figure 2-35: Deformation and forces of a Cornering Tyre. Wheel side slip angle is .
It is essential to distinguish between the steer angle and wheel (lateral or side) slip angle of the tyre.
Lower right part of Figure 1-43 shows this difference. The steer angle, or d, is the angle between ve-
hicle longitudinal direction and tyre longitudinal direction. The wheel side slip angle, is the angle be-
tween tyre longitudinal direction and the tyre translational velocity (=wheel hub velocity).
Assuming no longitudinal tyre slip, 𝑅 , the relation between the lateral force of a tyre and the
wheel side slip angle is typically as shown in Figure 2-39. The behaviour of the curve is similar to that
exhibited for longitudinal forces Figure 2-29 and Figure 2-30. It becomes even more similar if lateral
slip angle is replaced by lateral wheel slip, 𝑠 , which is tan( ) ≈ for small lateral slip.
𝑓𝑠 0
𝑠 sign(𝑅 ∙ ) {
. 𝑒. 𝑅
} tan( ) [2.28]
|𝑅 ∙ | 𝑅∙
Using magnitude in denominator |𝑅 ∙ | gives same sign of 𝑠 and for all combinations of signs of
and , which leads to easier formulas.
122
Vehicle Interactions and Subsystems
Drawn for 𝑅
x and >0
pressure
x
stick
slip
“mode”
x
shear stress
(lateral, i.e. perpendicular
to top drawing)
Figure 2-36: The brush model’s physical model for lateral slip. The bristles are to be thought of as
the tread in series with sidewall lateral elasticity.
Each bristle in Figure 2-36 is thought of as a part of the tread in series with a part of the sidewall. Fur-
ther, the bristles are considered as independent of each other, which is debatable. However, it is
enough for a quantitative explanation of the brush model for lateral slip. The derivation of model equa-
tions becomes similar as for the longitudinal model:
∙ 𝑡(𝜉 ) ∙ 𝑡(𝜉 ) 𝑡 (𝜉 ) 𝜉 ⁄ 𝑇 𝑛 𝑝
𝜏 ∙𝛾 ∙ ∙ { }
𝐻 𝐻 𝑇 𝑛 𝑝 ≈ |𝑅 ∙ |
∙𝜉 𝑊 𝐿2
∙ ∙ ∙𝜉 { 𝑠 } ∙𝑠 ∙𝜉
𝐻 |𝑅 ∙ | 𝐻 |𝑅 ∙ | 𝐻 |𝑅 ∙ | 𝑊 𝐿2
Note that subscript 𝑦 has been introduced where we need to differ towards the longitudinal brush
model. Correspondingly, subscript 𝑥 should be used in longitudinal model. As for longitudinal model,
we have to express the force differently for when friction limit is not reached within the contact and
when it is.
123
Vehicle Interactions and Subsystems
∙ ∙
∙𝑠 𝑓 |𝑠 | ≤ ⇔| |≤
∙
∙
sign(𝑠 ) ∙ ∙ ∙( ∙ ) 𝑒 𝑠𝑒
{ 4∙ |𝑠 | [2.29]
2
∙𝑊∙𝐿
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒] 𝑛 𝑠
∙𝐻 |𝑅 ∙ |
(Only valid for pure lateral slip, i.e.: 𝑅 )
Note that the lateral tyre slip 𝑠 is the sliding speed of tyre over ground in lateral direction, divided by
the same “transport speed” as for longitudinal slip, i.e. the longitudinal transport speed 𝑅𝑤 . Note
also that the lateral force and lateral tyre slip are counter-directed, which is logical since it is of fric-
tion nature.
Tyre Lateral Slip vs Wheel Slip Angle
The lateral tyre slip 𝑠 ⁄|𝑅 ∙ | in Eq [2.29] and Eq [2.31] can be compared with lateral wheel slip
angle arctan( ⁄ ), mentioned in context of Figure 2-35:
• “Lateral wheel slip” 𝑠 𝑤 𝑤⁄ 𝑤 tan( ), is how wheel (hub and tyre) moves over
ground and independent of wheel rotational speed.
• “Lateral tyre slip” 𝑠 ⁄|𝑅 ∙ |, is the slip used in the constitutive relation (𝑠 ).
• If no longitudinal tyre slip 𝑠 0, i.e. if 𝑅 , we have 𝑠 𝑤 𝑠 . Then 𝑠 𝑤 can be used in
the constitutive relation.
For a linearization, the most correct way is that lateral force is 𝑠 , as opposed to α. Often one
finds α as starting point in the literature, but this compendium uses 𝑠 . (In (Pacejka, 2005),
pp184-185, there is also a note that 𝑠 is more appropriate than .)
A difference is how one linearizes a vehicle model in and . A non-steered axle modelled with
α , needs to be approximated with “ 𝑟 𝑡 𝑛( 𝑣 ⁄ 𝑣 ) ≈ 𝑣⁄ 𝑣 ” to make the vehicle
model linear, see derivation of Eq [4.49]. However, with 𝑠 ; the vehicle model becomes linear
without further approximations. For a steered axle it is less obvious, but it does not help the lineariza-
tion to use α.
2.2.4.1.2 Model with Dependent Bristles, String Model
Opposed to the assumption in 2.2.4.1.1, the lateral deformations of the bristles are dependent on each
other, especially since the tread is mounted on the belt and the belt is rather like a string. So, we as-
sume a certain deformation of the sidewall, expressed in 𝜀 𝑛 and 𝜀 for the belt=”string” in Figure
2-37. This gives a slightly different model compared to 2.2.4.1.1. Models with such belt deformation
are called “tyre string models”. The shape of the string is dependent on the sidewall elasticity, e.g. tyre
profile height, but also of the side force itself. So, the model is intrinsically implicit; the string shape
influences the side force and the side force influences the string shape.
The derivation of model equations becomes similar as for the longitudinal model. Here is an interme-
diate result, an expression for the shear stress 𝜏 :
∙ 𝑡 (𝜉 ) (𝑅 tan(𝜀 𝑛 ) ) ∙ 𝑡(𝜉 )
𝜏 ∙𝛾 ∙ ∙
𝐻 𝐻
𝑡(𝜉 ) 𝜉⁄ 𝑇 𝑛 𝑝 (𝑅 tan(𝜀 𝑛 ) )∙𝜉
{ } ∙
𝑇 𝑛 𝑝 ≈ |𝑅 ∙ | 𝐻 |𝑅 ∙ |
𝑊 𝐿2
∙ tan(𝜀 𝑛 ) sign( ) ∙𝜉 { }
𝐻 |⏟
𝑅∙ | 𝐻
( 𝑦 )
124
Vehicle Interactions and Subsystems
(𝑠
∙⏟ tan(𝜀 𝑛 ) sign( )) ∙ 𝜉
𝑊 𝐿2
𝑦𝜀
𝑅
Drawn for 𝑅 and >0
𝜉
pressure
𝜉 stick
slip
“mode”
𝜉 shear
stress
(lateral)
Figure 2-37: Tyre sidewall deformation and tread deformation with the belt (the “string”) in
between. The drawn bristles are here assumed to represent only tread parts, while the sidewall is
treated as an elastic structure between rim and belt.
Comparing to 2.2.4.1.1, we can note that subscript 𝑡𝑟 is added to underline that and /𝐻
now means only the tread, not including the sidewall. Sidewall elasticity is instead handled with 𝜀 𝑛 .
The variable 𝑠 𝜀 is an auxiliary mathematical variable introduced only to make the expressions more
manageable; it will be eliminated later.
∙
∙𝑠 𝜀 𝑓| |≤
∙
sign(𝑠 𝜀) ∙ ∙ ∙( ∙ ) 𝑒 𝑠𝑒
4∙ |𝑠 𝜀 |
{
∙ 𝑊 ∙ 𝐿2
ℎ𝑒𝑟𝑒 𝑒𝑛𝑠 𝑜𝑛: [𝑓𝑜𝑟 𝑒] 𝑛 𝑠 𝑛 𝑠 𝜀 𝑠 tan(𝜀 𝑛 ) sign( )
∙𝐻 |𝑅 ∙ |
Now, this model is still implicit because 𝜀 𝑛 depends on 𝜏(𝜉 ). Introducing simplest possible (linear)
constitutive equation for this dependency as in Eq [2.30]:
125
Vehicle Interactions and Subsystems
∫𝜏 𝜉 𝑤𝜀 𝑛 tan(𝜀 𝑛 ) [2.30]
Figure 2-38: Comparison of model with independent and dependent bristles (String model).
The influence of vertical load was discussed in 2.2.3.1.2 but is better explained with dependent bris-
tles. Assumes that only the tread stiffness (and not 𝑤 𝜀 𝑛 ) varies with contact length, and that
this variation is proportional, ( ) , as we found in 2.2.3.1.2. This indicates a degressive char-
acteristics of ( ), which is also observed in measurements.
The model with dependent bristles is probably more correct. Anyway, we will use the other most in
this compendium, since it is much easier to combine with the longitudinal model (Eq [2.9]) to model
combined (longitudinal and lateral) slip.
126
Vehicle Interactions and Subsystems
∙𝑠 or 𝛼 ∙ [2.32]
For the brush model, or any other model which describes (𝑠 … ) or (α … ), one can de-
fine the “Lateral tyre slip stiffness” or “Tyre Cornering Stiffness”, or 𝛼 , which have the unit N/1 or
N/rad. It is the derivative of force with respect to slip or slip angle. Reference (ISO 8855) defines the
cornering stiffness as 𝛼 for slip angle 0. It is implicit that also 𝑠 0. And 0 when 𝑠 0:
𝜕 𝜕
𝛼 ( )| (𝜕 )| [2.33]
𝜕𝛼 𝛼= = 𝑦
𝑦= =
When using only small slip, it does not matter if the cornering stiffness is defined as the slope in an
versus diagram or versus 𝑠 𝑡 𝑛( ) diagram. Therefore, the notation for cornering stiffness
varies between 𝛼 and . Cornering stiffness has the unit 𝑁 which can be interpreted differently: as
[𝑁 𝑁⁄ 𝑁/(( /𝑠)⁄( /𝑠)) ] or as [𝑁 𝑁 ⁄𝑟 ].
The longitudinal tyre slip stiffness, , is normally larger than the lateral tyre slip stiffness, , which
can be explained with that the tyre is less stiff in lateral direction. Since it is the same rubber one could
argue that both and 𝐻 should be the same, but both due to longitudinal grooves in the tread and due
to lateral deformable sidewall, it is motivated to introduce different subscripts: ( ⁄𝐻) > ( ⁄𝐻) .
One could elaborate with different friction coefficients and , but in this compendium it is claimed
that friction is well modelled as isotropic. More about this in 2.2.4.7.5.
The cornering tyre forces initially exhibit a linear relation with the slip angle. A non-linear region is
then exhibited up to a maximum value. In Figure 2-39, the maximum slip angle is only 16 degrees (or
𝑠 tan( 6 𝑒 ) 0. 9) and one can expect that the tyre forces will drop as the slip angle ap-
proaches 90 degrees.
Adhesion Limit
Adhesion Limit
|Cornering Force| [kN]
0 4 8 12 16
Figure 2-39: Left: Influence of tyre design. Right: Influence of inflation pressure, (Gillespie, 1992).
127
Vehicle Interactions and Subsystems
Figure 2-40: Example of cornering stiffness versus vertical load for a truck tyre 295/80R22.5.
Figure 2-41: Cornering stiffness versus vertical load for some passenger car tyres. From flat track
tests.
128
Vehicle Interactions and Subsystems
Figure 2-42: § Preliminary measurements of lateral slip stiffness for twin-mounted trailer tyres for
heavy vehicles. Two tyres, “1” and “2”, are measured. “W” means worn to a small groove depth but
still legal. From Mattias Hjort and Sogol Kharrazi, VTI.
129
Vehicle Interactions and Subsystems
lost, the steering will tend to steer in the direction of body motion above the steered axle, which is nor-
mally relatively smooth and safe.
Figure 2-43 shows the combined response of lateral force and slip angle. It is interesting to note that
the steering torque reaches a peak before the maximum lateral force capacity of the tyre is reached. It
can be used by drivers to find, via steering wheel torque, a suitable steer angle which gives a large lat-
eral force but still does not pass the peak in lateral force. The reason why pneumatic trail can become
slightly negative is because pressure centre is in front of wheel centre, see Figure 2-13.
Tyre Aligning Moment in due to Lateral Shear in Brush Model
A model for (yawing) aligning moment around a vertical axle through centre of contact point, 𝑀 , will
now be derived. Any model for lateral shear stress can be used, but we will here only use the uniform
pressure distribution and independent bristles in 2.2.4.1.1. A corresponding expression as Eq [2.29] is
derived, but for 𝑀 instead of .
𝐿
𝐿
𝑀 𝑊 ∙ ∫ 𝜏 ∙ (𝜉 ) 𝜉 ⋯
𝐿
C ∙𝑠 𝑓𝑜𝑟 < ⇔𝑠 <
6 [2.36]
2 2
∙𝐿
( ) 𝑒 𝑠𝑒
{ 8∙ ∙𝑠 ∙ ∙𝑠
𝑊 𝐿2
ℎ𝑒𝑟𝑒
𝐻
steering axis
shear stress,
intersection
𝜏
with ground 𝑡𝑝
contact
patch
𝑡𝑝 𝑡𝑝 +
tyre
steering moment peaks at lower slip than lateral tyre force
steering moment + 𝑡𝑝
130
Vehicle Interactions and Subsystems
Figure 2-44: Aligning moment (𝑀 ) around contact patch center for uniform pressure distribution.
The lateral force and the aligning torque can be used to calculate the steering forces. If also the steer-
ing assistance is known, the steering wheel torque can be calculated. It can be noted that the model
does not include the moment from steering rotation itself, i.e. the torque counteracting .
2.2.4.7.2 Influence from Longitudinal Tyre Force
Figure 2-43 does not show effects from 𝜏 , but this is shown in Figure 2-45.
2.2.4.7.3 Tyre Spin Torque
The none-symmetry in shear stress around wheel centre in Figure 2-45 can appear as steady state;
non-symmetry in 𝜏 due to brush model and in 𝜏 due to non-symmetric vertical pressure. But there is
one additional reason to yaw moment in tyre contact patch. That is the friction yaw velocity of the
wheel 𝑤 𝑙 . The additional moment from this effect is often called (tyre) spin torque. One way to
model it is, are conceptually an elastic torsional spring in series with friction. The influence from spin
torque is only important at low speed, e.g. steering in low speed, where 𝑤 𝑙 ≈ 𝑤 𝑙 .
moment moment around
shear around wheel steering axis
stresses force center intersection
shear stress, 𝜏 𝑠
steering axis
intersection 𝑠 𝑠+
with ground
contact
patch
shear stress, 𝜏
tyre
𝑡𝑝
𝑡𝑝 𝑡𝑝 +
Figure 2-45: Influence from both lateral and longitudinal shear stress on aligning moment.
2.2.4.7.4 Camber Force
Camber force (also called Camber thrust) is the lateral force caused by the cambering of a wheel. Cam-
ber thrust 𝑚 is approximately linearly proportional to camber angle 𝑤 for small angles:
131
Vehicle Interactions and Subsystems
ground, so
≈ 𝑅𝑤
outlet
𝑅𝑤 𝑤
inlet same forces
𝑅𝑤 cos
𝜉
𝜉 𝐿 𝜉 0 𝜉 𝐿⁄
𝜉 0
Moving
ground 𝑚 Fixed ground
𝑤 𝑤 𝑚
Contact
ground rotates Upper bristle ends follows 𝑤 sin 𝑚 ≈
patch
this straight line, so ≈ 𝑤 𝑚
𝑝𝑝 𝜉 0
Lower bristle follows the rotating ground, so
Deformation 𝜉 𝜉 𝐿⁄
𝑙 𝑤 𝑤 𝑚
and 𝜏 𝜉 have
𝑚
this shape Conditions with low and/or very stiff bristles
Conditions where 𝒚 is saturated by 𝒑: (as a “metal-to-metal rolling contact”):
stick slip stick 𝜏
𝜏 + 𝑝
𝜏 0 𝑚 0
𝜏 0 𝑀 𝑚
𝜏 𝑝
𝑚 𝜏 𝑝 𝜏
Figure 2-46: Brush modes for Camber force (or Camber thrust) 𝑚 . For intuitive
understanding, note that the ground is thought of as rotating opposite to the vertical component
of wheel rotation ⃗ 𝑤 .
𝑡 𝑡 𝑘𝑒𝑠 𝑡 𝑒 𝑡𝜉 𝜉 𝜉
∫0 (𝑣 𝑜𝑤 𝑟 𝑦 𝑣𝑢𝑝𝑝 𝑟 𝑦(𝜉))∙ 𝜉 ∫0 ((𝜉 𝐿 ⁄2) 𝜔𝑧𝑔 )∙ 𝜉
𝜏 (𝜉 ) ∙ 𝛾 (𝜉 ) { 𝑓𝑜𝑟 𝑟 𝑠𝑡 𝑒 } ∙ 𝐻𝑦
∙ 𝐻𝑦
𝑓𝑟𝑜 𝑛 𝑒𝑡 𝑡𝑜 𝜉
𝜉 𝑛 𝑝 𝑡𝜉 𝑦 𝑦
2
𝜉 𝐿
𝜉
𝑦
{ } ∙ ∫ 𝜉(𝑅𝑤 𝑤 𝑡𝜉 𝐿⁄ ) ∙ 𝑡𝜉 ∙ [𝑅𝑤 𝑤 𝑡𝜉 ] ∙
𝑅𝑤 𝑤 𝑡𝜉 𝐻𝑦 𝐻𝑦 2 2 2 𝐻𝑦
2 2
𝜉 𝜉 𝑦 𝜉
(𝑅𝑤 𝑤 𝐿 𝑡𝜉 ) {𝑡𝜉 ≈ 𝑤 𝑚 } 𝑤 𝑚 ∙ (𝑅𝑤 𝑤 ( ) 𝐿
2 𝑤 𝜔𝑤 2 𝐻𝑦 𝑤 𝜔𝑤
𝜉 𝑦
) 𝑦
∙ (𝜉 2 𝐿 𝜉) ≈ ∙ 𝑚 (𝜉 2 𝐿 𝜉)
𝑤 𝜔𝑤 2 𝐻𝑦 𝑤 2 𝐻𝑦 𝑤
𝐿
Integration 𝑚 ∫ 𝜏 (𝜉 ) ∙ 𝜉 gives:
𝑊 𝐿3
𝑚 𝑚 ∙ 𝑚 ℎ𝑒𝑟𝑒 𝑚
𝐻 𝑅𝑤 [2.38]
Note: Only valid when no part of the contact patch is saturated by friction.
132
Vehicle Interactions and Subsystems
Equation [2.39] can be plotted as a circle, called the “Friction Circle”. Since the lateral and longitudinal
properties are not isotropic (due to carcass deflection, tread patterns, camber, etc) the shape may be
better described as a “Friction Ellipse” or simply “Friction limit”.
When not cornering, the tyre forces are de- 1
scribed by a position between -1 (braking) and maximum turning left
+1 (acceleration) along the Y-axis. Note that without propulsion and
braking
the scales of both axes are normalized to the 0.5 turning left while
maximum value for friction. braking
partial propulsion
Fy/(mu*Fz)
133
Vehicle Interactions and Subsystems
𝑡𝑒𝑟 𝑠 𝑛 𝑠𝑝𝑒𝑒
[2.40]
𝑜𝑛 𝑡𝑢 𝑛 𝑠 𝑛 𝑠𝑝𝑒𝑒 𝑅∙
When considering a combined slip model, one can establish the “total slip” 𝑠 𝑠, through the defini-
tion 𝑠 2
𝑠 2
𝑠 + 𝑠 . Then, it can be tempting to look for a function
2 2
𝑓(𝑠 ) and then decom-
pose into and through [ ] [+𝑠 ⁄𝑠 𝑠 ⁄𝑠 ] . This is not fully physical, which is
easiest understood by looking at the brush model for uniform pressure distribution: When 𝑠 is small
enough, there will be no slip in the contact, because 𝜏 < 𝑝 in whole contact. For such conditions and
isotropic linear deformation model of the bristles, the longitudinal and lateral models from before are
valid. So, is independent of 𝑠 and is independent of 𝑠 . So, using 𝑠 gives a non-physical depend-
ence. Anyway, such approximate models can be useful for conditions with larger slip, see 2.2.5.3.1.
tyre 𝑅 W=width
outlet inlet
𝜉
𝜉 𝐿 𝜉 𝜉 𝜉 0
H=height
road (relative
(relative
to wheel
to wheel hub) hub)
Contact patch, Velocities of bristle ends in
view from above: stick zone:
𝑅
force on bristle
𝐻 𝑠 ⁄
𝑠 𝑠 ⁄𝐻 Drawn for 𝑅 > and >0
𝑠 𝑠 𝑠 ⁄𝐻 Positions of bristle ends in
𝐻 𝑠 ⁄ 𝑠 stick zone, for a bristle which
entered contact time 𝑡 ago:
x
𝑅 𝑡
pressure
x stick arctan ⁄ 𝑡
slip 𝑡
“mode” Drawn for 𝑅 > and >0
x 𝜏 𝑙𝑝 longitudinal
𝜏 shear stress Continuous
2
𝜏 across break-away
2
point (𝜏 𝑙 𝑝 + 𝜏 𝑙 𝑝
x 𝜏 𝑙𝑝 lateral 𝜏 2 + 𝜏 2 ), but redistribution
𝜏 towards sliding direction in slip zone.
shear stress
Figure 2-48: Physical model for deriving brush model for combined slip.
134
Vehicle Interactions and Subsystems
𝜏 ∙𝑠 ∙𝜉 𝑛 𝜏 ∙𝑠 ∙𝜉
𝑊 𝐿2 𝑊 𝐿2
If no slip in contact:
𝐿 𝐿 𝐿
𝐿2
𝑊 ∙∫𝜏 ∙ 𝜉 𝑊∙∫ ∙𝑠 ∙𝜉∙ 𝜉 ∙ 𝑠 ∙ ∫𝜉 ∙ 𝜉 ∙𝑠 ∙ ∙𝑠
𝑊 𝐿2 𝐿2 𝐿2
𝐿 𝐿 𝐿
𝐿2
𝑊 ∙ ∫𝜏 ∙ 𝜉 𝑊∙∫ ∙𝑠 ∙𝜉∙ 𝜉 ∙ 𝑠 ∙ ∫𝜉 ∙ 𝜉 ∙𝑠 ∙ ∙𝑠
𝑊 𝐿2 𝐿2 𝐿2
Now, if there is a break-away point (𝜉 ) where slip starts, we can find it from 𝜏(𝜉 ) 𝑝 ⇒ 𝜏2 +
𝜏 2 2
𝑝 . Introducing an auxiliary parameter, 𝑘
2 ⁄ , and an auxiliary variable, 𝑠𝑘
+ ∙
[ 𝑦
] 𝑓 <0
∙
+ 𝑠 ∙ 𝑧 2 2 ∙ 𝑧
[ 𝑠 ] 𝑒 𝑠𝑒 𝑓 𝑠𝑘 ≤ 2∙ ⇔ √ + ≤
[ ] 𝑦 2
𝜉 2 𝐿 𝜉
+ (( 𝐿 ) ∙𝑠 + ∙ 𝐿
)
[ ] 𝑒 𝑠𝑒 [2.41]
𝜉 2 𝐿 𝜉 𝑦
{ (( 𝐿 ) ∙𝑠 + ∙ 𝐿
)
𝑅 2 2
ℎ𝑒𝑟𝑒 𝑠 𝑠 𝑠 √𝑠 +𝑠
|𝑅 ∙ | |𝑅 ∙ |
𝐿 ∙ 𝑊 ∙ 𝐿2
𝑘 𝑠𝑘 √(𝑘 𝑠 )2 + 𝑠 2 𝜉
𝑠𝑘 ∙𝐻
The terms with (𝜉 ⁄𝐿)2 comes from shear stress in the stick zone, and the direction is in proportion to
the stiffnesses and (anisotrop elasticity). The terms with (𝐿 𝜉 )⁄𝐿 comes from the shear stress
in the slip zone and it is counter-directed to the relative velocity between the surfaces (isotropic fric-
tion).
Figure 2-49 shows results from the model. It can be observed that is independent of 𝑠 and is in-
dependent of 𝑠 for ≤ ∙ ⁄𝑠. This is a reasonable consequence of that no sliding occurs so that
forces are purely defined by the elasticity, not the friction. At 𝑠 ≈ 0.0 . .0.04, we see that increases
with utilization of | | at some areas. This is a redistribution of force from longitudinal to lateral, due
to that the tyre stiffer in longitudinal than lateral.
135
Vehicle Interactions and Subsystems
𝑠 sign 𝑅
leaning
Anti-symmetric,
𝑠
quadrant
in the 3rd
𝑠
Symmetric, in
quadrant
the 2nd
Figure 2-49: Tyre (𝑠) with iso-curves for longitudinal slip 𝑠 and lateral tyre slip 𝑠
⁄|𝑅𝑤 | (5 left diagrams) and for 𝑠 (𝑅𝑤 )/|𝑅𝑤 | and slip angle
𝑟 𝑡 𝑛 ( ⁄ ) (5 right diagrams). Tyre model from Eq [2.41] used. Levels used for slip and
𝑡 𝑛(𝑠 𝑝 𝑛 𝑒): 0 ±0.0 ±0.0 ±0.05 ±0. ±0. ±0. ±0.4.
∙ 𝑧
Note that, for √ 2
+ 2
≤ 2
, is independent of 𝑠 and of 𝑠 . This is because for such low
slip and uniform pressure distribution, there is no slipping part in the contact patch. It is the slipping
part that creates the ”cross-dependence” that depends on 𝑠 and on 𝑠 . The independence ap-
pears as the perpendicular iso-𝑠 and iso-𝑠 curves for < in the upper diagrams in Figure
2-49.
136
Vehicle Interactions and Subsystems
137
Vehicle Interactions and Subsystems
leaning
𝑠 sign 𝑅
Anti-symmetric,
𝑠
quadrant
in the 3rd
𝑠
sign 𝑅
sign 𝑅
Symmetric, in
quadrant
the 2nd
𝑠
𝑠
Figure 2-51: Tyre (𝑠) with iso-curves for longitudinal slip 𝑠 and lateral tyre slip 𝑠
⁄|𝑅𝑤 | (5 left diagrams) and for 𝑠 (𝑅𝑤 )/|𝑅𝑤 | and slip angle
𝑟 𝑡 𝑛 ( ⁄ ) (5 right diagrams). Tyre model from Eq [2.42] and Eq [2.47] used. Levels used for
slip and 𝑡 𝑛(𝑠 𝑝 𝑛 𝑒): 0 ±0.0 ±0.0 ±0.05 ±0. ±0. ±0. ±0.4.
If 𝑠 is varied while 𝑠 0 in Eq [2.42], leads exactly to the longitudinal slip model with peak (Eq
[2.17]). Similar peak in is found for varying 𝑠 while 𝑠 0. The peaks in and are equally large,
but the peak in occurs at larger slip: 𝑠 𝑝 𝑘 > 𝑠 𝑝 𝑘 . Also for combined slip, a maximum in oc-
curs, but on a “2-dimensional ridge” in the “cake plot” in Figure 2-52.
138
Vehicle Interactions and Subsystems
Figure 2-52: Plots of the vector field [ ] 𝑓([𝑠 𝑠 ]) or 𝑓 (𝑠), using the results from
Figure 2-51. “Cake-plot” is an expression from (Weber, 1981).
2.2.5.2.1 § Some Specific Operating Conditions
𝑣
𝑅 . 3
≈ 0.9709
𝑅
0.
tan 0. ≈
≈ 0. 00
0
0. 𝑟
+
Figure 2-53: § Part from Figure 2-51. Shows specific conditions: (rotationally) locked ( 0),
infinite spin ( ± ), and 𝑥| |. The 𝑥| | appears at slight negative ≈ 500 𝑁.
139
Vehicle Interactions and Subsystems
A problem with Eq[2.43] is that anisotropy cannot be represented. Inspired by Eq [2.41] and Eq [2.42],
we can adjust it to Eq [2.44], which has only one example of function 𝑓. Eq [2.44] is not fully physically
motivated. However, it can be tuned to experiment data which has different and .
+𝑠 𝑅 ∙
[ ] [ 𝑠 ] [ 𝑤 ]
𝑠 2
√(𝑅𝑤 ∙ )2 + ( )
𝑅𝑤 ∙
ℎ𝑒𝑟𝑒 𝑠 2 𝑠2 𝑠2 + 𝑠2 ℎ𝑒𝑟𝑒 𝑠 𝑛 𝑠
|𝑅𝑤 ∙ | |𝑅𝑤 ∙ |
𝑓 <0 [2.44]
𝑛 {
𝑓( 𝑠𝑘 ) 𝑒 𝑠𝑒
ℎ𝑒𝑟𝑒 𝑒. . 𝑓 min( 𝑠𝑘 )
ℎ𝑒𝑟𝑒 𝑠𝑘 √(𝑘 𝑠 )2 + 𝑠 2 𝑘 ⁄
140
Vehicle Interactions and Subsystems
|
§ Comparison with Physically Motivated Model
𝑅 𝑅 𝑅
𝑠 𝑠 𝑠
;
𝑅 𝑠 𝑅 𝑠 𝑅 𝑠
𝑠 𝑠 𝑠
𝑠 𝑠 𝑠 ≈
𝑅 𝑅
Figure 2-54: § Left and middle shows that causality can be changed for a physically motivated
model, such as in Eq [2.41] and Eq [2.42], even if the middle generates an algebraic loop. The right
shows how Eq [2.45] and Eq [2.46] “cheats”, by using an 𝑠 definition which assumes 𝑅 ≈ ,
which is only motivated without combined slip.
Figure 2-55 shows experiment data on how slip stiffness varies with vertical force.
Contact lenght [mm]
Lateral,
Figure 2-55: Measurements for varying vertical force. Left: Slip stiffnesses. Right: Contact length.
2.2.5.4.1 Explanation of Higher Slip Stiffness Longitudinal than Lateral
The model in 2.2.4.1.1.1 explains why tyre is more slip stiff in longitudinal than lateral direction, i.e.
why > . We can then assume iso-tropic brush bristles in both friction and shear stiffness, i.e.
⇒ and 𝑤𝜀 𝑛 ⁄( 𝑤 𝜀 𝑛 + ) . From experiments, we typi-
cally find ⁄ { 𝑡 𝑛𝑜 𝑛 𝑁 𝑚 } 𝑁 𝑚 ⁄ 𝑁 𝑚 𝑁 𝑚⁄ 𝑁 𝑚 𝑘𝑁 𝑚 .5 . . .
This could be explained with the model if 𝑤 𝜀 𝑛 ⁄ ⁄(𝑘𝑁 𝑚 ) . . . This shows that neither
of 𝑤 𝜀 𝑛 nor is neglectable.
2.2.5.4.2 Model for how Lateral Slip Stiffness is Degressive with Vertical Force
The model in 2.2.4.1.1.1 can also explain why lateral slip stiffness is degressive with , as indicated
already in 2.2.3.1.3. The was found to be proportional to , due to that contact length increase pro-
portional to √ . A very conceptional reasoning of how 𝑤 𝜀 𝑛 could vary with follows now: We de-
fined 𝑤 𝜀 𝑛 𝜀 𝑛 ⁄ . For a fix lateral deformation 𝜀 𝑛 ≈ ⁄𝐿 and ≈ 𝐿. So, 𝑤 𝜀 𝑛 ⁄𝐿2
⁄ . With and 𝑤 𝜀 𝑛 ⁄ , and given 𝑁 𝑚 and 𝑁 𝑚 at given nominal
𝑁 𝑚 , the model gives Eq [2.47] and Figure 2-56. So, by knowing Nom and measuring 𝑁 𝑚 and
𝑁 𝑚 one gets 𝑘𝑁 𝑚 as 𝑘𝑁 𝑚 𝑁 𝑚 ⁄ 𝑁 𝑚 and thereby a quantified model of ( ) and ( ).
141
Vehicle Interactions and Subsystems
𝑛
𝑁 𝑚
𝑁 𝑚 𝑘𝑁 𝑚 2
𝑁 𝑚 [2.47]
𝑁 𝑚 +
𝑁 𝑚
Figure 2-56: Model with Dependent Bristles explains that Lateral Slip Stiffness is degressive with
. The is (direction independent) slip stiffness due to tread (𝑡𝑟).
142
Vehicle Interactions and Subsystems
60000
60
0
0
Figure 2-58: Different tyre models which will filter road irregularities differently. Picture from
Peter Zegelaar, Ford Aachen.
143
Vehicle Interactions and Subsystems
𝑊𝑒 𝑟𝑅 𝑡𝑒 𝑟 𝑡 𝑜𝑛 𝑜𝑟 𝑒 ∙ 𝑆 𝑛 𝑆𝑝𝑒𝑒 ⇒ 𝑊𝑒 𝑟𝑅 𝑡𝑒 𝑘𝑤 𝑣 ∙ ∙ ≈
)2 + 2 )2 + 2
≈𝑘∙ ∙ √(𝑅 ≈ 𝑘𝑤 𝑣 ∙ ( ∙ 𝑠) ∙ √(𝑅 ≈
≈ 𝑘𝑤 𝑣 ∙ ( ∙ 𝑠) ∙ (𝑠 ∙ 𝑇 𝑛 𝑝 ) ⇒ 𝑊𝑒 𝑟𝑅 𝑡𝑒 ≈ 𝑘𝑤 𝑣 ∙ ∙ 𝑠2 ∙
[2.49]
2 2
where √ + 𝑠 √𝑠 2 + 𝑠 2; 𝑇 𝑛 𝑝 defined as in Eq Error!
Reference source not found. and 𝑘𝑤 𝑣 is a constant for a certain tyre with a
certain temperature, rolling on a certain road surface, which characterises
the wear averaged over the contact patch.
road surface
Figure 2-59: Wheel/axle suspension described as modular sub-model per axle. It may be noted that
both wheel model (main geometry such as wheel radius) and tyre model (how and vary with
tyre slip and ) is a part of each Wheel & Tyre sub-model.
The simplest view we can have of a suspension system is that it is an individual suspension between
the vehicle body and each wheel, consisting of one linear spring and one linear damper in parallel.
Chapter 5 uses this simple view for analysis models, because it facilitates understanding and it is
enough for a first order evaluation of the functions studied (comfort, road grip and fatigue load) dur-
ing normal driving on normal roads.
The full 3D aspect of suspension is not covered here in 2.3. Instead, a division into 2D is done in 2.3.3
Suspension -- Heave and Pitch and 2.3.4 Suspension -- Heave and Roll, aiming at Longitudinal and Lat-
eral dynamics, respectively. The full 3D aspects are briefly addressed in 4.5.3.2.5 and 0.
144
Vehicle Interactions and Subsystems
2.3.1.1 Linkage
Linkage, which has the purpose to constrain the relative motion between wheel and body via kinemat-
ics to one dof (approximately vertical translation), or, for a steered axle, also allow one more dof (ap-
proximately yaw rotation) per axle. The linkage defines how longitudinal and lateral tyre forces are
brought to the body (sprung mass).
The linkage consists of links (or members) and joints; mainly ball joints, but sometimes others, such as
hinge joints. The coordinates (“hard-points”) of these joints are the real design parameters, but the dy-
namic behaviour of a complete vehicle model can be expressed in much fewer parameters, namely the
“effective pivot points”. These effective points are used in 2.3.3 and 2.3.4.
2.3.1.3 Dampers
Dampers have the purpose to dissipate energy from any oscillations of the vertical displacement of the
wheel relative to the body. High damping damps oscillations, but high damping also increases the
shock force transmittance (with this reasoning, the name “shock-absorber” is misleading). The most
common design is the hydraulic piston type. Simplest understanding or modelling is a linear relation
145
Vehicle Interactions and Subsystems
between the vertical deformation speed and force. But the following non-linearities should be men-
tioned:
• The dampers are normally intentionally designed to be different in different deformation di-
rection. Typical values for passenger cars are about 3 times more damping in rebound (exten-
sion) than compression (bump). For heavy vehicles, the difference is often even larger, maybe
a factor 5 or 10.
An intuitive explanation to the asymmetry can be found by comparing driving over different
transient unevenness, a bump or a hole. In both cases, large spring stiffness and large damping
cooperate to transfer the road unevenness and worsen comfort. But average damping cannot
be too low, since damping damps oscillations. When driving over a bump, the force can be very
large due to the bump stop in the spring. When driving over a hole, the force is limited by
wheel lift from ground. The spring, including bump stop and wheel lift, is therefore generally
stiffer for a bump than for a hole. So, to compensate for this, the damping is made opposite:
more damping when driving over a hole, i.e. rebound (extension).
• Damping in leaf springs is non-linear since they work with dry friction.
• The dampers can be designed to be controllable during operation of the vehicle. This can be
used to change the damping characteristics to adjust for varying roads or varying weight of ve-
hicle cargo or to be controllable in a shorter time scale for compensating in each oscillation cy-
cle. The latter is called “semi-active suspension” and is available on some high-end vehicles on
market.
on axle i on axle i 𝑟 𝑤 𝑙
With these definitions, the vertical wheel forces for a two-axle vehicle will be as follows. It is assumed
that the vehicle is symmetrical and that there are connections between the wheels only as anti-roll
bars on each axle. The body is fixed and road under the wheels are displaced with . Notation 𝑙 is
146
Vehicle Interactions and Subsystems
axle rate and 𝑤 𝑙 is wheel rate for axle . The time derivative of spring force and vertical veloci-
ties are used to avoid involving pre-tension in the springs.
𝑙 𝑙 𝑙
𝑪 𝟎 𝑤 𝑙 ( 𝑤 𝑙 )
[ 𝒘𝟏 ] [ ] ℎ𝑒𝑟𝑒 𝑪𝒘𝒊 [ ]
𝟎 𝑪𝒘𝟐 2𝑙 𝑙
2𝑙 ( 𝑤 𝑙 ) 𝑤 𝑙
[ 2 ] 2
If both body and road under the wheels is moving, we simply exchange [ 𝑙 2𝑙 2 ]𝑇 with
[ 𝑙 2𝑙 2 ] 𝑇 [ 𝑙 2𝑙 2 ]𝑇
. The axle roll stiffness becomes 𝑙𝑙
( ℎ 𝑥 ⁄ 4) 𝑊 [Nm/rad] for 𝑀 𝑙 .
Similarly, for axle , we can define axle (damping) rate 𝑙 , wheel (damping) rate 𝑤 𝑙 and axle roll
damping 𝑙𝑙 . However, 𝑙 is often simply 𝑤 𝑙 since there are typically no dampers connect-
ing left and right wheel. Corresponding 4 × 4 damping matrix becomes a diagonal matrix, since 𝑙
normally is 𝑤 𝑙.
If we know the design parameters (stiffness and location of the actual spring) we can calculate the
rates. This will be exemplified in 2D and briefly discussed for general (3D) below.
2.3.2.1 Explanation in 2D
A very simplified suspension is assumed in Figure 3-30. The stiffnesses 𝑓 and are the axle spring
rates (or effective stiffnesses of each axle). The real spring may have another stiffness 𝑙 , but its effect
on vertical force is captured in the effective stiffness. An example of how the effective stiffness is found
for a 2D model (heave and pitch) from a real suspension design is given in Figure 2-61.
Transverse (Panhard) rod Strut
Shock absorber rod 𝑝 𝑛 𝑝
Coil 2 3
1 spring
𝑝
Drive
axle a
physical spring
4 “effective suspension”
5
virtual spring
with stiffness 𝑝 𝑛 with stiffness
𝑝 𝑛
𝑝
“trivial
suspension”
6a 6b
Moment equilibrium of arm:
𝑝 𝑛 𝑝 ⇒ 𝑝 𝑛 𝑝
𝑝
Compatibility: 𝑟𝑟
𝑟
Equivalence in stiffness:
2
⇒ 𝑝 𝑛 ⇒ 𝑝 𝑛
𝑟 𝑟 𝑟
Figure 2-61: From suspension design spring stiffness to effective stiffness. 6a: Final Mathematical
model. 6b: Interpretation (“Reverse modelling”) back from Mathematical to Physical model,
showing “Trivial (linkage) suspension”.
147
Vehicle Interactions and Subsystems
Note that the factor ( / )2 is not the only difference between effective and real stiffness, but the ef-
fective can also include compliance from other parts than just the spring, such as bushings and tyre.
There will also be a need for a corresponding effective damping coefficient, see 3.4.5.2.2, or axle
(damping) rate. How forces in road plane is transferred is not well captured in this model, see in
Figure 2-61. Compared to suspension models later in compendium, Figure 2-61 ends with a “trivial
(linkage) suspension model”. In 3.4.5.2, the suspension linkage is better modelled, which allows valid-
ity for (propulsion and braking). Corresponding for is modelled in 4.3.10.3.
2.3.2.2 § Expansion to 3D
It should be noted that above is one of 2 possible 2D views of explaining effective stiffness. A similar
reasoning could be done in view from rear. Then we would see the same physical spring but other lev-
ers ( and ). So, that view would give another value of effective stiffness. A realistic model for both
and has to use pivot axis in 3D instead. If a prototype is available, one can also measure effective
stiffness, because it is the actual stiffness that one would feel if one tries to lift the wheel by grabbing
the tyre-to-ground contact surface. In a first approach one can explain a 3D model as moment equilib-
rium around a moment-free pivot axis. However, for a general 3D motion, there should be a screw
joint along the pivot axis, see mathematical formulation of equilibrium in 4.5.3.2.5. The lead of the
screw is often assumed to be zero and it is no examples below where a non-zero lead is identified in an
example suspension.
gf
90 deg
ef
ef
gf
Figure 2-62: Example of typical front axle suspension, and how pivot point is found. The example
shows a McPherson suspension. From Gunnar Olsson, LeanNova.
148
Vehicle Interactions and Subsystems
V-stay
Stabilizer bar
Bump stop
=0
𝑇 𝑇
Figure 2-63: Axle suspensions/installations for double rear axle heavy vehicles.
Figure 2-64 and Figure 2-65 show design of two axles with independent wheel suspensions. Figure
2-66 shows an axle with dependent wheel suspension. These figures show how to find wheel pivot
points and roll centre. In the McPherson suspension in Figure 2-65, one should mention that the strut
is designed to take bending moments. For the rigid axle in Figure 2-66, one should mention that the
leaf spring itself takes the lateral forces. Symmetry between left and right wheel suspension is a rea-
sonable assumption and it places the roll centre symmetrically between the wheels, which is assumed
in the previous models and equations regarding roll centre.
Pivot point
motion for points for wheel
90 deg
moving with hub
90 deg
Figure 2-64: Example of how to appoint the pivot point for one wheel, and roll centre height, for an
axle with double wishbone suspension.
149
Vehicle Interactions and Subsystems
motion for
90 deg
points moving
with hub Pivot point
for wheel
motion for a
point moving
with hub, where
the wheel which
is in contact
with ground
From the other
90 deg Roll centre wheel on same
for axle axle
90 deg
Figure 2-65: Example of how to appoint the pivot point for one wheel, and roll centre height, for
axle with double McPherson suspension.
Generally, a “rigid axle” gives roll centre height on approximately the same magnitude as wheel radius,
see Figure 2-66. With individual wheel suspension one has much larger flexibility, and typical chosen
designs are 30..90 mm front and 90..120 mm rear.
Figure 2-66: Example of how to appoint the pivot point for one wheel and roll centre for axle with
rigid axle suspended in leaf springs. From (Gillespie, 1992).
The target for roll centre height is a trade-off. On one side, high roll centre is good because it reduces
roll in steady state cornering. On the other side, low roll centre height is good because it gives small
track width variations due to vehicle heave. Track width variations are undesired, e.g. because it
makes the left and right tyre lateral force fight against each other, leaving less available friction for lon-
gitudinal and lateral grip. Roll centre is normally higher rear than front. One reason for that is that the
main inertia axis leans forward, and parallelism between roll axis and main inertia axis is desired.
150
Vehicle Interactions and Subsystems
𝑇 𝑇
(on one
driven axle)
Body
Prime mover
(ICE, internal Wheels
combustion
engine)
Transmission
(Clutch & Gearbox & Final gear)
151
Vehicle Interactions and Subsystems
𝐴 𝑇𝐴 𝐴 𝐴2 𝑇𝐴2 𝐴2
𝑇 2 𝑇 2 2
Figure 2-69: Alternative notation and sign conventions for Propulsion system models, following
”1.6.1.1.2 Multiple cut forces notation”.
𝑇
𝑇
𝑢𝑟 𝑡 𝑜𝑛 [𝑠]
0 0 00 03 04 05
152
Vehicle Interactions and Subsystems
gear 1 0
( typically between 0.1 and 0.5.)
gear 2
0
0 Speed, v
Figure 2-71: Left: Fuel consumption map. Curves with constant specific fuel consumption
[ ⁄(𝑘𝑊 ℎ )], which is ⁄efficiency. Middle: Specific fuel consumption curves transformed to
Traction Diagram, for different gears. Right: How efficiency with efficiency-optimal gear ratio
drops when < 𝑚 .
153
Vehicle Interactions and Subsystems
Propulsion
force, F
gear 1
gear 2
0
0 Speed, v
Figure 2-72: Left: Efficiency map for a typical brushless DC motor, from (Boerboom, 2012). Elliptic
curves show where efficiency, in [1/W/W], is constant. Right: The efficiency curves can also be
transformed in Traction Diagram, see 3.2.1, for a given gear.
Quadrant 1 and 3:
Motor operation,
converting electric
power to mechanical:
𝜂 𝐼⁄ 𝑇
Quadrant 2 and 4:
Generator operation,
converting mechanical
power to electrical:
𝜂 𝑇 ⁄ 𝐼
Figure 2-73: § 4 quadrant efficiency map for electric motor, from Ref (Brian Bole, 2012).
2.4.2.2.1 § Efficiency Model for Electric Machine
A model 𝜂 𝜂(𝑇 ) is useful when studying selection of gears and combination of multiple electric
drivetrains. For this, interpolation in a map from empiric measurements can be used. However, a sym-
bolic expression based on physics can have advantages in computation performance and validity
range. If introducing 2 linear resistors ( 𝑅 𝐼) one can explain some of the losses, see following fig-
ure.
154
Vehicle Interactions and Subsystems
𝑅𝑝
𝜂
0 [𝑉] (Assuming 𝑇 has same sign as 𝐼)
Numerical solution:
with 00 𝑅 0. 𝑅𝑝 5000
𝑇 𝑅𝑠 𝑅𝑠 ( + 𝑅𝑠 ⁄𝑅 𝑝 )
𝜂 𝑝 ( + +√ 𝑇 ) 𝜂 𝑝 (𝑇 𝑅𝑠 𝑅𝑝 )
⁄𝑅𝑝 + 𝑇 𝑅𝑝
Already in the mathematical model in figure above and from expression for 𝜂 𝑝,
we see that 𝑇 and
only apears in the combination 𝑇 , which is the mechanical power. So, the model with 2
resistors gives constant 𝜂 along curves with 𝑇 𝑜𝑛𝑠𝑡 𝑛𝑡. So, the single 𝜂 peak at a cer-
tain ( 𝑇) cannot be reflected by this model. The compendium does not go further in physical
modelling of EMs, so a non-physical factor is added with 2 tuning parameters, :
𝜂 𝜂 𝑝 (𝑇 𝑅𝑠 𝑅𝑝 ) ( 𝑘 ) ( 𝑘𝑇 𝑇 )
An example of parameterisation of this model is plotted in the following figure.
155
Vehicle Interactions and Subsystems
Figure 2-75: § Model of power efficiency of an electric machine based on 2 resistors and 2 tuning
parameters. Parameters: 00 [𝑉] 𝑅 0.06 [𝛺] 𝑅𝑝 500 [𝛺] 𝑘𝜔 9.88
8 [ ⁄(
0 𝑟 ⁄𝑠) ] 𝑘 𝑇 4.50 0 6 [ ⁄(𝑁 )2 ]
2
2.4.3 Transmissions
In some contexts, “transmission” means the 1-dimensional transmission of rotational mechanical
power from an input shaft to one output shaft. Such are called “Main transmissions”. In other contexts,
“transmission” means the system that distributes the energy to/from an energy buffer and to/from
multiple axles and/or wheels. Such are called “Distribution transmissions”.
156
Vehicle Interactions and Subsystems
powershifting transmission for a hybrid propulsion system is seen in Figure 2-85. A dummy sequence
of shifting is simulated in Figure 2-86.
Interpret pedals.
Request engine torque Decides how to shift. Engages
and when to shift gear. and disengages the clutches
C3
e in C2
out Driven wheel
Vehicle body
Prime
mover C1 w x
157
Vehicle Interactions and Subsystems
A locked differential, it is generally more difficult to analyse and understand in a vehicle manoeuvre.
Here, the wheels are forced to have same rotational speed, and, in a curve, that involves the tyre longi-
tudinal slip characteristics, so the explicit form solution involves more equations.
So, open/locked differential is the basic concept choice. But there are additions to those: One can build
in a friction clutch which is either operated automatically with mechanical wedges or similar or oper-
ated by control functions. Another was is to build in one-way clutches which can be coupled in dynam-
ically, see (Lidberg & Alfredsson, 2009). One can also build in an EM which moves torque from one
wheel to the other. However, the compendium does not go further into these designs.
Shafts are also parts of the distribution transmissions. If oscillations are to be studied, these has to be
modelled with energy storing components:
• Rotating inertias or Flywheels ( 𝑇 𝑛 𝑇 ) and
• Elasticities, compliances or springs: (𝑇 ∙( 𝑛 )).
𝑇 𝑇 𝑙 >0
𝑥 1: 𝑇 < 𝑥 =0: 𝑥 =+1:
2 𝑇 𝑙 0 𝑇>+ 𝑇 +
𝑙 2 𝑙 <0
𝑙 𝑙
𝑥
𝑥
𝑇<
Figure 2-77: Model of a clutch. The 𝑥 is a discrete state, declared as Integer.
𝑓𝑥 𝑡ℎ𝑒𝑛 𝑇 +
(Continuous) Equations
{𝑒 𝑠𝑒 𝑓 𝑥 0 𝑡ℎ𝑒𝑛 𝑙 0
using the discrete state 𝑥 :
𝑒 𝑠𝑒 𝑓 𝑥 + 𝑡ℎ𝑒𝑛 𝑇 𝑒𝑛 𝑓
(
ℎ𝑒𝑛 𝑥 𝑛 )
𝑙 > 0 𝑡ℎ𝑒𝑛 𝑥 0 [2.54]
(Discrete State) 𝑒 𝑠𝑒 ℎ𝑒𝑛 (𝑥 0 𝑛 𝑇< ) 𝑡ℎ𝑒𝑛 𝑥
ransition Equation: 𝑒 𝑠𝑒 ℎ𝑒𝑛 (𝑥 0 𝑛 𝑇 > + ) 𝑡ℎ𝑒𝑛 𝑥 +
{𝑒 𝑠𝑒 ℎ𝑒𝑛 (𝑥 + 𝑛 𝑙 < 0 ) 𝑡ℎ𝑒𝑛 𝑥 0 𝑒𝑛 ℎ𝑒𝑛
158
Vehicle Interactions and Subsystems
…
calculates “clutch torque capacity”, 𝑡 .
… …
…
Further
components
… J J …Further
components elasticity clutch elasticity
flywheel clutch flywheel
…
…
… …
… J …
damper clutch damper
(e.g. 𝑇( ) prime (e.g. (𝑠 )
flywheel clutch elasticity mover model) tyre model)
Figure 2-78: Examples of differently modelled surroundings of a clutch.
For complete main transmissions, as automatic transmissions, there are several clutches involved, the
implementation of the ideal model in Eq [2.54] can be very demanding. It is modelled in Modelica. The
discrete dynamics is modelled with the discrete state (xd declared Integer) and the operator
“pre(z)”, which holds the value from last time instant or, in an event, from the last event iteration.
2.4.4.2.1 Clutch and Elasticity in Series
The easiest surrounding to a clutch is in series with elasticity, assuming velocities can be input. Figure
2-79 shows such. It also adds 𝑠𝑡 𝑡 𝑜𝑛 𝑠𝑡 > ; a different static and dynamic friction, 𝑘 𝑠𝑡
𝑙𝑝 .
159
Vehicle Interactions and Subsystems
12 3 temporary elasticity
𝑡 (rigid shaft when
(permanent) 𝑡 clutch is slipping)
elasticity 2
2
𝑇 𝑇 𝑇 𝑇
𝑇 𝑇
2 𝑇
clutch 2 𝑇
clutch
//Clutch:
if slip==0 then 1 //Clutch and Elasticity:
if slip == 0 then
3
w1 = wx; w1 = wx;
else der(state) = cs*(wx - w2); T = state;
der(T) = sign(slip)*der(cc); else // if abs(slip) == 1 then
end if; T = +sign(slip)*cc;
0 = (wx - w2); //rigid elasticity
der(state) = void; //state not interpreted as T
end if;
when pre(slip)== 0 and T >+st*cc then slip=+1; when pre(slip)== 0 and T >+st*cc then slip=+1;
reinit(T, +cc); reinit(state, +cc);
elsewhen pre(slip)== 0 and T <-st*cc then slip=-1; elsewhen pre(slip)== 0 and T <-st*cc then slip=-1;
reinit(T, -cc); reinit(state, -cc);
elsewhen pre(slip)==+1 and w1<wx then slip= 0; elsewhen pre(slip)==+1 and w1<wx then slip= 0;
elsewhen pre(slip)==-1 and w1>wx then slip= 0; end when; elsewhen pre(slip)==-1 and w1>wx then slip= 0; end when;
//Elasticity:
der(T) = cs*(wx-w2);
1.0
1 3
1.0
0.5 0.5
0.0 0.0
-0.5 -0.5
-1.0 -1.0
-1.5 -1.5
0 5 10 15 0 5 10 15
w1 w2 wx w1 w2 wx
2 1 2 3
0 0
-2 -2
0 5 10 15 0 5 10 15
//Clutch
5if
slip and Elasticity:
T > +st*cc and w1 > w2 then der(T) = +der(cc); 2 slip
Figure 2-79: Model of clutch and elasticity in series, modelled for speed input from both sides.
Implemented in Modelica. Implementation 1,2 and 3. Without discrete state and without when (2)
only works well (without chattering) for 𝑠𝑡 .0. Switched physical model (3) has a state
(state) depending on the discrete state 𝑠 𝑝; elasticity is modelled only when slip=0.
2.4.4.2.2 Clutch between Inertias
When connecting 2 inertias or 2 elasticities with a clutch is more complicated. For clutch between in-
ertias, it is proposed to use 𝑙 as state variable, because it makes it easy to keep 𝑙 0 during
stick.be modelled, see Figure 2-80. The values 𝑥 ± are temporary during a state event; it enables a
total state transition during one event between 𝑥 and 𝑥 + , without (wrongly) logging an
intermediate 𝑥 0. With 𝑠𝑡 < , there is risk for chattering solutions.
160
Vehicle Interactions and Subsystems
<0
𝑙
𝑙 <0
𝑥 +
𝑙
𝑥 2: 𝑇 < 𝑠𝑡 𝑥 =0: 𝑥 =+2:
0 𝑇 0 𝑇 > +𝑠𝑡 𝑇 +
𝑙
𝑙 𝑙 𝑙 >0
𝑥
>0
𝑥 𝑙 <0 𝑥 1: 𝑇 𝑠𝑡 𝑙 <0
𝑙
𝑥 𝑙 <0
(𝑥 and + are only temporary during the event.)
𝑇 < 𝑠𝑡
Figure 2-80: Modelica of a clutch which works to be connected between inertias.
2.4.4.2.3 Functional or Inverse Clutch Model
In some cases, it can be suitable to prescribe how the clutch is operates in terms of 𝑙 (𝑡) instead of
(𝑡). This means that we rather assume a successful engagement as that 𝑙 → 0 during an assumed
engagement time, than assuming a (𝑡) and find out how long time it takes to engage (or fail to en-
gage. The 𝑙 (𝑡) is then declared as input and the (required) torque 𝑇(𝑡) and the torque capacity
(𝑡) becomes an output. This can simplify the modelling. This is not further discussed in this compen-
dium but see http://blog.xogeny.com/blog/part-2-kinematic/.
𝜀 | 𝑙|
max ( ) 𝑒 𝑠𝑒
𝜀 𝑛 𝑚 𝜀 𝑛 𝑚
+ 𝜀 (| )
{ 𝑙|
161
Vehicle Interactions and Subsystems
𝑇
𝑇
𝑓
𝑓
𝑇𝑙
𝑇𝑙
𝑙 𝑙
𝑛 𝑚 𝑛 𝑚
Figure 2-81: Example of approximation of clutch model. Strategies: 𝑛 𝑚 = typical slip speed of
clutch, 𝜀 ≪ . Dashed curve shows before approximation.
Figure 2-82 shows an example that the ideal and approximate models can give comparable results
with respect to torques and speeds. About computational efficiency, the ideal needs around 5 𝑠 time
step with Euler forward integration, while the approximate needs 100 times smaller time step. If bet-
ter agreement than in Figure 2-82 is needed, 𝜀 needs to be reduced, which slows down the approxima-
tion even more. Note also that, for energy dissipation, the approximate model of course calculates a
higher energy dissipation, since it assumes the clutch has to slip to transfer torque.
approxClutch.Aped approxClutch.c0
approxClutch.Aped
2 approxClutch.c0
2 𝐴 𝑒 𝑒𝑟 𝑡𝑜𝑟 𝑝𝑒
1
1 𝑢𝑡 ℎ 𝑒𝑛 𝑒 𝑒𝑛𝑡
0
0 0 2 4 6 8 10 12 14 16
0 2 4 6 8 10 12 14 16
approxClutch.w eng approxClutch.w gin idealClutch.w eng idealClutch.w gin
approxClutch.w
600eng approxClutch.w gin idealClutch.w eng idealClutch.w gin
600
550 /[𝑟 /𝑠]
𝑛 𝑛
550
500
500
450
450 /[𝑟 /𝑠]
𝑛
400
400
350
350
300
300
250
250
200
200 0 2 4 6 8 10 12 14 16
0 2 4 6 8 10 12 14 16
Modelled in two ways: approxClutch.vx idealClutch.vx
approxClutch.vx idealClutch.vx
• Approximation, 𝜀 0.05 20
/[ /𝑠]
• Ideal 20
0
0 0 2 4 6 8 10 12 14 16
There are clutch models also in the standard Modelica (see 1.5.4.5) library, see Figure 2-83. Note that
the library is built such that the “small inertia” is needed, which forces down computational efficiency
during clutch slip.
162
Vehicle Interactions and Subsystems
idealGear
engine engine_flywheel small_inertia body
spring damper
𝑇 𝑇𝑛
𝑇𝑛 𝜆 2
𝑛
stepped 1
e in out gearbox
(with planetary
lock-up gears) 1 1
clutch ⁄
𝜈 ⁄ 𝑛 𝜈 𝑛
Figure 2-84: Left: Traditional automatic transmission. Right: Conceptual curves of the torque
converter.
Fuel tank for fossil fuel or hydrogen, battery, flywheel, hydro static accumulator and super capacitor
are examples of energy storages. An energy buffer often refers to an energy storage that can not only be
emptied (during propulsion), but also refilled by regenerating energy from the vehicle during deceler-
ation. With that nomenclature, a fuel tank is an energy storage, but not an energy buffer. Also, a battery
which can only be charged from the grid, and not from regenerating deceleration energy, is not an en-
ergy buffer. Today, the most common energy storage is fossil fuel tank and the most common energy
buffer is battery.
Eq [2.57] gives a simple model of an energy buffer. Also the energy converters to mechanical rotation
(electric machines, if energy buffer is a battery) is modelled.
163
Vehicle Interactions and Subsystems
164
Vehicle Interactions and Subsystems
89
82
85
44
41
37
Electric
motor
m 5 2
𝑚
9
C1
4
3
ICE 8
e 1 C0 C2 6 B1 x 2
𝑅𝑤
7
Figure 2-85: § Hybrid propulsion system with powershifting, designed using planetary gears.
Upper left: Design. Upper right: Gear/Clutch schedule. Lower half: Dynamic model.
C0.c C0.Tleft wm Tm
400 1500
200
1000
0
500
-200
0 10 20 30 40 50 60 70
0
C1.c C1.Tleft
400 -500
0 10 20 30 40 50 60 70
200
0 we Te
-200
0 10 20 30 40 50 60 70 400
C2.c C2.Tleft 0
200 0 10 20 30 40 50 60 70
0 vx Fx_kN
20
0 10 20 30 40 50 60 70
15
B1.c B1.Tleft
10
400
200 5
0 0
-200 -5
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Figure 2-86: § Simulation of the transmission in Figure 2-85 (with approximate clutch models
from Eq [2.55]). Example sequence of shifts: Simply shift each 10th second, in order as in table in
Figure 2-85.
165
Vehicle Interactions and Subsystems
Design point 3
Design point 2
Design point 1
Figure 2-87: § Efficiency maps for a Single Drivetrain: 1 electric motor (20 kW), gearbox with 2
gears.
166
Vehicle Interactions and Subsystems
Figure 2-88: § Efficiency maps for a Multiple Drivetrains: 3 electric motors (3×10 kW), no
gearboxes. Efficiency maps for 3 Modes are shown.
The figure shows 3 “Modes”. A Modes is a selection of which drivetrains (which EMs) to use, and with
which distribution, to generate the required force on the vehicle. The 3 modes in the figure are de-
fined as:
• Mode 1: Use only Drivetrain 1
• Mode 2: Use Drivetrain 1 and 2, with equal distribution of force and power
• Mode 3: Use Drivetrain 1 only, if it is enough for the required . If more is needed, add as much
as needed from Drivetrain 2. If even more is needed, add as much as needed from Drivetrain 3.
Note that one can define modes in more advanced way, such as in Ref (Utbult, Jonasson, Yang, &
Jacobson, accepted 2020). Also note, that anyway the modes are defined, they constitute discrete se-
lections which reminds very much of gear selection.
167
Vehicle Interactions and Subsystems
o Electric machines (machines can be used symmetrical, i.e. both for positive and nega-
tive torques, see Figure 2-70)
• Heavy vehicles often have Retarders. They normally use hydraulic or Eddy current to dissipate
engine, as opposed to dry friction. So, they cannot brake at low speeds or stand-still.
• Large steer angles will decelerate the vehicle, see 3.2.2.2.
This section is about Friction brakes, meaning Service brakes and Parking brake. In vehicle dynamics
perspective, these have the following special characteristics:
• Friction brakes are almost unlimited in force for a limited time since they can lock the wheels
for most driving situations and road friction (ICE and electric motors are often limited by their
maximum power, since it is often smaller than available road friction.) However, if the friction
brakes are used for a long time, the brake lining will start to fade. This means friction coeffi-
cient is lowered due to high temperature (oxidation and melting of pad/lining material).
• Friction brakes can only give torque in opposite direction to wheel rotation. (Electric motors
can brake so much that wheel spins rearwards.)
• Friction brakes can hold the vehicle at standstill. (Using electric machines for stand-still in a
slope, a closed loop control would be necessary, resulting in that vehicle “floats” a little.)
The basic design of a passenger car brake system is a hydraulic system is show in Figure 2-89. Here,
the brake pedal pushes a piston, which causes a hydraulic pressure (pressure = pedal force/piston
area). The hydraulic pressure is then connected to brake callipers at each wheel, so that a piston at
each wheel pushes a brake pad towards a brake disc ( Disc orce Pressure PistonArea). The brake
torque on each wheel is then simply: 𝑇 NumberOf rictionSurfaces DiscCoefficientOf riction
Disc adius Disc orce. (Normally, there are 2 friction surfaces, since double-acting brake calipers.) By
selecting different piston area and disc radii at front and rear, there is a basic hydro mechanical brake
distribution ratio between front and rear axle. There are normally two circuits for redundancy. It
should be mentioned that DiscCoefficientOf riction varies a lot; during one strong brake event, it can
typically drop 10..25% due to temperature rise and sliding velocity decrease.
Brake systems for heavy trucks are generally based on pneumatics, as opposed to hydraulics, see Fig-
ure 2-91. Ref (Tagesson K. , 2017), has a good descriptive chapter about brake systems for heavy vehi-
cles.
II HI
Hydraulics
X (common)
Pistons with
brake pad Front Front
Brake
disc HH
LL
Hydraulic pump for
ABS/ESC and hydraulic Rear
valves are not drawn.
Figure 2-89: Layouts of a hydraulically applied brake system, which is conventional on passenger
cars.
Brake systems for modern road vehicles are almost always mechatronic systems, i.e. they contain both
mechanical parts and control algorithms. As minimum, one can include the wheel slip control, see
4.6.2.1.4, or ABS/EBD, see 3.5.2.3/3.5.2.4.
168
Vehicle Interactions and Subsystems
ABS intervention 1.
2.
Master Cylinder
Inlet Valves
ESC intervention
(Pedal braking. One wheel 3. Wheel Cylinders/Callipers (No pedal braking. One wheel
brake controlled=released) 4. Brake Discs brake controlled=applied)
5. Outlet Valves
6. Tank/Low Pressure
Accumulator
1. 7. Pump 1.
8. Check Valve
9. 9. Isolation Valve/Traction 9.
Control Valve
8. 8.
7. 7.
2. 5. 5.
5. 2.
2. 5.
6. 6.
2.
3. 3. 3. 3.
controlled = controlled =
= released 4. … 4.
= applied 4. … 4.
(several wheels, only 2 drawn) (several wheels, only 2 drawn)
Figure 2-90: Concept of hydraulically applied brake system for ABS and ESC functions.
Figure 2-91: Pneumatically applied brake system for heavy vehicles. Electronics Brake System,
EBS, from Volvo GTT, Mats Sabelström.
169
Vehicle Interactions and Subsystems
di do 1 turn centre
a Ackermann error, front.
b Ackermann error, rear.
1 tan(d o ) = 1 tan(d i ) + w L
L
Figure 2-92: Ackermann steering geometry. Left: One axle steered. Right: Both axles steered and
including “Ackermann errors”. From (ISO 8855).
In traditional steering systems, the steering wheel angle has a monotonically increasing function of the
steer angle of the two front axle road wheels. This relation is approximately linear with a typical ratio
170
Vehicle Interactions and Subsystems
of 15..17 for passenger cars. For trucks the steering ratio is typically 18..22. In some advanced solu-
tions, steering on other axles is also influenced (multiple-axle steering, often rear axle steering). There
are also solutions for dynamically adding steer angle through a planetary gear and electric angle-con-
trolled motor on the steering shaft, so called Active Front Steering (AFS). In reference (Tagesson K. ,
2017), there is a good descriptive chapter about steering systems for heavy vehicles.
Ackermann 𝑤
steering
𝑓 𝑟𝑝 𝑛
(trapezoidal Rack 𝑙
𝑓 𝑟𝑝 𝑛
𝑤
𝑤
geometry)
(Erasmus Darwin
1758, Rudolph
Ackermann 1810)
Tie Rod
Steering
Knuckle
Steering
View from top Axis
and rear:
𝑙
0
symmetry
when 𝑤
𝑅𝑆𝐴
𝑙
𝛾 𝛾
𝑅𝑆𝐴 𝑟𝑝 𝑛 𝑤 𝑟𝑝 𝑛 𝑤
Figure 2-93: Left: Ackermann (Trapezoidal) Steering. Right: Rack Steering, common on passenger
cars.
Eq [2.59] shows the relation between steering angles for rack steering, with 𝑇𝑅𝐿=Tie Rod Length,
𝑆𝐴𝐿=Steering Arm Length, 𝑅𝑆𝐴=Rack to Steering Axis lengths.
2
𝑇𝑅𝐿2 (𝑅𝑆𝐴 𝑆𝐴𝐿 cos(𝛾 + 2
𝑙 )) + (𝑅𝑆𝐴 𝑆𝐴𝐿 sin(𝛾 + 𝑙) + 𝑟𝑝 𝑛 ∙ 𝑤)
[2.59]
𝑇𝑅𝐿2 (𝑅𝑆𝐴 𝑆𝐴𝐿 cos(𝛾 ))2 (𝑅𝑆𝐴
+ 𝑆𝐴𝐿 sin(𝛾 ) 𝑟𝑝 𝑛 ∙ 𝑤)
2
The 𝑇𝑅𝐿 determines a toe-in. To get toe-in , design (oradjust) the 𝑇𝑅𝐿 to:
𝑇𝑅𝐿 √(𝑅𝑆𝐴 𝑆𝐴𝐿 cos(𝛾 2
toe-in )) + (𝑅𝑆𝐴 𝑆𝐴𝐿 sin(𝛾 toe-in ))
2 .
171
Vehicle Interactions and Subsystems
Rack
Force Assisted
Assistance
UnAssisted
SteWhlTq
Figure 2-94: Left: Boost Curve with different working areas depending on the driving envelope.
Middle: Torque distribution between manual torque, FM, and assisting torque, FA, depending on
applied steering wheel torque. From Reference (Rösth, 2007). Right: Unassisted and assisted
steering wheel torque.
forward
Fx1=Tb/R
Castor offset at
wheel center
(drawn positive)
Wheel 1 with
Fx1 shaft torque, Ts:
172
Vehicle Interactions and Subsystems
from the contact patch centre, both longitudinally and laterally. The offsets are called Castor offset at
ground and Steering axis offset at ground, respectively. (Steering axis offset at ground is sometimes
called Scrub radius, but that is also used for the resulting distance of both offsets, so it is an ambiguous
name.) This figure also defines Normal steering axis offset at wheel centre. We will use the 3 latter
measures to explain why steering is affected by differences in shaft torque left/right, differences in
brake torque left/right, and lateral wheel forces. The steering axis distance from wheel centre in a side
view, is called Castor offset at wheel centre.
Often, the actual forces between tyre and ground are not in the exact centre of the nominal contact
patch, which also creates “effective” version of castor offset and scrub radius. For instance, the Pneu-
matic trail adds to the Castor offset at ground due to the lateral force distribution being longitudinally
offset from nominal contact point; so that the lever becomes not only , but + 𝑡. The moment, 𝑇 ,
on steering system (around Steering axis, turning towards increasing steer angle) is affected by tyre
forces on a steered axle as in Equation below.
Castor offset at ground is, on passenger vehicles, 15-20 mm (at motorcycles approximately 100 mm).
On rear wheel driven passenger vehicles, it is typically 5 mm due to higher Castor Angle which gives a
beneficial higher Camber angle gain at cornering. On a front wheel driven passenger vehicle, a non-
zero castor offset is not chosen due to drive axle lateral displacement. Castor offset gives a self-aligning
steering moment, which generally improves the steering feel.
(𝑇 +𝑇 ) 𝑇
𝑇 ∙ 𝑠 cos(𝐾 𝐼) + 𝑅 sin(𝐾 𝐼) +
𝑅 𝑅
(𝑇 2 + 𝑇 2 ) 𝑇2
+ 𝑠 cos(𝐾 𝐼) + 𝑅 sin(𝐾 𝐼) +
𝑅 𝑅 [2.60]
+( ) ( ( + 𝑡 ) + 2 ( + 𝑡2 )) cos( 𝐴)
𝑇2 𝑇 𝑇2 𝑇
𝑠 cos(𝐾 𝐼) + 𝑘 ( ( +𝑡 )+ 2 ( + 𝑡2 )) cos( 𝐴)
𝑅 𝑅
The equation shows that difference in both brake torque and shaft torque affects steering and so does
the sum of lateral forces. For reducing torque steer and disturbances from one-sided longitudinal
forces due to road irregularities, kingpin offset, scrub radius different road friction should be as small
as possible, but it is limited by geometrical conflicts between brake disc, bearing, damper, etc.
Positive scrub radius contributes to self-centring, thanks to lifting the car body, see below. Negative
scrub radius compensates for split- braking, or failure in one of the brake circuits. Hence, the scrub is
a balance between these two objectives. Scrub radius is often slightly negative on modern passenger
cars. Scrub radius is often positive on trucks, maybe 10 cm, due to packaging.
It should be noted that the Eq is not complete with respect to all “returning effects”. There are also ef-
fects from Castor angle and Castor trail as well as that the tyre has a width and radius. However, in to-
tal, these give rise to a returning steering torque which is depending on the steer angle.
It can also be noted that steering effort for low speed or stand-still is largely influenced by whether
brakes are applied or not, due to the magnitude (not sign) of scrub radius.
173
Vehicle Interactions and Subsystems
𝑠 ′′ sin 𝐾 𝐼
for 𝒕𝒆𝒆𝒓 𝟎
View from
rear: s
KPI
𝑠 ′′′
Fx1
or 𝑅 or 𝑅
ℎ ≈ℎ ℎ 0
Aerodynamic reference point or 𝑅
Centre of gravity, CoG
Figure 2-97: Force-equivalent ways to model longitudinal-wind aerodynamic forces in 𝑥 -plane.
𝑓 and differs, but is the same for all.
From vehicle motion point of view, variables for aerodynamic forces and moments are naturally de-
fined as positive when acting on vehicle in directions defined positive by (ISO 8855). They are denoted
and 𝑀 in 2.8. So, > 0 when the wind pushes the vehicle forward. Consequently, <0
when the wind is resisting forward motion of the vehicle. The longitudinal force play is special in the
sense that it often appears in discussions where it is natural to name it “resistance force” (resisting for-
ward motion), which makes it natural to use the opposite sign definition. So, we use a double notation
on the longitudinal aerodynamic force: 𝑅 , see Eq [2.62].
∙ 𝜌 ∙ 𝐴𝑓 𝑛
𝑅 ∙ 2
𝑙 sign( 𝑙) [2.62]
174
Vehicle Interactions and Subsystems
The 𝑙 is the longitudinal component of the wind velocity relative to the vehicle. The parameters
𝜌 and 𝐴𝑓 𝑛 represent the drag coefficient, the air density and a reference area of the vehicle, re-
spectively. The 𝐴𝑓 𝑛 is the area of the vehicle projected on a vehicle transversal plane. The most com-
∙𝜌∙𝐴𝑓𝑟𝑜
mon case is relative headwind, which makes sign( 𝑙) , which makes 𝑅 2
∙
2
𝑙 > 0.
Typical values of drag coefficients ( ) for cars can be found from sources such as: (Robert Bosch
GmbH, 2004), (Barnard R. , 2010), (Hucho, 1998), and (Schuetz, 2015). These coefficients are derived
from coast down tests, wind tunnel tests or CFD (Computational Fluid Dynamics) calculations. The air
resistance can often be neglected for city speeds, but not at highway speeds.
Since a car structure moving through the air is not unlike an aircraft wing, there are also an aerody-
namic lift force and pitch moment. This affects the vertical forces on front and rear axle, and conse-
quently the tyre to road grip. Hence, it affects the lateral stability.
𝑙 ∙ 𝜌 ∙ 𝐴𝑓 𝑛
∙ 2 𝑙
[2.63]
𝑝𝑚 ∙ 𝐿 ∙ 𝜌 ∙ 𝐴𝑓 𝑛
𝑀 ∙ 2 𝑙 sign( 𝑙 )
The coefficient 𝑙 represents the lift characteristics of the vehicle. For extreme vehicle, such as racing
cars, one can achieve negative 𝑙 , but often by sacrificing with higher . The forces and are
assumed to act through the same reference point, often centre of gravity (CoG), which defines 𝑀 .
One can replace [ 𝑀 ] with equivalent [ 𝑓 ] or
[ 𝑓 ], Figure 2-97 and Eq [2.64].
𝑙𝑓 ∙ 𝜌 ∙ 𝐴𝑓 𝑛 2
𝑓 ∙ 𝑙
𝑙 ∙ 𝜌 ∙ 𝐴𝑓 𝑛 2 [2.64]
∙ 𝑙
For a reference height ℎ , often ℎ 0
The speed 𝑙 is the lateral component of the vehicle velocity relative to the wind. Note that 𝐴 and
𝐿 may now have other interpretations and values than in Equations [2.62]-[2.64], e.g. 𝐴𝑓 𝑛 or 𝐴 .
175
Vehicle Interactions and Subsystems
The coefficients 𝑚 are typically more independent of sign( 𝑙 ) > 0 due to approximate sym-
metriy in vehicle body shape across an 𝑥 -plane.
Some of the equations above in 2.8 are selected to include “sign(⋯ )”. The corresponding coeffi-
ciencts typically, thereby, retain their (positive) sign. In the other equations, the corresponding coef-
ficiencts may or may not change sign.
The coefficients 𝑚 are almost independent of sign( 𝑙) but they still vary with relative wind
yaw angle 𝑙.
2.9.2 Driver
To study how different vehicle designs work in a vehicle operation, a driver model is needed. In its eas-
iest form, a driver model can be steering wheel angle 𝑤 0 or 𝑤 ̂ sin( 𝑡) . Another ex-
treme interpretation of what can be called a driver model is an implicit/inverse statement, like “driver
will push accelerator pedal so that speed 0 [ ⁄𝑠] during the manouvre”, which leads to that ac-
celerator pedal position becomes an output, as opposed to input, to the vehicle model. Beyond those
very extreme driver models, there is often need for a driver model which react on vehicle states in re-
lation to an environment or traffic. In this section, driver models are primarily thought of as models
of the driver of the subject vehicle, but when modelling surrounding traffic carefully, each object vehi-
cle can also use a driver model.
The driver interacts with the vehicle mainly through steering wheel, accelerator pedal and brake pe-
dal. In addition to these, there are clutch pedal, gear stick/gear selector, and various buttons, etc., see
Figure 2-98, but we focus here on the first 3 mentioned.
176
Vehicle Interactions and Subsystems
177
Vehicle Interactions and Subsystems
178
Vehicle Interactions and Subsystems
𝑝𝑟𝑒𝑠𝑒𝑛𝑡 𝑡 𝑒
𝑡 𝑒
𝑜𝑛 𝑡𝑢 𝑛
𝒗𝒆𝒍𝒐𝒄𝒊𝒕𝒚
subject vehicle
object vehicle
𝑡 𝑒
𝑜𝑛 𝑡𝑢 𝑛
𝒂𝒄𝒄𝒆𝒍𝒆𝒓𝒂𝒕𝒊𝒐𝒏
𝑡 𝑒
179
Vehicle Interactions and Subsystems
lateral force on the front axle, reduced with assistance from the power steering or steering servo, but
almost independent of the angles.
A frequently used model is that the driver decides a steering wheel angle and expects a certain steer-
ing wheel torque, 𝑇 𝑤 , feedback. Only in very driver-active steering situations, the causality is modelled
the other way, e.g. studying what happens if driver takes hands-off, i.e. steering wheel torque =0. It can
be that driver uses the effect that steering system returns to 0 without driver torque after a quick
step steering by driver.
2.9.2.3.2 § Steering Wheel Angle Driver Models
How driver operates the steering wheel can also be called a lateral driver model. But different ma-
noeuvres call for different driver models. Examples of manoeuvres/situations are:
• Going into curve from straight and vice versa
• Turning in intersection
• Avoidance manoeuvre
• Stabilizing vehicle when sudden lose yaw stability, e.g. in curve
• Compensating road unevenness which cause lateral disturbance
• Lateral speed control during overtaking
2.9.2.3.3 § Example of Steering Wheel Angle Driver Model Based on Biology
An example of a driver model is the Salvucci and Gray model in Figure and Eq below. It is motivated
from human’s biological cognition. The steering wheel angle is denoted 𝑤.
nearpoint
𝑊/ 𝑓
𝑛
farpoint
𝑤 𝑘𝑓 𝑓 + 𝑘𝑛 𝑛 +𝑘 𝑛 ⇔
[2.67]
⇔ 𝑤 (𝑡) 𝑤 (𝑡 ) + 𝑘𝑓 𝑓 (𝑡) + 𝑘𝑛 𝑛 (𝑡) +𝑘 ∫
0
𝑛 (𝑡 . . 𝑡) 𝑡
It is not well defined in figure above what it means that vehicle “points”. It can be either vehicle head-
ing or vehicle course angle. Probably, the course angle, is most motivated to use. The course angle is
different for different points in the vehicle, so one might need to define also a “mind-point” in the vehi-
cle. In many cases the mind-point is the position of the driver.
The 3 𝑘 parameters that have to be tuned to certain driver, vehicle and operation. Additionally, the dis-
tance to aim-points needs to be tuned, such as if there should be extra margins on top of 𝑊/ and
whether 𝑊 is vehicle, road or lane width. Further on, all these assumptions can change during a simu-
lation, e.g. as driving on a windy country road and suddenly a meeting vehicle shows up might make
the far point jump. So, even if the model equation appears simple, it is far from obvious how to tune
and use it. Hence, there are variants of it which has only one aim-point. We will refer to the reduced
model, with only one aim-point:
This kind of model is fed with variables from the vehicle and environment states. A drawback is that
the dynamic vehicle behaviour is built into the 𝑘 parameters, meaning that if vehicle dynamics is
changed, e.g. adding a trailer or reducing road friction, the 𝑘 parameters reasonable have to be re-
180
Vehicle Interactions and Subsystems
tuned. If identifying the 𝑘 parameters as physical parameters, using a dynamic vehicle model, one can
reduce the need for such re-tuning.
Above driver model can be categorized as having “look ahead”. Simpler driver model can be based on
only where the vehicle is at present time, such as present lateral position in lane. More advanced
driver models use a dynamic model for prediction. The vehicle dynamics model is especially needed in
predicting models.
181
Vehicle Interactions and Subsystems
182
Longitudinal Dynamics
3 LONGITUDINAL DYNAMICS
3.1 Introduction
The primary purpose of a vehicle is transportation, which requires longitudinal dynamics. The chapter
is organised with one group of functions in each section as follows:
• 3.2 Steady State Function
• 3.3 Functions Over (Long)
• 3.4 Functions in (Short) Events
• 3.5 Control Functions
𝝎⁄𝒓𝒂𝒕𝒊𝒐
prime mover characteristics:
Traction Diagram
𝑻 𝑭
𝒓𝒂𝒕𝒊𝒐
Multiply torque by
𝒓𝒂𝒅𝒊
𝒓𝒂𝒅𝒊
Multiply rotational speed by
𝝎
𝒓𝒂𝒕𝒊𝒐 𝒗
183
Longitudinal Dynamics
A traction diagram for a truck is given in Figure 3-2, which also shows that there will be one curve for
each gear.
lowest gear
highest gear
Figure 3-2: Example of engine map and corresponding traction diagram map from a truck.
(D13C540 is an I6 diesel engine of 12.8-litre and 540 hp for heavy trucks.)
Figure 3-3: Relations between diagrams. The circle markers show alternative gears for one
example of downhill operating point, where neutral gear is possible.
Losses in transmission can be included by loss models for transmission, such as:
𝑇
𝜂𝑇 ∙ 𝑟 𝑡 𝑜 ∙ ℎ𝑒𝑟𝑒 𝜂 𝑇 ≤
𝑟 𝑢𝑠
𝜂𝜔 ∙ 𝑟 𝑢𝑠 ∙ ℎ𝑒𝑟𝑒 𝜂𝜔 ≤ [3.2]
𝑟 𝑡𝑜
𝑣 𝑙 ∙
ℎ𝑒𝑟𝑒 𝜂 𝑇 ∙ 𝜂𝜔 𝜂 𝑙 ≤
𝑛 𝑛 𝑇∙
This will move the curves in the first quadrant downwards due to ηT < and to the left due to 𝜂𝜔 < .
Tyre rolling friction is a torque loss mechanism, which on its own yields 𝜂𝜔 and ηT < . Tyre lon-
gitudinal slip is a speed loss mechanism, which on its own yields 𝜂𝜔 < and ηT . See 2.2.1.6. The
multiplication with 𝜂 is only demonstrative and should be seen more generic: in many cases the losses
are additional instead of multiplicative, e.g. for rolling resistance: 𝑟 𝑡 𝑜 (𝑇 𝑇)⁄𝑟 𝑢𝑠
𝑟 𝑡 𝑜 (𝑇 𝑓 )⁄𝑟 𝑢𝑠 . Which wheels’ to use is discussed in 3.2.2.
184
Longitudinal Dynamics
A traction diagram is a kind of “one degree of freedom graphical model”. The traction diagram is on
complete vehicle level, so the force axis represents the sum of forces from all wheels. This can include
more than one propulsion system and also brakes.
VehicleForceRequest
VehicleForceRequest
VehicleSpeed VehicleSpeed
0 0
Gear=Neutral Gear=Neutral
-small -small
Gear=1 Gear=2 Gear=3 Gear=1 Gear=2 Gear=3
Figure 3-4: § Gear selection diagram. Right: with hysteresis to avoid non-stationary and too
frequent shifts
The curves in the diagram are also depending on other conditions, such as road grade and vehicle
gross weight. The diagram only indicates sequential gears, while, especially heavy vehicle, often jump
over gears, e.g. shifting directly from gear 1 to 3. The energy optimality of neutral gear is explained in
Figure 3-3. In essence, the neutral gear eliminate torque losses in propulsion system which otherwise
would remain even in the operating condition of 𝑉𝑒ℎ 𝑒 𝑜𝑟 𝑒𝑅𝑒𝑞𝑢𝑒𝑠𝑡 ≈ 0.
3.2.1.1.2 Gear Selection for Transient Operating Conditions
The previous description was made for a steady-state operating condition. Transients, i.e. variations
over time, add the problem to avoid non-stable gear shifting, and especially that gear is shifted too fre-
quent. This is because each shift causes energy losses, but also other negative consequences such as
reduced driveability of the vehicle and lifetime of the propulsion system components.
Too frequent shifting can be avoided by hysteresis in different ways. One way is to define separate up-
and down-shift curves, see previous figure. The hysteresis is achieved by putting these curves much
enough apart. This can be characterized as using a memory.
Another way to avoid too frequent shifting, is to use a timer which explicitly forbids shifting within a
certain time after the last shift. Using the time partly overcome the problem that the area in the dia-
gram where neutral gear is optimal is very narrow.
So, there is a conflict between selecting neutral gear for low energy consumption and avoiding too fre-
quent gear shifts.
185
Longitudinal Dynamics
With a predictive cruise, 3.5.2.2.1, it is much easier to utilize the neutral gear, since the algorithm can
simply look in the prediction and find out if a too frequent gear shift is predicted or not. In this way
one can allow neutral gear much more often without the drawbacks of too frequent shifting.
186
Longitudinal Dynamics
2
𝑅 𝑅 𝑙𝑙 + ∙ ∙ sin( )+ ∙ ∙𝜌∙𝐴∙( 𝑤𝑛 )
𝑅 𝑙𝑙 ∑ 𝑓 𝑤 𝑙 [3.3]
non-driven wheels
𝑓 ℎ𝑒𝑒 𝑠 𝑟𝑒 𝑢𝑛 𝑟 𝑒𝑛: 𝑅 𝑙𝑙 𝑓∙ ∙ ∙ cos( )
As seen in Figure 3-5, the supply and resistance curves are drawn in same diagram. The resulting in-
tersection identifies the top speed of the vehicle. This is a stable point in the diagram, so the vehicle
condition at top speed is for steady state (no acceleration).
Figure 3-5 also shows that the acceleration can be identified as a vertical measure in the traction dia-
gram, divided by the mass. However, one should be careful when using the traction diagram for more
than steady state driving. We will come back to acceleration performance later, after introducing the
two effects “Load transfer” and “Rotating inertia effect”.
𝑙 ∑ (𝑇 (𝑅 )) ∑ (𝑇 𝑠 |𝑅 |) ∑ ( |𝑅 |)
𝑅
= :𝑁 = :𝑁 = :𝑁
2 2
∑ ( |𝑅 |) ≈ ∑
𝑅 𝑅
= :𝑁 = :𝑁
An example with a fore-aft-symmetric 2-axle vehicle, propelled on one axle gives:
02 2 2
𝑙 ≈ ( + )
𝑅 ⁄ ⁄ 𝑅
The same vehicle, but propelled equally much on both axles gives:
( ⁄ )2 ( ⁄ )2 2
𝑙 ≈ ( + )
𝑅 ⁄ ⁄ 𝑅
So, twice as much energy is lost due to longitudinal tyre slip if propelling on 1 instead of 2 axles.
When negative wheel torque, one can brake with friction brakes and then there is no energy loss. How-
ever, if braking with electric propulsion, the loss can be negative, meaning that energy is regenerated
to electric energy storage. If braking so much with electric propulsion that wheel rotates rearwards,
there would again be an energy loss, 𝑙 𝑇 𝑛𝑒 𝑡 𝑒 𝑛𝑒 𝑡 𝑒 > 0.
187
Longitudinal Dynamics
⁄ 2 2
𝑙 𝑓 𝑟
𝑅 (( ) +( ) )≈ ( ) [3.4]
𝐿 𝑟 𝐿 𝑓 𝑦
𝑅 is such that the additional propulsion force due to cornering is ≈ 𝑅 or the additional
power is ≈ 𝑅 . During a transport operation, the cornering in each time instant is typi-
cally described by two variables, e.g. ( 𝑅𝑝 ), but only the combined scalar measure 2⁄
𝑅𝑝 influ-
ences 𝑅 . Hence, we can plot the following graph:
𝑅
Figure 3-6: Left: Cornering Resistance Coefficient. Right: Required steer angle.
Vehicle data: 500 𝑘 𝐿 𝑓 . 5 𝑓 60 [𝑘𝑁/ ] 80 [𝑘𝑁/ ].
Notes:
• The model used above is not advanced enough to differ between which axle is driven. For such
purpose, one would need e.g. the model in Figure 4-17.
• Normal driving is often below 2 or 3 ⁄𝑠 2 , so the coefficient typically stays below 0.01. So, the
influence on energy consumption, during such “maximum normal” negotiation of corners, is
still of the same magnitude as rolling resistance coefficient 𝑅𝑅 ≈ 0.005. .0.0 0.
• For ideally tracking axles, see 2.2.6, 𝑓 → and → , which gives that 𝑓 → 0 and conse-
quently no power loss and no required propulsion force. Therefore, high cornering stiffness is
fuel efficient when cornering.
• When driving extreme cornering, such as driving as fast as possible in a circle on a test-track,
one will experience that the top speed is much lower than driving straight ahead. That is NOT
explained by [3.4]. It would require inclusion of a combined tyre slip model.
3.2.3.2 Gradeability *
Function definition: Grade-ability is the maximum grade that a vehicle is capable to maintain the forward mo-
tion on an uphill road at a certain constant speed, at a certain road friction level and with a certain load. (from
Reference (Kati, 2013))
For vehicles with high installed propulsion power per weight, the road friction can be limiting, but this
is not visualised in Figure 3-7. Since the speeds are higher than for start-ability, the air resistance can-
not be neglected.
188
Longitudinal Dynamics
Road slope of this curve Road slope of this curve (Grade-ability here limited by
( 𝑟 𝑒2 ) is the Grade- ( 𝑟 𝑒3 ) is the Grade-
ability for speed 2 ability for speed 3 propulsion, not road friction)
Longitudinal
force
𝑅 for
varying
road grades
Road slope of this curve
( 𝑟 𝑒 ) is the Down-
grade holding capability
for speed
0
0 3
Top Speed,
2 speed
Figure 3-7: How Top speed, Gradeability and Down grade holding capability is read-out from
Traction diagram.
189
Longitudinal Dynamics
Longitudinal
force, 𝑭𝒙
ICE on continuously variable ratio
(can keep power at maximum ICE power, 𝑇 𝑚 / )
m*g
Figure 3-9: Free Body Diagram for steady state vehicle. With ISO coordinate system, the road
gradient is positive when downhill. (Rolling resistance force on non-driven axles is included in .)
From the free-body diagram we can set up the equilibrium equations as follows and derive the formula
for load on front and rear axle:
Moment equilibrium, around rear contact with ground ():
∙ ∙ ( ∙ cos( ) + ℎ ∙ sin( )) 𝑅 ∙ ℎ 𝑓 ∙(𝑓+ ) 0 ⇒
∙ cos( ) + ℎ ∙ sin( ) ℎ
⇒ 𝑓 ∙ ∙ 𝑅 ∙
𝑓+ 𝑓+
[3.5]
Moment equilibrium, around front contact with ground ():
∙(𝑓+ ) ∙ ∙ ( 𝑓 ∙ cos( ) ℎ ∙ sin( )) 𝑅 ∙ ℎ 0 ⇒
𝑓 ∙ cos( ) ℎ ∙ sin( ) ℎ
⇒ ∙ ∙ +𝑅 ∙
𝑓 + 𝑓 +
For most vehicles and reasonable gradients, one can neglect ℎ ∙ sin( ) since it is ≪ | 𝑓 ∙ cos( )| ≈
| ∙ cos( )|.
190
Longitudinal Dynamics
For a vehicle which is driven only on one axle, it is only the normal load on the driven axle, 𝑣 𝑛,
that is the limiting factor:
min( 𝑝 ∙ 𝑣 𝑛) [3.6]
One realises, from Figure 2-15, that the rolling resistance on the driven axle works as a torque loss and
that the road friction limitation will be limiting 𝑇 𝑛 ∙ 𝑟 𝑡 𝑜 𝑒 ∙ 𝑣 𝑛 rather than limiting 𝑇 𝑛 ∙
𝑟 𝑡 𝑜. Expressed using the rolling resistance coefficient, 𝑓 𝑙𝑙 , gives:
𝑇𝑛 ∙𝑟 𝑡 𝑜
min( 𝑝 ∙ 𝑣 𝑛) min ( 𝑓 𝑙𝑙 ∙ 𝑣 𝑛 ∙ 𝑣 𝑛) [3.7]
𝑟 𝑢𝑠
This is shown in the traction diagram in Figure 3-10, where it should also be noted that the rolling re-
sistance curve only consists of the rolling resistance on the non-driven axles. See also Figure 2-15.
𝑇𝑛 𝑟 𝑡𝑜 𝑒
force, Fx
Longitudinal
𝑣 𝑛
𝑝
𝑟 𝑢𝑠
𝑇𝑛 𝑟 𝑡𝑜
𝑓 𝑙𝑙 𝑣 𝑛
𝑟 𝑢𝑠
Friction limit: 𝑣 𝑛
𝑅
Maximum 𝑅 0.5 𝜌 𝐴 2
available 𝑅 𝑙𝑙 𝑓 𝑙𝑙 𝑛 𝑣 𝑛
𝑅 𝑙 𝑝 sin
0
0 Speed, 𝒗
Figure 3-10: Traction diagram with Road Friction limitation and Driving Resistance curves.
Figure 3-11 shows how we find the start-ability in the traction diagram. There are two phenomena
that can limit start-ability: propulsion system or road friction. Also, in each case, we can theoretically
reach somewhat higher start-ability by allowing clutch or tyre to slip. However, in practice the start-
ability has to require “forward motion without significant slip in clutch or tyre”, because there will be a
lot of wear and heat in the slipping clutch or tyre. Hence, the lower curves in Figure 3-11 are used. The
reduction is however very small, since the resistance curves does not change very much in this speed
interval (the resistance curves in the figure have exaggerated slope; the air resistance can typically be
neglected for start-ability).
However, the energy loss (heat, wear) in clutch and tyre should be limited also during the starting
sequence. This can limit the start-ability more severely than the slope of the resistance curves, but it
cannot be shown in the traction diagram, since it is limited by energy losses in clutch or tyre, which is a
time integral of 𝑇 ∙ 𝑙 and 𝑇𝑤 𝑙 ∙ 𝑤 𝑙 . There can also be quite other limitations of start-ability,
such as deliberately limited engine torque at low gears, to save drive shafts.
191
Longitudinal Dynamics
force
Longitudinal
Longitudinal
is the start-ability with
spinning tyres.
Friction limit Road slope of
for this curve is the
varying start-ability with
rolling tyres. 𝑅 for
road slope varying
Road slope of this curve is road slope
the start-ability (ending
with slipping clutch). Friction
limit
lowest gear lowest gear
Road slope of this curve is the start-
ability (ending with engaged clutch).
0 0
0 Speed, 𝒗 0 Speed, 𝒗
m*g
Figure 3-12: Towing Loads. The towing vehicle is front axle driven.
192
Longitudinal Dynamics
80 km/h
Figure 3-13: New European Driving Cycle (NEDC). From (Boerboom, 2012)
193
Longitudinal Dynamics
80 km/h
10 min
Time, s
T/T(max), %
Time, s
194
Longitudinal Dynamics
Figure 3-16: Top: WLTC cycle from http://www.unece.org, from 2015, for light vehicles.
Bottom: European Transient Cycle (ETC) for heavy commercial vehicles. The engine variants, 𝑇 ( )
and (𝑡), are used for certification from 2000. https://www.dieselnet.com/standards/cycles/etc.php
Figure 3-17: FTP cycle converted to a Driving pattern, i.e. a distribution of operating duration in
speed and acceleration domain. From http://www.epa.gov/oms/regs/ld-hwy/ftp-rev/ftp-tech.pdf.
Driving patterns can use more than 2 dimensions, such as [speed, acceleration, road gradient]. In prin-
ciple, they can also use less than 2 dimensions, maybe only [speed]. The (steady state) vehicle model
has to reflect the corresponding dimensions.
195
Longitudinal Dynamics
𝑇
𝑤 ∙
𝑤 ∙ 𝑇 ∙ 𝑅𝑤 [3.8]
𝑅𝑤 ∙ ⇒ 𝑅𝑤 ∙
Eliminate and gives:
𝑤 𝑤
( 𝑤 + 2
𝑤 ⁄𝑅𝑤 )∙ 𝑇⁄𝑅𝑤 or
2 [3.9]
𝑘 𝑤 ∙ 𝑇⁄𝑅𝑤 ℎ𝑒𝑟𝑒 𝑘 𝑤 𝑤 + ⁄𝑅𝑤
Figure 3-18: Rolling Note that we can keep and : ⁄𝑅𝑤 and 𝑇⁄(𝑘 𝑅𝑤 ) .
wheel
So, the rotating inertia makes the mass 𝑤 appear a factor 𝑘 larger and the reaction force (ex-
pressed in 𝑇) correspondingly lower. We call the factor 𝑘 the “rotational inertia coefficient”.
A vehicle with total mass will appear to have larger mass due to inertias in the propulsion system.
There are rotational inertias at two places: before transmission, i.e. rotating with same speed as en-
gine: and after transmission, i.e. rotating with same speed as the wheel: 𝑤 . The appearant mass, 𝑘 ∙
, will be dependent on the main transmission ratio 𝑟 as well:
𝑟 𝑤 ∙ 𝑟2
𝑘∙ ∙ 𝑇 ℎ𝑒𝑟𝑒 𝑘 ∙ + 2
+ 2
[3.10]
𝑅𝑤 𝑅𝑤 𝑅𝑤
Typically for a passenger car with traditional ICE propulsion, 𝑘 . . . .4 in the first gear and 𝑘 .
in the highest gear ( ≈ 0. [𝑘 2]
𝑟≈ 4 4 6 𝑅𝑤 0. [ ]). So, the phenomena is significant!
Typically for electrical propulsion of same vehicle, is smaller and there are fewer gears, so 𝑘 is lower
for low speed and higher for high speed.
We can now learn how to determine acceleration from the Traction Diagram, see Figure 3-19.
Propulsion
force,
Figure 3-19: Acceleration in Traction Diagram. Rotating inertia effects are shown assuming that
the engine is operated on its maximum curve and gear selection and clutch operation are made for
highest acceleration.
When the clutch is slipping, there is no constraint between engine speed and vehicle, so the term with
disappears from Equation [3.10]. If the wheel spins, both terms and 𝑤 disappear. It should there-
fore be noted that, if tyre slip is modelled, the effect of rotating inertia is regarded without using the 𝑘
factor above. That is, if the mathematical model in [3.11] is used instead of [3.8]. Eq [3.11] gives an ex-
plicit form model with two states [ ] and it becomes increasingly computational demanding the
larger is. (Similar decoupling between inertias and 𝑤 happens if torque converter or elastic
driveshafts are modelled between engine inertia and wheel.)
196
Longitudinal Dynamics
∙
∙ 𝑇 ∙ 𝑅𝑤
𝑅𝑤 [3.11]
𝑠 ℎ𝑒𝑟𝑒 𝑠
|𝑅𝑤 |
Gear 2
Speed,
Reverse gear
Braking friction limit,
Figure 3-20: Traction Diagram in 4 quadrants. One of two axles is assumed driven, which limits
propulsion to ≈half of braking friction limit. Up-hill slope is assumed, which is seen as asymmetric
resistance.
197
Longitudinal Dynamics
The final accumulated consumption [in kg or J] is often divided by the total covered distance in the
driving cycle, which gives a value in kg/km or J/km. If the fuel is liquid, it is also convenient to divide
by fuel density, to give a value in litre/(100*km). It can also be seen as a measure in 2, which is the
area of the “fuel pipe” which the vehicle “consumes” on the way.
For hybrid vehicles (with energy buffers) the same driving cycle can be performed (same (𝑡)) but
ending with different level of energy in the buffer. Also, the level when starting the driving cycle can be
different. This makes it unfair to compare energy consumption only as fuel consumption, one should
rather weight it to energy cost, €/𝑘 or €/(ton km).
3.3.4.2.1 Forward and Backward Simulation
We should note that the calculation scheme in Equation [3.12] does not always guarantee a solution.
An obvious example is if the driving cycle prescribes such high accelerations at such high speeds that
the propulsion system is not enough, i.e. we end up outside maximum torque curve in engine diagram.
This is often the case with “inverse dynamic analysis”, i.e. when acceleration of inertial bodies is pre-
scribed, and the required force is calculated. An alternative is to do a dynamic analysis, which means
that a driver model calculates the pedals in order to follow the driving cycle speed approximately, but
not exactly. Inverse dynamic analysis is often more computational efficient, but limits what can phe-
nomena that can be modelled in the propulsion and brake system. The computational benefit is espe-
cially large if state variables can be omitted, which is often the case but not always. Inverse dynamics
and dynamic simulations are sometimes referred to as backward and forward simulation, respectively;
see Reference (Wipke, Cuddy, & Burch, 1999).
§ Causality
198
Longitudinal Dynamics
The causality (data flow directions) in an Explicit form model changes a lot if changing between For-
ward and Backward simulation. However, the Mathematical model experiences minimal changes, basi-
cally only the parameter 𝑡 𝑛 changes from 𝑡 𝑛 > 0 to 𝑡 𝑛 0, i.e. the driver can be seen as “ideal”,
i.e., in some “magic” way, making the vehicle follow what the Environment stipulates: . The Envi-
ronment is where the Driving Cycle (𝑡) or (𝑠) is modelled. A big difference is that the Backward
simulation can often lack solution, e.g. if propulsion system or brake system is weaker than driving cy-
cle demands.
Forward simulation:
”Non-ideal Driver” Vehicle
For instance:
Environ- 𝑒
ment 𝐷 𝑣 Driver
𝑡𝐷 𝑣 Actuators
Interpretator
𝑒 𝑓𝑢𝑛 𝐷 𝑣
with 𝑡𝐷 𝑣 > 0
Backward simulation:
”Ideal Driver” Vehicle
For instance:
Environ- 𝑒
ment 𝐷 𝑣 Driver
𝑡𝐷 𝑣 Actuators
Interpretator
𝑒 𝑓𝑢𝑛 𝐷 𝑣
with 𝑡𝐷 𝑣 0
Figure 3-21: § Data flow diagrams showing causality for Forward and Backward simulation.
Backward leads to changes marked with yellow and that integration ∫ 𝑡 changes to
differentiation ⁄ 𝑡 in Vehicle.
3.3.4.2.2 Stepped, CVT and Energy Buffering Main Transmissions
For a given driving cycle (𝑡) there is a certain [ ] for each time instant. For different types of
main transmissions, one can select operating point [𝑇 ] in prime mover diagram differently, see
Figure 3-22. CVTs (or tightly stepped transmissions) enables operation close to optimal curve of en-
gine. However, CVTs typically have more losses than stepped transmissions. So, for commercial vehi-
cles, a higher number of gears is more realistic than real CVTs.
As mentioned in Figure 3-22, one can operate in the optimal point by an energy buffer or by propelling
intermittent with a varying vehicle speed. The latter case can be seen as if the vehicle’s own kinetic en-
ergy ( 2⁄
) is used to buffer energy, see (Hedman, 1994) which shows fuel savings in the magni-
tude of 10%. Intermittent operation is on market as part of predictive cruise, see 3.5.2.2.1.
199
Longitudinal Dynamics
𝑙
optimal curve for varying
(reachable with CVT)
Figure 3-22: Conceptual difference between Stepped, Continuously Variable and Energy Buffering
transmission. Operating intermittent would only follow the driving cycle with same average speed.
3.3.4.4 Emissions *
Function definition: As Energy consumption but amount of certain substance instead of amount of energy.
There are emission maps where different emission substances (NOx, HC, etc.) per time or per pro-
duced energy can be read out for a given speed and torque. This is conceptually the same as reading
out specific fuel consumption or efficiency from maps like in Figure 2-71 and Figure 2-72. A resulting
value can be found in mass of the emitted substance per driven distance.
Noise is also sometimes referred to as emissions. It is not relevant to integrate noise over the time for
the driving cycle, but maximum or mean values can have relevance. Noise emissions are very periph-
eral to vehicle dynamics.
200
Longitudinal Dynamics
3.3.4.6 Range *
Function definition: Range [km] is the inverted value of Energy consumption [kg/km, litre/km or J/km], and
multiplied with fuel tank size [kg or litre] or energy storage size [J].
The range is how far the vehicle can be driven without refilling the energy storage, i.e. filling up fuel
tank or charging the batteries from the grid. This is in principle dependent on how the vehicle is used,
so the driving cycle influences the range. In principle, the same prediction method as for energy con-
sumption and substance emissions can be used. In the case of predicting range, you have to integrate
speed to distance, so that you in will know the travelled distance.
201
Longitudinal Dynamics
Longitudinal
force, 𝑭𝒙
force, 𝑭𝒙
propulsion system propulsion system
Mass∙Acceleration
Mass∙Acceleration gear 1 Reserve on gear 1
Reserve on gear 1 Mass∙Acceleration
gear 1 Mass∙Acceleration Reserve on gear 2
Reserve on gear 2
Short term
Acceleration
gear 2 Reserve
gear 2
0 0
0 Speed, 𝒗 0 Speed, 𝒗
Figure 3-23: Acceleration reserves for different gears. Large dots mark assumed operating points.
x
y
Figure 3-24: Free Body Diagram for accelerating vehicle. Rolling resistance in 𝑓 and .
Note that the free-body diagram and the following derivation is very similar to the derivation of Equa-
tion [3.5], but we now include the fictive force ∙ .
Moment equilibrium, around rear contact with ground:
𝑓 ∙𝐿+ ∙ ∙ ( ∙ 𝑜𝑠( ) + ℎ ∙ 𝑠 𝑛( )) 𝑅 ∙ ℎ ∙ ∙ℎ 0 ⇒
∙ 𝑜𝑠( ) + ℎ ∙ 𝑠 𝑛( ) ℎ ℎ
⇒ 𝑓 ∙( ∙ ∙ ) 𝑅 ∙
𝐿 𝐿 𝐿
[3.13]
Moment equilibrium, around front contact with ground:
+ ∙𝐿 ∙ ∙ ( 𝑓 ∙ 𝑜𝑠( ) ℎ ∙ 𝑠 𝑛( )) 𝑅 ∙ ℎ ∙ ∙ℎ 0 ⇒
𝑓 ∙ 𝑜𝑠( ) ℎ ∙ 𝑠 𝑛( ) ℎ ℎ
⇒ ∙( ∙ + ∙ )+𝑅 ∙
𝐿 𝐿 𝐿
These equations confirm what we know from experience, the front axle is off-loaded under accelera-
tion with the load shifting to the rear axle. The opposite occurs under braking.
202
Longitudinal Dynamics
The load shift has an effect on the tyre’s grip. If one considers the combined slip conditions of the tyre
(presented in Chapter 2), a locked braking wheel limits the amount of lateral tyre forces. The same is
true for a spinning wheel. This is an important problem for braking as the rear wheels become off-
loaded. This can cause locking of the rear wheels if the brake pressures are not adjusted appropriately.
See more in 3.4.4.
If the problem is instead a “natural causality” problem, e.g. simulation of acceleration. Then is given
𝜕
(independent) and is dependent of 𝑅 , via 𝑚
𝑖𝑟
hence 𝜕 𝑟𝑧 𝑖𝑟
𝐿
.
𝑖𝑟
So, depending on which of these engineering problem one has in mind, it can either be true that 𝑅
influences only if ℎ ≠ 0 or only if ℎ ≠ ℎ.
𝑓 ≈ ≈ ⁄𝑅𝑤 ℎ
⇒ 𝑓 ≈{ }≈ ( ∙( ))
𝑓 ≈ ≈ 𝐿 ℎ 𝑅𝑤
With realistic values for a passenger vehicle, ≈ 500 𝑘 and 0.5 𝑘 2
, the influence from
wheel rotational inertia is ( )⁄( ℎ 𝑅𝑤 ) ( 0.5)⁄( 500 0.5 0. ) ≈ 0.005 0.5% on the
load transfer which in most cases can be considered as neglectable. In a strong take-off or brake apply,
when ≫ ⁄𝑅𝑤 , the equation would give a much higher influence but just for a short while. To cap-
ture the short transient in such cases, one should model also suspension as in 3.4.5.2.2.
″ 2)
∙ cos( ) + h ∙ sin( ) ℎ ℎ
𝑓 ∙ (( + ∙ ∙ ) 𝑅 ∙
𝐿 L L
[3.14]
″ 2) 𝑓 ∙ cos( ) h ∙ sin( ) ℎ ℎ
∙ (( + ∙ + ∙ )+𝑅 ∙
𝐿 𝐿 𝐿
2
Assuming that we have the road as (𝑠), then arctan ( ) ≈ and ″
2 .
Note that this model is assuming that vertical variations of road are much larger than wheelbase and
track width and same on left and right side of the road/vehicle. Else the variation would be called road
unevenness, which will be more treated in Chapter 5.
203
Longitudinal Dynamics
If models with body vertical and pitch motion and suspension springs, such as in 3.4.5 and 3.4.5.2 it is
often suitable to express the vertical fictive force, ̈ with instead of ″ 2
. The fictive
force downwards will then be 2
𝜅 instead. This can be understood from basic
geometry, ″ ≈ 𝜅 , where 𝜅 is the road pitch curvature [ ⁄ 𝑒𝑛 𝑡ℎ], see Figure 3-25.
crest sag
s
𝑦
′ ′
s s
• where ′ is a function of s.
• Vertical acceleration in inertial coordinate
system, ̈ ≈ 2 ,
̈
2
• where 2 is a function of s.
Figure 3-25: Free Body Diagram for driving over non-flat vertical road profile.
3.3.6 Acceleration
Acceleration performance like, typically, 0-100 km/h over 5..10 s, will be addressed in this section.
These accelerations are relatively steady state (vehicle pitch and heave are relatively constant), so the
suspension compliance is not considered.
Accelerations will also be covered in 0, as being shorter events. The vehicle pitch and heave vary more
and, consequently, the suspension compliance becomes important to model. This modelling is also
more suited for braking, which typically involve suspension more than propulsion.
204
Longitudinal Dynamics
Fx=mu*Ffz;
ax=(Fx-Rres)/m;
end
ax1=ax; %(clutch engaged)
ratio=ratios(2);
… then similar as for gear 1
ax2=ax;
%if gear 3 (clutch engaged)
ratio=ratios(3);
… then similar as for gears 1 and 2
ax3=ax;
%if clutch slipping on gear 2
ratio=ratios(2);
wc=vx*ratio/radius; %speed of output side of clutch
Te=max(Engine_T);
we=Engine_w(find(Engine_T>=Te)); %engine runs on speed where max torque
Fx=Te*ratio/radius;
ax=(Fx-Fres)/(m+Jw/(radius^2));
Ffz=m*(g*lr/L-ax*h/L);
if Fx>mu*Ffz
Fx=mu*Ffz;
ax=(Fx-Rres)/m;
end
if wc>we %if vehicle side (wc) runs too fast, we cannot slip on clutch
ax=-inf;
end
ax0=ax;
[ax,gear_vec(i)]=max([ax0,ax1,ax2,ax3]); vx_vec(i+1)=vx+ax*dt;
end
Phenomena that are missing in this model example are:
• Gear shifts are assumed to take place instantly, without any duration
• The option to use slipping clutch on 1st and 3rd gear is not included in model
• The tyre slip is only considered as a limitation at a strict force level, but the partial slip is not
considered for simplification. The code line “we=vx*ratio/radius;” is hence not fully cor-
rect. Including the slip, the engine would run at somewhat higher speeds, leading to that it
would lose its torque earlier, leading to worse acceleration performance.
• Load transfer is assumed to take place instantly quick; delays due to Suspension compliance, as
described in 3.3, are not included.
205
Longitudinal Dynamics
Engine 4
Traction diagram
x 10
2
Torque[Nm] gear 1
160
Power[kW] 1.8 gear 2
gear 3
140 1.6
resistance
slip limit
120 1.4
1
80
0.8
60
0.6
40
0.4
20
0.2
0 0
0 100 200 300 400 500 600 700 0 5 10 15 20 25 30 35 40
rotational speed [rad/s] speed [m/s]
Simulation result
14
12
10
vx [m/s]
6
gear (0=clutch slips on gear 2)
0
0 1 2 3 4 5 6 7 8 9 10
t [s]
Figure 3-26: Example of simulation of acceleration, using the code in Equation [3.15].
If simple mathematic functions can be used to describe ( ) and ( ) the solution can even be on
closed form.
206
Longitudinal Dynamics v_xReq v_x a_x
30
′ -20
∙ 𝑅 𝑅 -10 0 10 20 30
𝑅 𝑅 (𝑠 ) 𝑅 𝑅 (𝑠 ) 0
s
10
t [s]
20 30
𝑠 𝑡′ ⁄
200
′ s
0
200
0 10 20 30
0 t [s]
207
Longitudinal Dynamics
208
Longitudinal Dynamics
𝑙 𝑙𝑓
𝑓 ∙ ∙(𝑟 ) and ∙ ∙( + )
𝐿 𝐿 𝐿 𝐿
or, if eliminating : 𝑟
𝑓𝑥 √𝑟 4 ℎ 𝐿
𝑓𝑥
[3.16]
ℎ ℎ
𝑙𝑟 𝑙𝑓
or: 𝑓 ⁄
𝐿 𝑚 𝐿
and ⁄
𝐿
+
𝑚 𝐿
209
Longitudinal Dynamics
⁄𝑁
Conceptual purpose
with pressure limiting
valve or EBD
𝑓 ⁄𝑁
Figure 3-28: Brake Proportioning diagram. The curved curves mark optimal distribution for some
variation in position of centre of gravity.
The proportioning is done by selecting pressure areas for brake calipers, so the base proportioning
will be a straight line, marked as “Hydrostatic brake proportioning”. For passenger cars, one typically
designs this so that front axle locks first for friction below 0.8 for lightest vehicle load and worst vari-
ant. For heavier braking than 0.8 , or higher (or front-biased) centre of gravity, rear axle will lock
first, if only designing with hydrostatic proportioning.
To avoid rear axle lock-up, one restricts the brake pressure to the rear axle. This is done by pressure
limiting valve, brake pads with pressure dependent friction coefficient or Electronic Brake Distribu-
tion (EBD). In principle, it bends down the straight line as shown in Figure 3-28. With pressure de-
pendent values one gets a piece-wise linear curve, while pressure dependent friction coefficient gives
a continuously curved curve. EBD is an active control using same mechatronic actuation as ABS. EBD is
the design used in today’s passenger cars, since it comes with ABS, which is now a legal requirement
on most markets.
On heavy vehicles with EBS (Electronic Brake System) and vertical axle load sensing, the brake pres-
sure for each axle can be tailored. For modest braking (corresponding to deceleration ≤ ⁄𝑠 2) all
axles are braked with same brake torque, to equal the brake pad wear which is importance for vehicle
maintenance. When braking more, the brake pressure is distributed more in proportion to each axle’s
vertical load.
210
Longitudinal Dynamics
Ideal curve
ECE regulation
limits to this region
Figure 3-29: Brake Proportioning. From (Boerboom, 2012). If looking carefully, the “HydroStatic”
curve is weakly degressive, thanks to brake pad material with pressure dependent friction
coefficient.
Suspension model with “trivial suspension”
𝑥 𝑓
ℎ
ℎ
0 𝑓 0 𝑓
𝑓
𝐿
𝑥 𝑓 , are displacements from a Quasi steady-state assumed, so that longitudinal
static stand-still position. acceleration ( ) may be non-zero, but vertical
𝑓 0 means that road is smooth. and pitch acceleration are zero.
Figure 3-30: Model for steady state heave and pitch due to longitudinal wheel forces.
There is no damping included in model, because their forces would be zero, since there is no displace-
ment velocity, due to the steady-state assumption. As constitutive equations for the compliances
(springs) we assume that displacements are measured from a static condition and that the compli-
ances are linear. The road is assumed to be smooth, i.e. 𝑓 0.
𝑓 𝑓 + 𝑓∙( 𝑓 𝑓) 𝑛 + ∙( )
[3.17]
ℎ𝑒𝑟𝑒 𝑓 + ∙ 𝑛 𝑓 ∙ 𝑓 ∙ 0
211
Longitudinal Dynamics
We see already in free-body diagram that 𝑓 and always act together, so we rename 𝑓 +
𝑤 , where w refers to wheel. The assumption of “trivial linkage” explains how longitudinal forces are
transferred between wheels and body. Equilibrium then gives:
∙ + 𝑤 0
∙ 𝑓 0 [3.18]
∙ 𝑓 ∙ 𝑓 𝑤 ∙ℎ ∙ (ℎ ℎ) 0
Compatibility, to introduce body displacements, and 𝑝 , gives:
𝑓 𝑓 ∙ 𝑛 + ∙ [3.19]
Solving constitutive relations, equilibrium, compatibility using Matlab Symbolic toolbox gives:
>> clear, syms zf zr Ffz Frz Ffz0 Frz0 ax z py
>> sol=solve( ...
'Ffz=Ffz0-cf*zf', ...
'Frz=Frz0-cr*zr', ...
'Ffz0+Frz0=m*g', ...
'Ffz0*lf-Frz0*lr=0', ...
'-Fair-m*ax+Fxw=0', ...
'm*g-Ffz-Frz=0', ...
'Frz*lr-Ffz*lf-Fxw*h-Fair*(hair-h)=0', ...
'zf=z-lf*py', ...
'zr=z+lr*py', ...
zf, zr, Ffz, Frz, Ffz0, Frz0, ax, z, py);
The solution is given in Eq [3.20]:
𝑤
𝑓 ∙ 𝑓 ∙
∙( 𝑤 ∙ℎ+ ∙ (ℎ ℎ )) [3.20]
𝑓 ∙ ∙ 𝐿2
𝑓 +
𝑝 ∙( 𝑤 ∙ℎ+ ∙ (ℎ ℎ ))
𝑓 ∙ ∙ 𝐿2
In agreement with intuition and experience the body dives (positive pitch) when braking (negative
𝑤 ). Further, the body centre of gravity is lowered (negative z) when braking and weaker suspension
front than rear ( 𝑓 ∙ 𝑓 < ∙ ), which is normally the chosen design for cars.
The air resistance force is brought into the equation. It can be noted that for a certain deceleration,
there will be different heave and pitch depending on how much of the decelerating force that comes
from air resistance and from longitudinal wheel forces. But, as already noted, heave and pitch does not
depend on how wheel longitudinal force is distributed between the axles.
212
Longitudinal Dynamics
useful and not much more computational demanding (and probably easier intuitively), that model is
prioritized in this compendium.
h-hPC
h-hPC
PC
lPCf lr lf
lPCf
PC=Pitch Centre
Figure 3-31: Models for including suspension effects in longitudinal load transfer
3.4.5.2.2 Model with Axle Pivot Points
Behold the free-body diagram in Figure 3-32. The road is assumed flat, 𝑓 0. The force play in
the rear axle is shown in more detail. and are reaction forces in the pivot point. The is the
force in the elasticity, i.e. where potential spring energy is stored. The torque 𝑇 is the shaft torque, i.e.
from the propulsion system. Both torque from propulsion and brake system contribute to . But
torque from friction brake 𝑇 is not visible in free-body diagram, unless decomposed in suspension
and wheel, as in the right-most part of Figure 3-32. This is because the friction braking appears as in-
ternal torque (or, depending on the brake design, probably forces) between brake pad and brake calli-
per. Any part of the longitudinal wheel force that is applied via reaction torque to unsprung parts will
not add to shaft torque, such as an electric motor mounted on unsprung parts or propulsion via longi-
tudinal propeller shaft to a final gear (as usual for rigid propelled axles).
The term is easy to forget, but it does influence especially when is large. The term can be ex-
plained as the other centripetal accelerations (giving centrifugal (fictive) forces) in 4.4.2.3. The term
is generally much smaller. Both the terms appear due to that velocities and accelerations are
expressed as components in the vehicle fix (and hence rotating) coordinate system. We could intro-
duce also velocities and accelerations in ground fix coordinate system, with subscripts differing be-
tween [ 𝑣 𝑣 ] and [ ] and between [ 𝑣 𝑣 ] and [ ], in a similar way shown in
4.4.2.3.3 for yaw rotation where it is much more important to differ between vehicle and ground fix.
We assume that displacements are measured from the forces 𝑓 and , respectively, and that the
compliances are linear. The total constitutive equations become:
𝑓 𝑓 + 𝑓 ∙( 𝑓 𝑓) 𝑛 + ∙( ) [3.21]
Now, there are two ways of representing the dynamics in spring-mass systems: Either as second order
differential equations in position or first order differential equations in velocity and force. We select
the latter, because it is easier to select suitable initial values. Then we need the differentiated versions
of the compliance’s constitutive equations:
𝑓 0+ 𝑓 ∙( 𝑓 𝑓) 𝑓 ∙ 𝑓 𝑛 0+ ∙( ) ∙ [3.22]
213
Longitudinal Dynamics
Decomposed in smaller
Rear axle: free body diagrams, to
+ explain 𝑇 and 𝑇 .
𝑓 𝑓
Suspension:
+
ℎ 𝑇
𝑒𝑓 𝑒
Wheel:
0 𝑓 0
0 𝑓 0 𝑓 𝑇
𝑓
𝑓
𝐿 𝑓
𝑓 𝑓 ∙( 𝑓 𝑓) 𝑓 ∙ 𝑓 𝑛 ∙( ) ∙ [3.23]
Now, 3 equilibria for whole vehicle and 1 moment equilibria for each axle around its pivot point gives:
∙ + 𝑓 + 0 ( 𝑡ℎ ≈ )
∙ ∙ + 𝑓 + 0 ( 𝑡ℎ )
∙ + ∙ 𝑓 ∙ 𝑓 ( 𝑓 + )∙ℎ 0 [3.24]
( )∙ ∙𝑒 +𝑇 0
( 𝑓 + 𝑓 𝑓 )∙ 𝑓 𝑓 ∙ 𝑒𝑓 + 𝑇 𝑓 0
It can be noted that the unsprung parts are considered massless. Also note, that a “trivial suspension
model” (as used in Eq [3.17] and see also Figure 2-61) is a special case which falls out from the equa-
tions by letting → , so that + . With such trivial suspension, there is no difference in
vehicle motion if actuation is done with shaft torque 𝑇 or via reaction to unsprung parts, 𝑇 .
Equilibria for wheels on each axle:
𝑇 +𝑇 𝑅𝑅 𝑅𝑤
𝑅𝑤 𝑤 0
{ ⇒
𝑇𝑓+𝑇 𝑓 𝑓 𝑓𝑤 𝑓 𝑅𝑅 𝑓 𝑅𝑤
𝑅𝑤 0
[3.25]
neglect rotational inertia 𝑇 +𝑇 𝑅𝑤 0
⇒{ }⇒ {
neglect rolling resistance 𝑇𝑓 +𝑇 𝑓 𝑓 𝑅𝑤 0
Compatibility, to connect to body displacements, and , gives:
𝑓 𝑓 ∙ 𝑛 + ∙
𝑓 𝑓 ∙ 𝑛 + ∙ [3.26]
𝑛
By combining constitutive relations, equilibrium and compatibility we can find explicit function so:
• 𝑆𝑡 𝑡𝑒𝐷𝑒𝑟 𝑡 𝑒𝑠 𝑥𝑝 𝑡 𝑜𝑟 𝑢𝑛 𝑡 𝑜𝑛(𝑆𝑡 𝑡𝑒𝑠 𝐼𝑛𝑝𝑢𝑡𝑠)
• States: [ 𝑓 ]
• State Derivatives: [ 𝑓 ] (if wheel rotational inertias neglected)
• Inputs: [𝑇 𝑓 𝑇 𝑇 𝑓 𝑇 ] or [ 𝑓 𝑇𝑓 𝑇 ]
The ExplicitFormFunction can be integrated with well-established methods for numerical ODE. Such
simulation of is shown in 3.4.8.1. Note that the model is linear.
214
Longitudinal Dynamics
%Results:
% sol.ax = (Ffx+Frx)/m
% sol.z = -(Tsr*cf*gf*lf^2 -Tsf*cr*gr*lr^2 +Tsr*cf*gf*lf*lr
[3.27]
-Tsf*cr*gr*lf*lr -Frx*cf*er*gf*lf^2 +Ffx*cr*ef*gr*lr^2
+Ffx*cf*gf*gr*h*lf +Frx*cf*gf*gr*h*lf -Ffx*cr*gf*gr*h*lr
-Frx*cr*gf*gr*h*lr -Frx*cf*er*gf*lf*lr +Ffx*cr*ef*gr*lf*lr)
/(cf*cr*gf*gr*(lf + lr)^2)
% sol.py = -(Tsr*cf*gf*lf +Tsr*cf*gf*lr +Tsf*cr*gr*lf
+Tsf*cr*gr*lr +Ffx*cf*gf*gr*h +Frx*cf*gf*gr*h
+Ffx*cr*gf*gr*h +Frx*cr*gf*gr*h -Frx*cf*er*gf*lf
-Ffx*cr*ef*gr*lf -Frx*cf*er*gf*lr -Ffx*cr*ef*gr*lr)
/(cf*cr*gf*gr*(lf+lr)^2)
% sol.Ffz = -(Ffx*h +Frx*h -g*lr*m)/(lf+lr)
% sol.Frz = +(Ffx*h +Frx*h +g*lf*m)/(lf+lr)
The solution can be compared with corresponding solution in Equation [3.20]. The is exactly the
same. Then, a general reflection is that the displacement, z and py, in Equation [3.27] follows a com-
plex formula, but that they are dependent on how the 𝑤 𝑓 + is applied: both dependent on
distribution between axles and dependent on how much of the axle forces ( 𝑓 and ) that are actu-
ated with shaft torques (𝑇 𝑓 and 𝑇 , respectively). In Figure 3-33, dashed lines show the solutions
from Equation [3.20].
215
Longitudinal Dynamics
Now, study the suspension at front axle in Figure 3-32. When the axle is braked, 𝑓 will be negative
and push the axle rearwards, i.e. in under the body. The front of the vehicle will then be lifted as in pole
jumping. This means that this design counter-acts the (transient) dive of the front. (Only the transient
dive will be reduced, while the dive after a longer time of kept braking is dependent only on the stiff-
nesses according to Equation [3.20].) The design concept for front axle suspension to place the pivot
point behind axle and above ground is therefore called “anti-dive”.
If the braking is applied without shaft torque 𝑇 𝑓 , a good measure of the Anti-dive mechanism is 𝑒𝑓 / 𝑓 .
This is the normal way for braking, since both the action and reaction torque acts on the axle. For in-
board brakes, or braking via propulsion shaft, the reaction torque is not taken within the axle, but the
reaction torque is taken by the vehicle body. The action torque 𝑇 𝑓 then appears in the equilibrium
equation for the axle, as shown in Equation [3.24]. If we neglect the wheel rotational dynamics for a
while, we can insert 𝑇 𝑓 𝑓 ∙ 𝑅𝑤 in the equation with 𝑇 𝑓 in Equation [3.24]:
( 𝑓+ 𝑓 𝑓 )∙ 𝑓 𝑓 ∙ 𝑒𝑓 + 𝑇 𝑓 0 𝑡ℎ 𝑇 𝑓 𝑓 ∙ 𝑅𝑤 ⇒
⇒( 𝑓+ 𝑓 𝑓 )∙ 𝑓 𝑓 ∙ 𝑒𝑓 + 𝑓 ∙ 𝑅𝑤 0 ⇒ [3.28]
⇒( 𝑓+ 𝑓 𝑓 )∙ 𝑓 𝑓 ∙ (𝑒𝑓 𝑅𝑤 ) 0
We can then see that a good measure of the Anti-dive mechanism is (𝑒𝑓 𝑅𝑤 )/ 𝑓 instead.
216
Longitudinal Dynamics
important, and this is so fast dynamics that the suspension mechanisms of Anti-lift and Anti-dive influ-
ences. The position of the load in the vehicle will influence, since it influences the load transfer.
We will now set up a mathematical model, see Equation [3.29], which shows how the normal forces
change during a braking event. It is based on the physical model in Figure 3-32. Driving resistance con-
tributes normally with a large part of the deceleration, but we will neglect this for simplicity, just to
show how the suspension mechanism works. The equations in the model are presented in the dynamic
modelling standardized format “Modelica”, and are hence more or less identical to Equation [3.22] to
[3.26]. (The term is included but makes no visible difference.)
//Actuation:
Ffx = if 1 < time and time < 3 then -0.4*m*g else 0;
Frx = if 3 < time and time < 7 then -0.4*m*g else 0;
Tsf/Rw = 0;
Tsr/Rw = if 5 < time and time < 7 then -0.4*m*g else 0;
//Motion equations:
der(z) = vz; der(py) = wy;
//Constitutive equations for the springs:
der(Fsf) = -cf*vfz; der(Fsr) = -cr*vrz;
//Constitutive equations for the dampers: [3.29]
Fdf = -df*vfz; Fdr = -dr*vrz;
//(Dynamic) Equilibrium equations:
-m*(der(vx)-vz*wy) + Ffx + Frx = 0;
-m*(der(vz)-vx*wy) - m*g + Ffz + Frz = 0;
-Jy*der(wy) + Frz*lr - Ffz*lf - (Ffx + Frx)*h = 0;
(Frz - Fsr - Fdr)*gr - Frx*er + Tsr = 0;
(Fsf + Fdf - Ffz)*gf - Ffx*ef + Tsf = 0;
//Compatibility:
zf = z - lf*py; zr = z + lr*py;
vfz = vz - lf*wy; vrz = vz + lr*wy;
The simulation results are shown in Figure 3-33. It shows a constant deceleration, but it is changed
how the decelerating force is generated. At time=3 s, there is a shift from braking solely on front axle
to solely on rear axle. The braking is, so far, only done with friction brakes, i.e. generating torque by
taking reaction torque in the axle itself. At time=5 s, there is a shift from braking with friction brakes to
braking with shaft torque. It should be noted that if we shift axle or shift way to take reaction torque,
gives transients even if the deceleration remains constant.
One can also see, at time=1 s, that the normal load under the braked axle first changes in a step. This is
the effect of the Anti-dive geometry. Similar happens when braking at rear axle, due to the Anti-squat
geometry. Since brake performance is much about controlling the pressure rapidly, the transients are
relevant, and the plots should make it credible that it is a control challenge to reach a high braking effi-
ciency.
217
Longitudinal Dynamics
w ithCentrifugalForce.vx
50
0
0 1 2 3 4 5 6 7 8 9
w ithCentrifugalForce.z [m] w ithCentrifugalForce.py [rad] w ithCentrifugalForce.z
0.03
0.01
py= /[𝑟 ]
0.00 /[ ]
/[ ]
-0.01
𝑓 /[ ]
-0.02
0 1 2 3 4 5 6 7 8 9
w ithCentrifugalForce.Fzf w ithCentrifugalForce.Fzr w ithoutCentrifugalForce.Fzf
1.4E4
1.2E4 𝑓
1.0E4
8.0E3
Same steady state vertical
6.0E3
forces, but transiently different
4.0E3
0 1 2 3 4 5 6 7 8 9
Braking on front axle Braking on rear axle Braking on rear axle
with friction brakes with friction brakes with shaft torque
Figure 3-33: Deceleration with constant vehicle deceleration but varying ways of actuation. (Eq
[3.29]. With the centripetal term (solid) and without (dashed). Dotted shows without anti-
dive/-squat geometry, i.e. 𝑓 . The term makes no visible difference.)
218
Longitudinal Dynamics
219
Longitudinal Dynamics
A development of CC has varying set speed trajectory ahead. The trajectory is typically optimized for a
predicted operation, about a minute ahead. The optimization typically considers road topography. In-
termittent propulsion and traffic are other phenomena that can be considered in optimization. Such
products are on the market, e.g. Volvo iSee and Scania Active Prediction for heavy vehicles.
§ Concepts of Predictive Optimization for Predictive CC
A predictive CC works in principle as figure below. It is an automated longitudinal driving function. It
selects an optimal trajectory for a certain horizon ahead, e.g. 1 km ahead and minimizing transport
cost including energy cost. The figure assumes it defines candidates to select by trying different (𝑠),
but there are also many other ways to define candidates, such as different driver model parameters.
The figure indicates an optimization by trial-and-error, but more sophisticated methods can be used,
such as Dynamic Programming, Quadratic Programming, Linear Programming, MPC, etc. Other quanti-
ties, such as gear shifting, including Neutral gear selection, 𝑒 𝑟(𝑠) and state of charge 𝑆𝑜 (𝑠) if the
vehicle is a hybrid vehicle, can be treated in primarily two different ways. They can be part of candi-
date definition, which increases the search space for the optimization. Alternatively, they can be calcu-
lated (e.g. to give momentaneous optimum) in the model used for optimization, which makes them
available as additional optimum trajectories beside 𝑂𝑝 (𝑠).
Macro-level information about traffic density and traffic mean velocity or the route ahead can be con-
sidered in optimization. But taking micro-level traffic information, such as nearest vehicle ahead, is
more challenging, especially predicting the velocity of the object vehicles.
220
Longitudinal Dynamics
RouteManager (executes seldom, e.g. once per 𝑠, and predicts over a horizon, e.g. 𝑘 ) Vehicle
Model (simulated to calculate 𝑜𝑠𝑡) Vehicle & Actuators Model
𝑠
Environment “Ideal following
Model (map with model” Actuators
route and/or
environment sensors) 𝑛 𝑠
Figure 3-34: § Environment, Driver and Vehicle, where the vehicle has a Predictive CC. See also
Figure 3-42.
§ Optimization by Varying Acceleration Trajectory
An example of how the cost (as sum of energy cost and fuel cost) can be reduced is given in the 3 fol-
lowing figures.
Figure 3-35: § Simulation with full model, including a driver model and actuator delays.
221
Longitudinal Dynamics
Figure 3-36: § Simulation with the model used for optimization with varying 𝑛 (𝑠). The
acceleration trajectory from previous figure used, discretized every 25 m.
Figure 3-37: § Simulation with the model used for optimization with varying 𝑛 (𝑠). The
acceleration trajectory from optimization used, discretized every 25 m. Cost reductions of ≈40%
and ≈10%, respectively, is found with the parameters used in this example. The candidates
are marked to envision that a next step optimization could be done with a model including
actuator dynamics.
The optimizations are made with Matlab function fmincon, with (𝑠) (discretized to parameters
[ (𝑠 ) ⋯ (𝑠𝑁 )]) as parameters to vary. The cost function included a simulation of the system, includ-
ing calculation of cost and constraints. Note that the simulations in the cost functions can include “in-
verse dynamics” if the candidate prescribe the whole trajectory of a state, which can enable more sta-
ble derivative approximations.
222
Longitudinal Dynamics
Note also that the above figure with optimisation result is not using the full model, i.e. delays in VMC2
and actuators are not considered. So, the end result in cost reduction will not be seen until running it
in a more accurate vehicle model or a real vehicle.
§ Optimization by Varying Force Request Trajectory
If we instead use the (𝑠) as theway to define the candidate, we can still pick an initial guess from
same simulation as before. But the
Figure 3-38: § Simulation with the model used for optimization with varying 𝑛 (𝑠),
discretized every 25 m.
223
Longitudinal Dynamics
Figure 3-39: ABS control. Principle and control sequence, from Ref (Gillespie, 1992)
-0.05
-0.1
sxfl/[1]
-0.15 sxfr/[1]
sxrl/[1]
-0.2 sxrr/[1]
-0.25
120
100
80
60
pfl/[bar]
40 pfr/[bar]
prl/[bar]
20 prr/[bar]
vx/[km/h]
0
1 1.5 2 2.5 3 3.5 4 4.5
time [s] 5
Figure 3-40: ABS control, Data log from passenger car test.
224
Longitudinal Dynamics
There are other side functions enabled by having ABS on-board. Such are “select low”, which means
that the brake pressure to both wheels on an axel is limited by the one with lowest pressure allowed
from ABS. So, if one wheel comes into ABS control, the other gets the same pressure. This is most rele-
vant on rear axle (to reduce risk of losing side grip) but one tries to eliminate the need of it totally, be-
cause it reduces the brake efficiency when braking in curve or on different road friction left/right.
It is often difficult to define strict border between functions that is a part of ABS and which is part of
EBD, which is why sometimes one say ABS/EBD as a combined function.
225
Longitudinal Dynamics
Figure 3-41 shows a diagram where different condition areas are marked. The sectioned area shows
where AEB will be triggered, using above rules. The smaller of the sectioned areas shows where it also
will be possible to trigger AEB so timely that a collision is actually avoided; with the assumed num-
bers, this is for speeds up to 6.4 m/s≈23 km/h.
𝑥
(or lead vehicle)
𝑇𝑇
object vehicle
𝑥 0.4 𝑠
Collision if normal
lateral avoidance
𝑦 𝑚
( 6 2)
𝑦 𝑙
. /𝑠 6.4 /𝑠
AEB triggered AND avoids collision
(or own vehicle)
⁄ ⁄
subject vehicle
226
Longitudinal Dynamics
many pieces of information and simple models; vehicle dynamics and driver behaviour (in both sub-
ject and object vehicles) as well as road characteristics. AEB function has to be designed together with
other similar functions, such as ACC and Forward Collision Warning (FCW).
AEB is legal requirement for both passenger vehicles and heavy vehicles (ISO19377, 2017).
Related functions are, e.g., an automatic extra force assistance in brake pedal when driver steps
quickly onto brake. Another related function is automatic braking triggered by a first impact and in-
tended to mitigate or avoid secondary accident events, starts to appear at market, see Reference (Yang,
2013). In semantic meaning, this could be seen as AEB, but they are normally not referred to as AEB;
AEB normally refers to functions that use environment sensors (forward directed radar, camera, etc.).
When designing and evaluating AEB, it is important to also know about the function Forward Collision
Warning, FCW. FCW is a function that warns the driver via visual and/or audio signals when a forward
collision is predicted. FCW is typically triggered earlier than AEB.
AEB as described above could be called “Rear-end AEB”. Another similar functionality, not included in
today’s AEB, could be called “Intersection AEB” and include braking for intersecting traffic, see
(Sander, 2018).
227
Longitudinal Dynamics
Maps
… Traffic Situation Layer Forrward Direction
Arb(Min)
Figure 3-42: Functional architecture for conventional front axle driven passenger car. Mainly
longitudinal functions (plus ESC, RSC) are shown, e.g. no steering. Cf. Figure 1-56.
If a reference architecture is used, it can assist function developers from OEM’s Electrical, Powertrain,
and Chassis departments and suppliers to have a common view of how vehicle’s embedded motion
functionality is intended to be partitioned and to understand how different functions relate and inter-
act with each other and what responsibilities they have.
228
Lateral Dynamics
4 LATERAL DYNAMICS
4.1 Introduction
The lateral motion of a vehicle is needed to follow the road curves, select route in intersections and
laterally avoid obstacles, which all involve steering. Vehicle steering is studied mainly through the ve-
hicle degrees of freedom: yaw rotation and lateral translation .
A vehicle can be steered in different ways:
• Applying steer angles on road wheels. Normally both of front wheels are steered with approxi-
mately same angle. Steering system described in 0.
• Applying longitudinal forces on road wheels; directly by unsymmetrical between left and right
side of vehicle, e.g. one-sided braking, or indirectly by deliberately use up much friction longi-
tudinally on one axle in a curve, so that that axle loses lateral force.
• Articulated steering, where the axles are fixed mounted on the vehicle body, but the vehicle
itself can bend.
The turning manoeuvres of vehicles encompass two sub-attributes. Handling is the driver’s perception
of the vehicle’s response to the steering input. Cornering is the physical response (open loop) of the
vehicle independent of how it influences the driver.
The lateral dynamics of vehicles is often experienced as the most challenging for the new automotive
engineer. Longitudinal dynamics is essentially motion in one plane and rectilinear. Vertical dynamics
may be 3 dimensional, but normally the displacements are small and in this compendium the vertical
dynamics is mainly studied in one plane as rectilinear. However, lateral dynamics involves motion in
the vehicle coordinate system which introduces curvilinear motion since the coordinate system is ro-
tating as the vehicle yaws.
The chapter introduces the models, more or less, in order of increasing number of states. For some
readers, it might be comprehensive to start with a look-ahead on the “linear one-track model” in 4.4.2.
This is the simplest model which yet captures the essential lateral motion quantities and as
states. The models earlier in the chapter can then be seen as simplifications.
229
Lateral Dynamics
230
Lateral Dynamics
Turning diameter, but also for any other certain manoeuvre. It can be defined for kerb and wall; dis-
tance between wheels or body points. The SPW should be small for good manoeuvrability.
x[m]
Figure 4-1: Paths for wheels and body points (added to result in Figure 4-5).
231
Lateral Dynamics
in the manoeuvre. The states in the simulation are the path coordinates with orientation (𝑥 𝑦 ) and
articulation angles ( 2 …).
The steady state is typically approached asymptotically, so the corresponding circle values requires
either long simulations or inserting state derivatives zero and algebraic solution. From geometry in
Figure 4-2 on can find an expression for (Circle) Off-tracking Δ:
(𝑅𝑓 𝐿) 𝑅𝑓 𝑅 𝑅𝑓 √𝑅𝑓2 𝐿2
[4.2]
( 𝑓 𝐿) 𝑅𝑓 𝑅 𝐿⁄sin( 𝑓 ) 𝐿⁄tan( 𝑓 )
𝑓 𝑓
𝑝 𝑝 𝑓 𝑙𝑙 𝑓 𝑤 𝑓 𝑓
𝑓 𝑤 𝑓 𝑙𝑙
Figure 4-3: Smaller turning circle diameter for front axle propulsion, as compared to rear axle
propulsion due to rolling resistance on the un-driven axle.
232
Lateral Dynamics
4.2.3.2 Model
The model in Eq [4.1] predicts a motion without involving forces, or actually assuming forces are zero.
To get a more complete model, where more variables can be extracted, we can set up the model in Fig-
ure 4-4.
Velocities: 𝑓 arctan 𝑠 𝑤
𝑓 arctan
𝑓 𝑣 Forces:
𝑓 𝑣 𝑓 𝑣
𝑓 𝑓
𝑓 𝑣
0
𝑓 𝑣
𝑟𝑦
𝑓 𝑓
𝑓 𝑣
𝑓 𝑓
𝑓 𝑓
𝐿 𝐿
Figure 4-4: One-track model with ideally tracking axles. Lower view of front wheel shows
conversion between wheel and vehicle coordinate systems.
The “physical model” in Figure 4-4 gives the following “mathematical model”:
Equilibrium (longitudinal, lateral and yaw around CoG):
0 𝑓 𝑣 +
0 𝑓 𝑣 +
0 𝑓 𝑣 ∙ 𝑓 ∙
Transformation between vehicle and wheel coordinate systems:
𝑓 𝑣 𝑓 𝑤 ∙ cos( 𝑓 ) 𝑓 𝑤 ∙ sin( 𝑓 )
𝑓 𝑣 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓 )
𝑓 𝑣 𝑓 𝑤 ∙ cos( 𝑓 ) 𝑓 𝑤 ∙ sin( 𝑓 )
𝑓 𝑣 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓 )
Compatibility between CoG and axles:
𝑓 𝑣 𝑛 𝑓 𝑣 + 𝑓∙
𝑛 ∙ [4.3]
Ideal tracking (Constitutive relation, but with infinite lateral slip stiffness):
𝑓 𝑤 0 𝑛 0
Path with orientation (compatibility), from Eq [1.5]:
𝑥 ∙ cos( ) ∙ sin( )
𝑦 ∙ cos ( ) + ∙ sin( )
233
Lateral Dynamics
actual assumption about ideal tracking lies in that 𝑓 𝑤 0. Global coordinates from Figure
1-49 is also used. A driving resistance of 100 N is assumed on the rear axle ( 00 ) to exemplify
that forces do not need to be zero.
//Equilibrium:
0 = Ffxv + Frx;
0 = Ffyv + Fry;
0 = Ffyv*lf - Fry*lr;
//Ideal tracking (Constitutive relation, but without connection to forces):
vfyw = 0; vry = 0;
//Compatibility:
vfxv = vx; vfyv = vy + lf*wz;
vrx = vx; vry = vy - lr*wz;
//Transformation between vehicle and wheel coordinate systems:
Ffxv = Ffxw*cos(df) - Ffyw*sin(df);
Ffyv = Ffxw*sin(df) + Ffyw*cos(df); [4.4]
vfxv = vfxw*cos(df) - vfyw*sin(df);
vfyv = vfxw*sin(df) + vfyw*cos(df); //or atan(vfyv/abs(vfxv))=df+atan(sfy); sfy=0;
//Path with orientation:
der(x) = vx*cos(pz) - vy*sin(pz);
der(y) = vy*cos(pz) + vx*sin(pz);
der(pz) = wz;
// Prescription of actuation:
df = if time < 4.5 then (35*pi/180)*sin(0.5*2*pi*time) else 35*pi/180;
//Rear axle undriven, which gives drag from roll resistance:
Frx = -100;
4.2.3.3 Simulation
The longitudinal speed is a parameter, 0 𝑘 /ℎ. A simulation result from the model is shown in
Figure 4-5. It shows the assumed steer angle function of time, which is an input. It also shows the re-
sulting path, 𝑦(𝑥).
df [rad] y [m]
35 deg
x[m]
time [s]
Figure 4-5: Simulation results of one-track model with ideal tracking tyres.
The variables 𝑥 𝑦 𝑝 are the only state variables of this simulation. If not including the path
model (Eq [1.5]), the model would be only an algebraic system of equations. That system of equations
could be solved isolated for any value of steer angle without knowledge of time history.
234
Lateral Dynamics
Examples are a two-axle vehicle which has parallel steering and truck with 3 axles, whereof the two
rear are non-steered, respectively.
Ffyv
vx df Fr2x Fr1x df
vr2x vr1x
𝑓
lr lf 𝑓
L
Δ𝐿𝑟 Δ𝐿𝑟
2 2 DeltaL
Figure 4-6: Non-Ackermann geometry, due to un-steered rear axles. Top: Rigid Truck with 3 axles,
whereof only the first is steered. Bottom: One-track model.
The changes we have to do in the model appear as underlined in Equation [4.5]. There has to be double
variables for , denoted 1 and 2 respectively. Also, we cannot (mathematically) use
𝑓 𝑤 0 anymore, but instead we introduce a lateral tyre force model, as described in 2.2.4.
//Equilibrium:
0 = Ffxv + Fr1x + Fr2x; //grade resistance could be added here, e.g. “+500”
0 = Ffyv + Fr1y + Fr2y;
0 = Ffyv*lf - Fr1y*(lr - DLr/2) - Fr2y*(lr + DLr/2);
//Constitutive relation (with slip, as opposed to Ideal tracking):
Ffyw = -Cf*sfy; sfy = vfyw/abs(vfxw);
Fr1y = -Cr1*sr1y; sr1y = vr1y/abs(vr1x);
Fr2y = -Cr2*sr2y; sr2y = vr2y/abs(vr2x);
//Compatibility:
vfxv = vx;
vfyv = vy + lf*wz;
vr1x = vx;
vr2x = vx;
vr1y = vy - (lr – DLr/2)*wz;
vr2y = vy - (lr + DLr/2)*wz; [4.5]
//Transformation between vehicle and wheel coordinate systems:
Ffxv = Ffxw*cos(df) - Ffyw*sin(df);
Ffyv = Ffxw*sin(df) + Ffyw*cos(df);
vfxv = vfxw*cos(df) - vfyw*sin(df);
vfyv = vfxw*sin(df) + vfyw*cos(df);
//Path with orientation:
der(x) = vx*cos(pz) - vy*sin(pz);
der(y) = vy*cos(pz) + vx*sin(pz);
der(pz) = wz;
// Prescription of steer angle:
df = if time < 4.5 then (35*pi/180)*sin(0.5*2*pi*time) else 35*pi/180;
//Rear axles undriven, which gives drag from rolling resistance:
Fr1x = -100/2*sign(vx);
Fr2x = -100/2*sign(vx);
235
Lateral Dynamics
The new result is shown in Figure 4-7, which should be compared to Figure 4-5. The radius in the cir-
cle increases a little, which is intuitive, since the double rear axle makes turning less easy.
df [deg] y [m]
35 deg
vx=2.778m/s
vy=1.036m/s
wz=0.652rad/s
𝐿𝑟
2
236
Lateral Dynamics
3: Instantaneous
centre of rotation for
semi-trailer (1 axle and
1 coupling points)
237
Lateral Dynamics
4.2.6 Reversing
Low speed manoeuvring is often about both driving forward and reversing. The derived models work
formally also for < 0. Assume that, after some driving forward, is changed to a negative value.
Assume that steering is same for same position along the path, (𝑠). The vehicle will then reverse
in approximately same path as it first drove forward. If ideal tracking tyres, such as Eq [4.1], the paths
will be identical. If adding forces to the model, the reverse path can deviate from forward path. A small
such deviation can be seen in figure due to a large rolling resistance coefficient (𝑅𝑅 0. 0).
In reality, the reverse path for an articulated vehicle often differs more from the forward path. This is
mainly due to back-lash in couplings; even if only some centimetre backlash it can influence a lot. The
backlash can be modelled by letting coupling point position be dependent of sign(coupling force).
𝑓
𝑓
𝐿 𝑓
𝐿
Figure 4-12: A low speed model of tractor with semi-trailer. Small angles 𝑓 are assumed.
The mathematical model for lateral position variables (𝑦 ) becomes:
𝑙
𝑦 𝐿 𝑓+ ∙ (from Eq [1.5], with small )
1
𝑣1 𝛿
𝐿1
(rotational velocity = tangential speed / radius to instantaneous centre)
238
Lateral Dynamics
𝑣1 𝛿 𝑣2𝑦
𝐿1 𝐿2
(articulation angle is difference between units)
𝑙
2 +𝐿 (units have same velocity in coupling point, small , small 𝑓)
1
Eliminating 2 gives:
𝑦 0 0 𝑦 ⁄𝐿
[ ] [0 0 0 ] [ ]+ [ ⁄𝐿 ] 𝑓
⏟0 0 ⁄𝐿2 ⁄𝐿 ⁄(𝐿 𝐿2 )
𝑨
The eigenvalues to 𝑨 becomes 𝜆 2 ±0 𝜆3 ⁄𝐿2 . So, the system is unstable when < 0, be-
cause it makes e(𝜆3 ) > 0. It is probably intuitive for may readers that the vehicle is unstable, as the
semi-trailer pushed rearwards via a moment-free joint. However, the analysis was included in the
compendium to show how stability appears for low speed models, i.e. without inertial terms 𝑠𝑠
𝑒 𝑒𝑟 𝑡 𝑜𝑛. The model can be extended with other multiple-unit vehicles, multiple axles in axle
groups and driver models. The state variables will remain as 𝑦 and one articulation angle for
each coupling.
239
Lateral Dynamics
Vehicle Handling
Dynamics track
Area
Skid pad
High speed
track
Figure 4-14: An example of test track and some parts with special relevance to Vehicle Dynamics.
The example is Hällered Proving Ground, Volvo Car Corporation.
Figure 4-15: An example of test track. The example is AstaZero (Active Safety Test Arena), owned
by Research Institute of Sweden (RISE) and Chalmers University of Technology.
240
Lateral Dynamics
High
speed
circle
Skid pad
Low
friction
Vehicle Dynamics Area
strips
Hill
strips
Figure 4-16: An example of test track. The example is CASTER’s (virtual) test track. Used for
CASTER’s driving simulator at Chalmers University of Technology.
241
Lateral Dynamics
𝑓 𝑣
𝑓 𝑣
+ cos( ) centre
𝑓
vrx 𝑓 𝑣 𝑓 𝑣
𝐿
𝑓 𝑓 𝑓
cos
sin
b
Figure 4-17: One-track model for Steady State Cornering. Dashed forces are “fictive forces”.
The model in Figure 4-17 has the mathematical form in Eq [4.6] (in Modelica format). Longitudinal
speed is assumed to be positive. The subscripts v and w refer to vehicle coordinate system and
wheel coordinate system, respectively. A driving resistance of 100 N is assumed on the rear axle
(Frx=100;). The longitudinal speed is a parameter, 00 𝑘 /ℎ.
//Equilibrium:
m*ax = Ffxv + Frx; //Air and grade resistance neglected
m*ay = Ffyv + Fry; Jz*0 = Ffyv*lf - Fry*lr; // der(wz)=0
-ax = wz*vy; +ay = wz*vx;
//Constitutive relation, i.e. Lateral tyre force model:
Ffyw = -Cf*sfy; sfy = vfyw/vfxw;
Fry = -Cr*sry; sry = vry/vrx;
//Compatibility:
vfxv = vx; vfyv = vy + lf*wz;
vrx = vx; vry = vy - lr*wz;
//Transformation between vehicle and wheel coordinate systems:
Ffxv = Ffxw*cos(df) - Ffyw*sin(df); [4.6]
Ffyv = Ffxw*sin(df) + Ffyw*cos(df);
vfxv = vfxw*cos(df) - vfyw*sin(df);
vfyv = vfxw*sin(df) + vfyw*cos(df);
//Path with orientation (from Eq [1.5]):
der(x) = vx*cos(pz) - vy*sin(pz);
der(y) = vy*cos(pz) + vx*sin(pz);
der(pz) = wz;
// Prescription of steer angle:
df = if time < 2.5 then (5*pi/180)*sin(0.5*2*pi*time) else 5*pi/180;
// Rear axle undriven, which gives drag from roll resistance:
Frx = -100;
A simulation result from the model is shown in Figure 4-18. Note that steering start to the left, but ve-
hicle path starts bending to the right. This comes from that it is a steady state model but used in a tran-
sient manoeuvre. The steady state cornering condition is found directly and turning left has the steady
state directed outwards, to the right, due to centrufugal force.
242
Lateral Dynamics
df [rad] y [m]
5 deg 2 +
2
=
8. 7 /𝑠
Ffyw
Ffxw
Frx-100 Nm
wz = 0.3995 rad/s
vy = -5.761 m/s
x[m]
time [s]
Figure 4-18: Simulation results of steady state one-track model. The vehicle sketched in the path
plot is not in scale, but correctly oriented.
Now, the validity of a model always has to be questioned. There are many modelling assumptions
which could be checked, but in the following we only check the assumption instead of the
more correct + , which we will learn in “4.4.2 One-Track Model”. Comparison of the
terms gives | |
𝑚
≈ | | 𝑚 ≈ 0 ⁄𝑠 2 , so | | is large and this jeopardizes the validity.
Large | | happens during 0 < 𝑡 <≈ 𝑠, so the model is not very valid there. But, at 𝑡 >≈ 𝑠, the
model is valid, at least in this aspect, since there | | ≈ 0 ≪ | |. So, the model is not so valid dur-
ing the initial sinusoidal steering. This shows that a steady state models should not be trusted outside
steady state conditions.
Eq [4.6] is a complete model suitable for simulation, but it does not facilitate understanding very well.
We will reformulate it assuming small 𝑓 (i.e. cos( 𝑓 ) , sin( 𝑓 ) 0, and 𝑓2 0). Eliminate slip, all
forces that are not wheel longitudinal, and all velocities that are not CoG velocities:
( +( + 𝑓 ) 𝑓)
𝑓 𝑤 ( +( + 𝑓 ) 𝑓) + 𝑓 ( + 𝑓 ) 𝑓 + ( +( + 𝑓 ) 𝑓)
( +( + 𝑓 ) 𝑓)
𝑣𝑦 𝑙𝑟 𝜔𝑧
𝑓 ( 𝑓 +( + 𝑓 )) + 𝑓 𝑤 𝑓 ( +( + 𝑓 ) 𝑓) [4.7]
𝑣
𝑓 ( 𝑓 +( + 𝑓 )) 𝑓 𝑓 𝑤 𝑓 𝑓
( +( + 𝑓 ) 𝑓)
Eq [4.7] is a complete model, which we can see as a dynamic system without state variables.
• Actuation: Steering and wheel torque on each axle: 𝑓 𝑓 𝑤 .
• Motion quantities:
For propulsion on both axles, it can be a reasonable case that we know 𝑓 𝑓 𝑤 and want to calcu-
late . Eq [4.8] is a rearrangement of Eq [4.7] for this purpose. It can be used for rear axle
243
Lateral Dynamics
drive using 𝑓 𝑤 0 and calculating how large needs to be. Front axle drive requires more rear-
rangements.
𝑓 𝐿+ 𝐿 𝑓 𝑤
2 2 𝑓
𝑓 𝐿 + ( 𝑓 𝑓)
( 2 ) + ( 2)
𝑓 𝐿 𝑓 𝐿 𝑓 𝑓 𝑤
𝐿2 + ( 2 𝑓 [4.8]
𝑓 𝑓 𝑓)
𝑓 ∙( + 𝑓 ∙ )∙ 𝑓
∙ ∙ 𝑓 𝑤
(( + 𝑓 ∙ )∙ 𝑓 + )
-0.2
-0.4
time [s]
-0.6
0 1 2 3 4
𝑓 ∙ ∙ 𝐿2 + ( ∙ 𝑓 ∙ 𝑓) ∙ ∙ 2
𝑠𝑠𝑢 𝑒 𝑠 ⇒
𝑓 ∙ ≈{ }≈
𝑓 ∙ ∙𝐿+( ∙ 𝑓 + ∙ )∙ 𝑓 𝑤
⇒ ⁄ ≈ 𝑅𝑝
∙ ∙ 2 𝑠𝑠𝑢 𝑒:
𝐿 𝑓 𝑓 ∙
≈ ∙ + ∙ ≈{ }≈
+ 𝑓 𝑤⁄ 𝑓 𝑅𝑝 ( 𝑅𝑝 𝑤⁄ 𝑓 ≈ 0
+ 𝑓 𝑤) ∙ ∙𝐿 𝑓
𝑓
2
[4.9]
𝐿 ∙ 𝑓 ∙ 𝑓 ∙ ∙ 𝑓 ∙ 𝑓 𝑓
≈ + ∙ {𝑢𝑠𝑒: 𝐾 }
𝑅𝑝 𝑓 ∙ ∙𝐿 𝑅𝑝 𝑓 ∙ ∙𝐿 𝑓 ∙𝐿 ∙𝐿
2
𝐿 ∙
+𝐾 ∙
𝑅𝑝 𝑅𝑝
The coefficient 𝐾 is the understeer gradient and it will be explained more in 4.3.3.
𝑓 ∙ ∙ 𝐿2 + ( ∙ 𝑓 ∙ 𝑓) ∙ ∙ 2
𝑓 2)
∙ ≈
𝑓 ∙( ∙𝐿 𝑓 ∙ ∙ 2) + ( 𝐿 𝑓 ∙ 𝑓 𝑤
𝑓 ∙ ∙ 𝐿2 ( 𝑓 ∙ 𝑓 ∙ )∙ ∙ 2
≈ {𝑢𝑠𝑒: 𝑓 𝑤 0} ≈ 2
∙ ≈
∙ ∙ ∙𝐿 ∙ 𝑓∙ ∙
𝑓 𝑓 [4.10]
𝐿
𝑓 → ∙
𝑣 →
⇒ ∙ ∙ 𝐿
𝑓 𝑓
𝑓 → ∙ 𝐾 ∙ ∙ ∙
{ 𝑣 →∞ 𝑓 ∙ 𝑓 𝑓
244
Lateral Dynamics
𝑣𝑦
We can see that there is a speed dependent relation between steer angle and side slip, . The side slip
𝑣
𝑦 𝑣
can also be expressed as a side slip angle, arctan ( 𝑣 ). Since normally 𝐾 > 0, the side slip changes
sign, when increasing speed from zero to sufficient high enough. This should feel intuitively correct, if
agreeing on the conceptually different side slip angles at low and high speed, as shown in Figure 4-20.
We will come back to this equation in context of Figure 4-31.
v v
vehicle
vehicle path
path centre
centre
path of
front axle
path of
rear axle
Figure 4-20: Body Slip Angle for Low and High Speed Steady State Curves
245
Lateral Dynamics
Note: We can still identify an 𝐾 , but the reference angle 𝐴 , see also 4.3.3, is not as simple as 𝐿⁄𝑅𝑝 .
𝐾 3 is the definition used in (ISO 8855). For 𝐾 3 , one can sometimes see the unit “rad/g” used, which
present compendium recommended to not use.
If vertical loads on axles are only due to gravity ( (𝐿 )⁄𝐿) and tyres linear with vertical
load ( ) we can express 𝐾 2 ⁄ 𝑓 ⁄ .
246
Lateral Dynamics
Figure 4-22: Under- and over-steering for a two-axle vehicle with 𝑓 . It does not depend on
which axle is steered, but which axle is first in the direction of motion. Figure drawn for vehicle
driving forward.
247
Lateral Dynamics
𝐿 𝐿∙ 𝐿∙
𝑛 𝑙 ≈{ ≈𝑅∙ }≈ ≈{ ≈ ∙ }≈ 2 [4.15]
𝑅
If the actual vehicle has | 𝑓 | < | 𝑛 𝑙 | the vehicle oversteers, and vice versa. This is often very prac-
tical since it only requires simply logged data, 𝑓 and . Note that when 𝑓 and 𝑛 𝑙 have differ-
ent signs, neither understeer nor oversteer is suitable as classification, but it can sometimes be called
“counter-steer”. An example of applying Eq [4.15] is shown in Figure 4-23, where one also see that the
ESC system does not follow the Eq [4.15] when deciding ESC interventions; ESC has more advanced
“reference models”, see 4.6.2.1.
A second look at Equation [4.9] tells us that we have to assume absence of propulsion and braking on
front axle, 𝑓 𝑤 0, to get the relatively simple final expression. When propulsion on front axle
( 𝑓 𝑤 > 0), the required steer angle, 𝑓 , will be smaller; the front propulsion pulls in the front end of
the vehicle. When braking on front axle ( 𝑓 𝑤 < 0), the required steer angle, 𝑓 , will be larger; the
front braking hinders the front end to turn in. To keep constant, which is required within definition
of steady state, one have to propel the vehicle because there will always be some driving resistance to
overcome. Driving fast on a small radius is a situation where the driving resistance from tyre lateral
forces becomes significant, which is a part of driving resistance which was only briefly mentioned in
3.2.
Figure 4-23: Log data from passenger car with ESC in a double lane change. Upper: Vehicle
motion. Middle: 𝑛 𝑙 from Eq [4.15] used to define “instantaneous under-/over-steering”
(US/OS). Lower: Pressure to each wheel brake.
248
Lateral Dynamics
We can see that the understeer gradient from steady state cornering model appears also in the for-
mula for neutral steering point position, . Since 𝑓 𝑛 𝐿 are positive, the neutral steering point is
behind of CoG for understeered (two-axle) vehicles, and in front of CoG for oversteered (two-axle) ve-
hicles. This is why and 𝐾 can be said to be alternative measures for the same vehicle function/char-
acter, the yaw balance.
Physical model: Mathematical model:
• Steady state ( 0) Equilibrium:
• Straight ahead driving (
𝑓 +
0) 0 +
• No steering
• Small tyre and vehicle side slip. Then,
0 𝑓 𝑓
angle=sin(angle)=tan(angle). Constitution: 𝑓 𝑓 𝑠𝑓 and 𝑠
𝑣𝑦
Fey Compatibility: 𝑠𝑓 𝑠
Fry Ffy 𝑣
𝑟 𝑙𝑟 𝑓 𝑙𝑓
Eliminate 𝑓 , 𝑠𝑓 𝑠 yields:
b 𝑟 𝑙𝑟
𝑓+ 𝑟
𝑓 𝑙𝑓
Identify understeering gradient, 𝐾
𝑓 𝑟 𝐿
0 Then:
𝑓 𝑓
L 𝐾 𝐿
𝑓 +
Figure 4-24: Model for definition and calculation of neutral steering point.
2.5
Ku = 2.525e-6 [1/N]
Understeered
2
required steering, df*R/L [rad]
1.5
Neutral steered
Ku = 0e-6 [1/N]
1
Oversteered
0.5 Ku = -1.794e-6 [1/N]
Figure 4-25: Normalized steer angle ( 𝑓 ∙ 𝑅 ⁄𝐿) for Steady State Cornering
249
Lateral Dynamics
𝑓
( ) { } 0
ℎ ℎ 0
( ∙ ) ( + ∙ )
𝑓
250
Lateral Dynamics
However, the cornering stiffness varies degressively, e.g. 𝑘𝑝∙ 𝑘 ∙ 2 . This is further stud-
ied in Reference (Drenth, 1993).
If taking the degressiveness of tyre cornering stiffness into account, the weight distribution plays a
role also without longitudinal load transfer; front biased weight distribution gives under-steered vehi-
cles and vice versa. Also, the number of wheels per axle influence stronger; single wheel front (or dou-
ble-mounted rear) gives under-steered vehicles and vice versa.
It should be noted that if the longitudinal acceleration is due to wheel torques, as opposed to road
grade or aerodynamic forces, the tyre combined slip effects will influence the curves which is not con-
sidered in Figure 4-26; the cornering stiffness of an axle will decrease with increased longitudinal
force.
𝑓⁄ 𝑓 𝑓 ⁄ =
Figure 4-26: Left: Under-steering gradient as function of longitudinal acceleration, , and static
load distribution, 𝑓 ⁄𝐿. Right: Critical and characteristic velocity as function of acceleration and
load distribution.
4.3.5.1.1 § Influence of Longitudinal Tyre Forces
With known combined slip effect, we could distribute longitudinal force so that 𝐾 is unchanged
and/or we could avoid braking so that becomes lower than the actual speed . To examplify
this, we use the simple combined slip model in Eq [2.46] and require 𝐾 0. We also assume flat road
and no air resistance. Resulting distribution is compared with two other distribution rules in next fig-
ure. For low | |, we have to apply wheel torque in different directions to keep 𝐾 0.
251
Lateral Dynamics
⁄
⁄
⁄
⁄
⁄
𝑓
⁄
𝑓
⁄𝑓
⁄
𝑓
Figure 4-27: Three wheel torque distribution strategies: Distribution as static distribution of
vertical forces, Equal utilization of (as “ideal curve” in 3.4.4) and Neutral steered.
252
Lateral Dynamics
30
Oversteered
25
/
Ku = -1.794e-6 [1/N]
wz/df [(rad/s)/rad]
20
⁄ 𝑓
Neutral steered
Ku = 0e-6 [1/N]
Characteristic
gain,
15
speed (for
yaw rate gain,
understeered
yaw velocity
vehicle)
10
Understeered
Ku = 2.525e-6 [1/N]
5
Critical speed (for
oversteered vehicle)
0
0 5 10 15 20 25 30 35 40 45 50
longitudinal vx
velocity,
[m/s] ⁄𝑠
Figure 4-28: Yaw velocity gain ( / 𝑓 ) for Steady State Cornering. Each “cluster of 3 curves”:
Mid curve 𝑓 𝑤 0. Upper 𝑓 𝑤 +0.5 ∙ 𝑓 . Lower 𝑓 𝑤 0.5 ∙ 𝑓 .
Oversteered
Ku = -1.794e-6 [1/N]
2.5
⁄ [1/(m*rad)]
𝑚
2
curvature gain, (1/R)/df 𝑓
1.5
curvature gain,
Neutral steered
0.5 Ku = 0e-6 [1/N]
253
Lateral Dynamics
500
Oversteered
𝑚⁄ 2
Ku = 0e-6 [1/N]
400
⁄ 𝑓
350
300
lateral acceleration gain,
250
Understeered
200
Ku = 2.525e-6 [1/N]
150
From the previous figures the responsiveness of the vehicle can be identified for different understeer
gradients. In all cases the vehicle which is understeered is the least responsive of the conditions. Both
the yaw velocity and lateral acceleration cannot achieve the levels of the neutral steered or over-
steered vehicle. The over-steered vehicle is seen to exhibit instability when the critical speed is
reached since small changes in the input result in excessive output conditions. In addition, the over-
steered vehicle will have a counter-intuitive response for the driver. To maintain a constant radius
curve, an increase in speed requires that the driver turns the steering wheel opposite to the direction
of desired path. The result of these characteristics leads car manufacturers to produce understeered
vehicles that are close to neutral steering to achieve the best stability and driver feedback.
It is not solely the understeering gradient that sets the curve shape, but we can still plot for some real-
istic numerical data, which are under-, neutral and over-steered, see Figure 4-31.
All cases in Figure 4-31 goes from positive side slip to negative when speed increases. This is the same
as we expected already in Figure 4-20.
We can also calculate and plot the longitudinal location of the motion centre, i.e. 𝑥𝑀 ⁄ , by
combining Eqs [4.22] and [4.20]. Note that 𝑅 is independent of 𝑓 , while the longitudinal location of
the motion centre, 𝑦𝑀 ⁄ , is not.
254
Lateral Dynamics
𝑚 √ 𝑚 𝑅𝑝 √min( 𝑓 ) 𝑅𝑝 [4.23]
𝑓
0.5 neutral steered
oversteered
𝑦𝑀
15
⁄𝑟
⁄ 𝑓 [1/rad]
0
𝑥𝑀 ⁄
[1/rad]
⁄𝑟
⁄ (vy/vx)/df
-0.5
10
Neutral steered
xMC = -vy/wz
-1
slip gain",
-1.5 5
gain,
side slip"side
-2
𝑥𝑀
0
understeered
-2.5
neutral steered
oversteered
-3 -5
0 5 10 15 20 25 0 5 10 15 20 25
longitudinalvx velocity,
[m/s] ⁄𝑠 longitudinal
vx velocity,
[m/s] ⁄𝑠
𝑣𝑦 𝑣𝑦
Figure 4-31: Left Side slip gain (𝑣 ∙𝛿𝑓
). Right: Motion centre longitudinal location (𝑥𝑀 𝜔𝑧
). For
Steady State Cornering.
255
Lateral Dynamics
.7 05 𝑁/𝑟
. 05 𝑁/𝑟
± 5 𝑘𝑁 ± 5 𝑘𝑁
Figure 4-32: The wheels cornering stiffness (𝜕 ⁄𝜕 ( )| ) changes degressively with vertical
𝑦=
load. The axle cornering stiffness therefore decreases with increased load transfer.
ℎ𝑒𝑟𝑒 𝑓 𝑛
+ +
𝑓 𝑓
For vehicles with largely varying vertical axle load (such as heavy trucks), one has to consider that the
contribution from tyre to axle cornering stiffness is rather proportional to vertical axle load, while the
contribution from side-force steering comes from suspension elasticities and is rather constant. So,
utilizing side-force steering makes the vehicles lateral manoeuvrability inconsistent with vertical load.
§ Influence of Translatory Compliance
256
Lateral Dynamics
One way to deliberately design for side-force steering is to let the axle or wheels have yaw pivot points
ahead or behind. This also leads to a lateral translatory compliance of the axle or wheels which,
strictly, also gives one more state variable. For lower frequencies, such as for handling, this compliance
can often be neglected.
4.3.7.3.2 Roll Steer Gradient
Roll steer gradient, 𝑘 , is defined for an axle and it is how much the wheels on an axle steers [deg]
per vehicle roll angle [deg]. Also, a non-steered axle can steer due to roll-steering. Roll-steering de-
pends on the suspension linkage geometry. The added steer angle can be expressed: Δ 𝑘 ∙ .
We will now derive the influence on steady state cornering. Add steering on rear axle to Eq [4.9]:
𝐿 𝑚∙𝑣 2 𝑟 ∙𝑙𝑟 𝑓 ∙𝑙𝑓
𝑓 ≈ +𝐾 𝑛 ∙ 𝐾 𝑛 (subscript “noRS” means “no Roll-Steer”)
𝑝 𝑝 𝑓 ∙ 𝑟 ∙𝐿
If we see as built up by one angle from the steering system and one part coming from the suspen-
sion, via roll-steering Δ :
𝐿 𝑚∙𝑣 2
𝑓 +Δ 𝑓 ( +Δ )≈ +𝐾 𝑛 ∙
𝑝 𝑝
257
Lateral Dynamics
force understeering rear at a two-axle vehicle. Using this concept can lead to very yaw stable vehicles.
The drawback is reduced yaw agility. If really exaggerated, it can take the rear axle to effectively nega-
tive cornering stiffness, which makes vehicle unstable.
258
Lateral Dynamics
Longer wheelbase (with unchanged yaw inertia and unchanged steering ratio) improves the transient
manoeuvrability, because the lateral forces have larger levers to generate yaw moment with.
259
Lateral Dynamics
can be designed from real vehicle tests as well. The slope in the handling diagram corresponds to un-
dersteering gradient 𝐾 3 in Equation [4.13].
𝐿
front axle, 𝑓 𝑓 𝑓
𝑙𝑟 𝑚
𝐿
rear axle,
𝑙 𝑚
From: Daniel A. Fittanto, et al. “Passenger Vehicle Steady-State Directional Stability Analysis
Utilizing EDVSM and SIMON”, Copyright 2004 by Engineering Dynamics Corporation
260
Lateral Dynamics
The relevance to study the load transfer during steady state cornering is to limit the roll during corner-
ing (for comfort) and yaw balance (understeering gradient, see 4.3.7.2). Additionally, the load transfer
influence the transient handling; see 4.4 and 4.4.3.5.
𝑙
w/2 w/2
Figure 4-35: A cornering vehicle. The is a fictive force. Subscript l and r mean left and right.
These equations confirm what we know from experience, the curve-inner side if off-loaded.
𝑙 𝑙 + ∙( 𝑙 𝑙) + ∙ (( 𝑙 𝑙) ( ))
{ 𝑙 0} 𝑙 ( + )∙ 𝑙 + ∙
[4.27]
⋯ ( + )∙ + ∙ 𝑙
ℎ𝑒𝑟𝑒 𝑙 + ∙ 𝑛 𝑙 ∙ ⁄ ∙ ⁄ 0
The stiffnesses and ( 𝑟 means anti-roll bar) are effective stiffnesses as measurable under
the wheels. The physical springs are mounted inside in some kind of linkage and have different stiff-
ness values, but their effect is captured in the effective stiffnesses. Some examples of different physical
spring and linkage design are given in 2.3.3.1 and 2.3.4.1.
We see already in free-body diagram in Figure 4-36 that 𝑙 and always act together, so we rename
𝑙 + . We see in Figure 2-61 that we have to assume something about how the lateral forces
are transferred from road to body. The “trivial linkage” from Figure 2-61 is assumed. Equilibrium
then gives:
∙ 0
∙ 𝑙 0 [4.28]
𝑙 ∙( ⁄ ) ∙ ( ⁄ ) + ∙ ℎ + ∙ ∙ ( 𝑦) 0
The term ∙ ∙ ( 𝑦) is taken as ∙ ∙ (ℎ ℎ ) ∙ sin( ) ≈ ∙ ∙ (ℎ ℎ ) ∙ . It assumes a height
for the point where the roll takes place, ℎ . We don’t know the value of it, until below where we study
261
Lateral Dynamics
the suspension design, but it can be mentioned already here that most vehicles have an ℎ ≪ ℎ. This
causes a “(roll) pendulum effect”, especially significant for heavy commercial vehicles due to their
large ℎ..
Compatibility, to introduce body displacements, and , gives:
𝑙 +( ⁄ )∙ 𝑛 ( ⁄ )∙ [4.29]
-y
(Anti-roll is drawn, only in left
picture and beside the vehicle,
ay z for better clarity in drawing.)
y m*ay
zl zr
, anti-roll bar
torsional spring
h m*g
zlr=0 Fzr
zrr=0 Fzl
w/2 w/2 w/2 w/2
Steady-state assumed, so that
z,y,px= ,zl,zr, are displacements lateral acceleration (ay) may be
from a static stand-still position. non-zero, but vertical and roll
zlr=zrr=0 means that road is smooth. acceleration are zero.
Figure 4-36: Model for steady state heave and roll due to lateral acceleration. Suspension model is
no linkage (or “trivial linkage”) and without difference front and rear.
Combining constitutive relations, equilibrium and compatibility, gives, as Matlab script:
syms …; sol=solve( ...
Flz==Flz0-(cside+carb)*zl+carb*zr, ...
Frz==Frz0-(cside+carb)*zr+carb*zl, ...
Flz0+Frz0==m*g, Flz0*w/2-Frz0*w/2==0, ...
Fy-m*ay==0, ...
m*g-Flz-Frz==0', ... [4.30]
Flz*(w/2)-Frz*(w/2)+Fy*h+m*g*(h-hRC)*px==0, ...
zl==z+(w/2)*px, zr==z-(w/2)*px, ...
Fzl1==-1/((Fzl/m-g/2)*w/(ay*h)), ...
zl, zr, Flz, Frz, Flz0, Frz0, Fy, z, px);
The results from the Matlab script in Equation [4.30]:
∙ 0
2 𝑚∙ 𝑦∙
𝑝 𝑥 ( 𝑖 )∙𝑤 2 2 𝑚∙ ∙(
+2∙ 𝑟 𝐶)
𝑦∙ 2 𝑚∙ 𝐶 [4.31]
𝑙 ∙ (2 ⁄( ∙ ))
𝑤 𝑖 +2 𝑟 𝑤2
𝑦∙ 2 𝑚∙ 𝐶
∙( + ⁄( ∙ ))
2 𝑤 𝑖 +2 𝑟 𝑤2
In agreement with intuition and experience the body rolls with positive roll when steering to the left
(positive 𝑤 ). Further, the body centre of gravity is unchanged in heave (vertical motion, ). The for-
mula uses ℎ which we cannot estimate without modelling the suspension. Since front and rear axle
normally are different, we could expect that ℎ is expressed in some similar quantities for each of
front and rear axle, which also is the case, see Equation [4.38].
4.3.10.2.1 Steady-State Roll-Gradient *
Function definition: Steady state roll-gradient is the body roll angle per lateral acceleration for the vehicle
during steady state cornering with a certain lateral acceleration and certain path radius on level ground.
262
Lateral Dynamics
263
Lateral Dynamics
RC h h h
ef ef hRCf
gf
w/2 w/2 w/2 w/2
Transversal sections from rear over rear axle:
Similar, but with subscript “r” instead of “f”. Typically, ℎ >ℎ 𝑓
Figure 4-37: Two alternative models for including suspension linkage effects (kinematics) in
lateral load transfer. Anti-roll bars not drawn.
4.3.10.3.2 Load Transfer Model with Roll Centre (One Pivot Point) per Axle
The model with 1 roll centres has some drawback as listen before. To mention some advantages, it is
somewhat less computational demanding. However, the main reason why the compendium uses this
model is to cover two different concepts with longitudinal and lateral load transfer.
Study the free-body diagrams in Figure 4-38.
Fsfl+Fsrl Fsfr+Fsrr
zflr=zrlr=0
+
zfrr=zrrr=0
𝑓
Figure 4-38: Model for steady state heave and roll due to lateral acceleration, using roll centres,
which can be different front and rear.
264
Lateral Dynamics
The road is assumed to be flat, 𝑓𝑙 𝑓 𝑙 0. In free-body diagram for the front axle,
𝑓 and 𝑓 are the reaction force in the rear roll-centre. Corresponding reaction forces are found for
rear axle. Note that roll centres are free of roll moment, which is the key assumption about roll centres.
The 𝑓𝑙 𝑓 𝑙 and are the forces in the compliances, i.e. where potential spring energy is
stored. One can understand the roll-centres as also unable to take vertical force, as opposed to con-
straining vertical motion (as drawn). Which of vertically force-free or vertically motion-free depends
on how one understands the concept or roll-centre, and it does not influence the equations.
Note carefully that the “pendulum effect” is NOT included here, in 4.3.10.3, as it was in 4.3.10.2. The
motivation is to get simpler equations for educational reasons.
There is no damping included in model, because their forces would be zero, since there is no displace-
ment velocity, due to the steady-state assumption. As constitutive equations for the compliances
(springs) we assume that displacements are measured from a static condition and that the compli-
ances are linear. Note that there are two elasticity types modelled: springs per wheel ( 𝑓𝑤 per front
wheel and 𝑤 per rear wheel) and anti-roll bars per axle ( 𝑓 front and rear). The road is assumed
to be smooth, i.e. 𝑓𝑙 𝑓 𝑙 0. The stiffnesses 𝑓𝑤 𝑤 𝑓 and are effective stiff-
nesses per wheel. We see already in free-body diagram that 𝑓𝑙 and 𝑓 always act together, so we
rename 𝑓𝑙 + 𝑓 𝑓 and 𝑙 + .
𝑓𝑙 𝑓𝑙 + 𝑓𝑤 ∙ ( 𝑓𝑙 𝑓𝑙 )
𝑓 𝑓 + 𝑓𝑤 ∙ ( 𝑓 𝑓 )
𝑙 𝑙 + 𝑤∙( 𝑙 𝑙)
+ 𝑤 ∙( )
[4.32]
𝑓 0+ 𝑓 ∙ (( 𝑓𝑙 𝑓𝑙 ) ( 𝑓 𝑓 ))
0+ ∙ (( 𝑙 𝑙) ( ))
∙ ∙ ∙ ∙ 𝑓
ℎ𝑒𝑟𝑒 𝑓𝑙 𝑓 𝑛 𝑙
∙𝐿 ∙𝐿
Equilibrium for whole vehicle (vertical, lateral, yaw, pitch, roll) neglecting body forces (air resistance
and gravity components in road plane) gives:
𝑓𝑙 + 𝑓 + 𝑙 + ∙
∙ 𝑓 +
0 ∙ ∙
𝑓 𝑓 [4.33]
( 𝑓𝑙 + 𝑓 )∙ 𝑓 +( 𝑙 + )∙ 0
( 𝑓𝑙 + 𝑙 )∙ ( 𝑓 + )∙ +( 𝑓 + )∙ℎ 0
Equilibrium for each axle (roll, around roll centre). Note that the axles are considered massless:
( 𝑓𝑙 𝑓𝑙 + 𝑓) ∙ ( 𝑓 𝑓 𝑓) ∙ + 𝑓 ∙ℎ 𝑓 0
[4.34]
( 𝑙 𝑙 + )∙ ( )∙ + ∙ℎ 0
𝑓𝑙 + ∙ 𝑓 ∙ 𝑛 𝑓 ∙ 𝑓 ∙
+ ∙ + ∙ 𝑛 ∙ + ∙ [4.35]
𝑙
𝑓𝑙 + 𝑓 0 𝑛 𝑙 + 0
The measure h is redundant and can be connected to the other geometry measures as follows. The
geometrical interpretation is given in Figure 4-39.
265
Lateral Dynamics
∙ℎ 𝑓 + 𝑓 ∙ℎ
ℎ ℎ [4.36]
𝐿
Combining Equations [4.32] to [4.36] gives, as Matlab script and solution:
syms …; sol=solve( ...
Fsfl==Fsfl0-cfw*zfl, Fsfr==Fsfr0-cfw*zfr, ...
Fsrl==Fsrl0-crw*zrl, Fsrr==Fsrr0-crw*zrr, ...
Faf==0-caf*(-zfl+zfr), Far==0-car*(-zrl+zrr), ...
Fsfl0==(1/2)*m*g*lr/L, Fsfr0==(1/2)*m*g*lr/L, ...
Fsrl0==(1/2)*m*g*lf/L, Fsrr0==(1/2)*m*g*lf/L, ...
Fflz+Ffrz+Frlz+Frrz==m*g, ...
m*ay==Ffy+Fry, ...
0==Ffy*lf-Fry*lr, ...
-(Fflz+Ffrz)*lf+(Frlz+Frrz)*lr==0, ...
(Fflz+Frlz)*w/2-(Ffrz+Frrz)*w/2+(Ffy+Fry)*h==0, ... [4.37]
(Fflz-Fsfl+Faf)*w/2-(Ffrz-Fsfr-Faf)*w/2+Ffy*hRCf==0, ...
(Frlz-Fsrl+Far)*w/2-(Frrz-Fsrr-Far)*w/2+Fry*hRCr==0, ...
zfl==z+(w/2)*px-lf*py, zfr==z-(w/2)*px-lf*py, ...
zrl==z+(w/2)*px+lr*py, zrr==z-(w/2)*px+lr*py, ...
zfl+zfr==0, zrl+zrr==0, ...
dh==h-(lr*hRCf+lf*hRCr)/(lf+lr), ...
zfl, zfr, zrl, zrr, Fsfl, Fsfr, Fsrl, Fsrr, ...
Faf, Far, Fsfl0, Fsfr0, Fsrl0, Fsrr0, ...
Fflz, Ffrz, Frlz, Frrz, Ffy, Fry, z, px, py, h);
The result from the Matlab script in Equation [4.37], but in a prettier writing format:
𝑙𝑟 𝑙𝑓
𝑓 ∙ ∙ 𝐿
𝑛 ∙ 𝐿
∙
𝑚∙ 𝑦 ∙ ( 𝑓𝑦 + 𝑟𝑦 )∙
0 𝑛 𝑝 𝑛 𝑝 0
𝑣 𝑖 𝑟𝑜 𝑣 𝑖 𝑟𝑜
∙ ℎ 𝑓 ∙ ℎ 𝑓 𝑙𝑙
𝑓𝑙 ∙( ∙( + ∙ ))
∙𝐿 𝐿∙ 𝑣 𝑙 𝑙𝑙
∙ ℎ 𝑓 ∙ ℎ 𝑓 𝑙𝑙
𝑓 ∙( + ∙( + ∙ ))
∙𝐿 𝐿∙ 𝑣 𝑙 𝑙𝑙
∙ 𝑓 ℎ ∙ 𝑓 ℎ 𝑙𝑙
𝑙 ∙( ∙( + ∙ )) [4.38]
∙𝐿 𝐿∙ 𝑣 𝑙 𝑙𝑙
∙ 𝑓 ℎ ∙ 𝑓 ℎ 𝑙𝑙
∙( + ∙( + ∙ ))
∙𝐿 𝐿∙ 𝑣 𝑙 𝑙𝑙
The axle roll stiffnesses, 𝑓 𝑙𝑙 and 𝑙𝑙 are identified beside vehicle roll stiffness 𝑣 𝑙 𝑙𝑙 . We
should compare Equation [4.38] with Equation [4.31]. Eq [4.31] considers the “(roll) pendulum effect”,
but not the differentiation between front and rear suspension. Eq [4.38] does the opposite.
Assume ℎ ℎ and look at the sum of vertical force on one side, 𝑙 in Eq [4.31]. Compare 𝑙 in Eq
[4.31] and 𝑓𝑙 + 𝑙 in Eq [4.38]; the equations agree if:
ℎ 𝑓 ∙ +ℎ ∙ 𝑓 ℎ ℎ
𝑓𝑙 + 𝑙 𝑙 ⇒ ∙( ∙( + )) ∙( ∙ ) 𝑙 ⇒
𝐿∙
ℎ 𝑓 ∙ +ℎ ∙ 𝑓 ℎ ℎ
⇒ + ⇒ℎ 𝑓 ∙ +ℎ ∙ 𝑓 (ℎ ℎ) ∙ 𝐿
𝐿∙
266
Lateral Dynamics
This is exactly in agreement with the definition of the redundant geometric parameter h, see Eq
[4.36]. This means that a consistent geometric model of the whole model is as drawn in Figure 4-39.
Here the artefact roll axis is also defined.
The terms of type ℎ ∙ ⁄(𝐿 ) in Eq [4.38] can be seen as the part of the lateral tyre forces that
goes via the stiff linkage. The terms of type (Δℎ ⁄ ) ( 𝑤 ⁄( 𝑤 + 𝑤 )) in Eq [4.38] can be seen as the
part of the lateral tyre forces that goes via the compliance. The latter part is distributed in proportion
to roll stiffness of the studied axle, as a fraction of the vehicle roll stiffness. This should agree with in-
tuition and experience from other preloaded mechanical systems (load distributes as stiffness).
Body rolls with positive roll when steering to the left, as long as CoG is above roll axle. Further, the
body centre of gravity is unchanged in heave (vertical z) because the model does not allow any vertical
displacements, which is a drawback already mentioned.
front axle
CoG roll centre
front axle
h
hRCf
rear axle
roll centre
h
rear axle
ℎ 𝑓 + 𝑓 ℎ
hRCr
ℎ ℎ
𝐿
Figure 4-39: Roll axis for a two-axle vehicle. (Note that the picture may indicate that the roll
centres and roll axis are above wheel centre, but this is normally not the case.)
Eq [4.44] was derived as a steady state out-of-road-plane model, but only the ratio between roll stiff-
nesses influence the lateral load transfer. So, if the roll stiffnesses are large, they can be neglected (con-
sidered infinite), if the ratios are given. Then, Eq [4.44] works also for transient manoeuvres.
4.3.10.3.3 Steady State Longitudinal and Lateral Distribution
If the vehicle has a steady state acceleration with combined and , we can combine Eqs [4.38]and
[3.27] (with 𝑓𝑥 + 𝑟𝑥
( ) ) to Eq [4.39]. Note:
• Body forces (air resistance and gravity components in road plane) are neglected.
• We have also assumed symmetric longitudinal load transfer. This is reasonable if no roll pre-
tension. Pre-tension could appear if uneven ground or unusual suspension, e.g. pre-tensioned
(or “active”) anti-roll bar.
∙𝑙 𝐶𝑓 ∙𝑙𝑟 𝑓 𝑟𝑜
𝑓𝑙 ∙ ( 2∙𝐿𝑟 2𝐿
∙( 𝐿∙𝑤
+ 𝑤
∙ ))
𝑣 𝑖 𝑟𝑜
∙𝑙 𝐶𝑓 ∙𝑙𝑟 𝑓 𝑟𝑜
𝑓 ∙ ( 2∙𝐿𝑟 2𝐿
+ ∙( 𝐿∙𝑤
+ 𝑤
∙ ))
𝑣 𝑖 𝑟𝑜
[4.39]
∙𝑙𝑓 𝐶𝑟∙𝑙𝑓 𝑟 𝑟𝑜
𝑙 ∙ ( 2∙𝐿 + 2𝐿
∙( 𝐿∙𝑤
+ 𝑤
∙ ))
𝑣 𝑖 𝑟𝑜
∙𝑙𝑓 𝐶𝑟∙𝑙𝑓 𝑟 𝑟𝑜
∙ ( 2∙𝐿 + 2𝐿
+ ∙( 𝐿∙𝑤
+ 𝑤
∙ ))
𝑣 𝑖 𝑟𝑜
267
Lateral Dynamics
268
Lateral Dynamics
Tracking ability
road and 𝑟 𝑡𝑦
is this swept width
translation
direction
cross-fall
direction
2 2
2
𝑤
2 2
2 2
+ 3
Figure 4-41: Tracking ability on straight path, TASP, for an “A-double”. Longitudinal forces
neglected. Axles within an axle group lumped together, ⋯ 5. (Longitudinal dimensions as
in Figure 1-50.)
An example is seen in Figure 4-41. If we neglect combined tyre slip and assume same cornering coeffi-
cient, , on all axles we can derive these equations, one for each axle group:
• 𝑤 ( )
• 2 2
• + + 𝑓𝑜𝑟 𝑢𝑛 𝑡 . .4
Since the levers for moment equilibria in road 𝑥𝑦-plane and in road 𝑥 -plane are equal, the distribu-
tion of the axles’ vertical forces and lateral forces becomes identical. So, relation between lateral force
𝑔𝑗𝑦 𝑔1𝑦𝑣 𝑔1𝑦𝑤
and vertical force becomes tan( )≈ for all axle groups 𝑗 . .5 and ≈ ≈ :
𝑔𝑗𝑧 𝑔1𝑧 𝑔1𝑧
•
𝑔1𝑦𝑤
≈ ≈ ( )
𝑔1𝑧
•
𝑔2𝑦
≈ ≈
𝑔2𝑧
•
𝑔 𝑖+1 𝑦
≈ ≈ 𝑓𝑜𝑟 𝑢𝑛 𝑡 . .4
𝑔 𝑖+1 𝑧
The model is very approximative, since it does not take the following aspects into account:
• Axles within an axle group can take different vertical force.
• Axles can have different cornering coefficient .
• Large road grades. This can influence via longitudinal forces and both articulation angles and
combined tyre slip. The effect would be larger if combined with low road-friction.
More accurate value of TASP can be found by simulation to a steady-state pose of the vehicle.
269
Lateral Dynamics
270
Lateral Dynamics
distribution changes, so that a “knee” on the curves appears, see Figure 4-42. So, the relation of type as
Equation [4.38] is no longer valid. For instance, it is not physically motivated to keep the roll-centre
model for an axle which has lifted one side. So, the prediction of critical lateral acceleration for roll-
over is not trivial, especially for heavy vehicles which has many axles, and often also a fifth wheel
which can transfer roll-moment to a certain extent. There are approximate standards for how to calcu-
late steady state roll-over thresholds for such vehicle, e.g. UN ECE 111 (http://www.unece.org/filead-
min/DAM/trans/main/wp29/wp29regs/r111e.pdf).
model Vehicle_1Unit_NAxles Vertical force on
…
inner wheels [N]veh.F_alz[2]
first inner
second inner
wheel lift
equation veh.F_alz[1] veh.F_alz[3]
wheel lift
//Test case (Steady state cornering, incr. a_y, until left lifts):
der(p_x)=1; // p_x=0.01; // 5E4
when sum(Lift)>=na then
terminate("Both left wheels lifted");
end when; 4E4
last inner
wheel lift
//Whole unit (=vehicle in this case):
//Equilibria irp: //Lateral: 3E4
0= sum(F_ay) - m*a_y;
//Yaw:
0 = sum(F_ay.*(x_a-x_CoG*ones(na))); 2E4
//Equilibria oorp: //Vertical:
0 = sum(F_alz+F_arz) - m*g;
//Pitch:
0 = sum((F_alz+F_arz).*(x_a-x_CoG*ones(na))); 1E4 Lateral
//Roll, around mid ground:
0=m*a_y*h + sum(F_alz)*w/2 - sum(F_arz)*w/2; acceleration
0E0
//For each axle:
for i in 1:na loop
//Equilibria: //Roll, around pivot point: -1E4
0 = F_ay[i]*h_aRC[i] + (F_alz[i] - F_arz[i])*w/2 - M_ax[i]; 0.0 2.5 5.0
//Constitution: //Tyre:
F_ay[i]=-CC_y*(F_alz[i]+F_arz[i])*s_ay[i]; veh.a_y [m/s²]
s_ay[i]=v_ay[i]/v_x;
v_ay[i]=v_y+(x_a[i]-x_CoG)*w_z;
//Suspension: Lateral force
der(M_ax[i]) = c_a[i]*(der(p_ax[i]) - der(p_x));
if Lift[i] > 0.5 then on axles [N]
veh.F_ay[1] veh.F_ay[2] veh.F_ay[3]
der(F_alz[i]) = 0;
else der(p_ax[i]) = 0; end if;
when (pre(Lift[i]) < 0.5) and F_alz[i] < 0 then
Lift[i] = 1; reinit(p_ax[i], 0); end when; 4E4
end for; Lateral
//Suspension Ctrl. Generates eqs only if number of axles=na>2:
for i in 2:(na-1) loop
acceleration
der(F_alz[i])+der(F_arz[i])=der(F_alz[i+1])+der(F_arz[i+1]); 0E0
end for;
end Vehicle_1Unit_NAxles; 0.0 2.5 5.0
veh.a_y [m/s²]
Figure 4-42: Example of 3-axle vehicle steady state roll-over wheel lift diagram.
4.3.12.1.1 § Roll-Over Model
One way to assess the steady state roll-over lateral acceleration is to model a dynamic system with roll
angle as independent variable (instead of time) and then simulate (i.e. study over a lapse with increas-
ing roll angle). When all wheels on one side lifts from ground, we define it as steady state roll-over lat-
eral acceleration. The model is then much more intuitively physical than, e.g. UN ECE 111 formulas.
Figure 4-42 shows a Modelica model of a 1-unit N-axles vehicle. Note that, since 3 axles means non-
Ackermann steering geometry, it necessarily involves the lateral tyre slip stiffnesses also. The model
can, with maintained physical interpretation, be extended to more units with roll-free or roll-rigid con-
nections. Phenomena as (roll) pendulum effect and lateral deformation of tyres can also be added.
An alternative way of computing the critical lateral acceleration is to use the same model, but not sim-
ulate it. Instead one could solve the states where assume that all axles except one are lifted and then
see how much lateral acceleration is needed to lift the last axle. Trying each axle to be the last to lift
gives several lateral acceleration values, whereof the smallest is the critical lateral acceleration.
271
Lateral Dynamics
the vehicle shall not roll-over for steady-state cornering on level ground with an enough friction coeffi-
cient. Another is that it should not roll-over in a tilt-table. Since the requirement is not truly perfor-
mance based, each interpretation will also stipulate a certain verification method; here it would be
theoretical verification using a rigid suspension model. Such model and threshold are shown in Figure
4-43.
The derivation of the SSF based requirement looks as follows:
∙ + ∙ ∙ℎ ∙ ∙
ℎ∙
𝑀𝑜 𝑒 : { + ∙ }⇒ ∙ ∙( )
⇒
∙ ∙( + )
𝑅𝑒𝑞𝑢 𝑟𝑒 𝑒𝑛𝑡: ≥0 } [4.42]
ℎ∙
⇒ 𝑅𝑒𝑞𝑢 𝑟𝑒 𝑒𝑛𝑡: > ⇒ 𝑆𝑆 >
∙ℎ
Maximum road friction, , is typically 1, which is why SSF>=1 would be a reasonable. However, typi-
cal values of SSF for passenger vehicles are between 0.95 and 1.5. For heavy trucks, it can be much
lower, maybe 0.3..0.5, much depending on how the load is placed. There are objections to use SSF as a
measure, because SSF ignores suspension compliance, handling characteristics, electronic stability
control, vehicle shape and structure.
view from rear, when turning left
ay
m*g
m*ay
h
Fy
Fiz≥0
Foz
(=roll-over w/2
threshold)
w
Figure 4-43: Model for verification of requirement based on Static Stability Factor, SSF.
272
Lateral Dynamics
∙ ℎ 𝑓 ∙ ℎ 𝑓 𝑙𝑙
𝑓𝑙 ∙( ∙( + ∙ ))
∙𝐿 𝐿∙ 𝑙𝑙 𝑣 𝑙
∙ ∙ ℎ 𝑓 ℎ 𝑓 𝑙𝑙
∙ ∙ ∙ ⏟∙ ∙ ∙
⏟ ∙𝐿 ⏟ 𝐿 ⏟ 𝑙𝑙 𝑣 𝑙
𝐿 𝑡𝑒𝑟 𝑜𝑟 𝑒 𝑛𝑉𝑒ℎ 𝑒
𝑉𝑒𝑟𝑡 𝑜𝑟 𝑒 𝑛𝐴𝑥 𝑒 𝐿 𝑡𝑒𝑟 𝑜𝑟 𝑒 𝑛𝐴𝑥 𝑒 𝑅𝑜 𝑆𝑡 𝑓𝑓𝑛𝑒𝑠𝑠 𝑟 𝑡 𝑜𝑛 𝑛𝐴𝑥 𝑒
ℎ ℎ 𝑙𝑙
𝑦( )∙ 𝑦( )∙ ∙
𝑙𝑙 𝑣 𝑙
If the vehicle has more axles, Eq [4.38] is generalized to Eq [4.43], which also is valid until first inner
wheel lifts from ground.
For axle of a roll-stiff vehicle:
ℎ ℎ 𝑙𝑙 [4.43]
𝑙 ( )∙ ( )∙ ∙
𝑙𝑙 𝑣 𝑙
For a vehicle with >2 axles, the parameter Δℎ can not be calculated from Eq [4.36], but can still be un-
derstood as the vertical distance between roll axis and the axles roll centre. It should be noted that the
pendulum effect is not included in Eq [4.43], and this is often a significant approximation if applied on
high CoG vehicles, like heavy trucks.
4.3.12.3.1 Model Assuming All Inner Wheels Lift at the Same Lateral Accelerat-
ion
An approximation of Steady state cornering roll-over acceleration (lateral acceleration when all
inner wheels lifted) can be found for vehicles where Eq [4.43] gives 𝑙 0 for all axles at same .
Then, summing the Eq [4.43] for all axles leads to the Eq [4.44] which is the same as the simple
SSF model in Figure 4-43 and 4.3.12.2 gives.
[4.44]
ℎ
In the following, we will elaborate with 4 additional effects, which marked in Figure 4-44.
• The tyre will take the vertical load on its outer edge in a roll-over situation. This suggests a
𝑦 𝑤+𝒘𝒕𝒚𝒓𝒆 𝑤+𝒘𝒕𝒚𝒓𝒆
change of performance and requirement to: < 2∙ and 2∙ > . This effect is accen-
tuated when low tyre profile and/or high inflation pressure. This effect decreases the risk for
roll-over.
• Due to suspension and tyre lateral deformation, the body will translate laterally outwards,
𝑦 𝑤 𝑫𝒆𝒇𝒚 𝑤 𝑫𝒆𝒇𝒚
relative to the tyre. This could motivate < 2∙
and 2∙
> . This effect increases the
risk for roll-over.
• Due to suspension linkage and compliances, the body will roll. Since the CoG height above roll
𝑦 𝑤 𝒉∙ 𝑤 𝒉∙
axis, ℎ, normally is positive, this could motivate < 2∙ 𝒙 and 2∙ 𝒙 > . This effect in-
creases the risk for roll-over. At heavy vehicle this “(roll) pendulum effect” is large.
• Due to suspension linkage and compliances, the body will also heave. This requires a suspen-
sion model with pivot points per wheel, as opposed to roll-centre per axle. The heave is nor-
𝑤 𝑦 𝑤
mally positive. This could motivate 2∙( + ) > and < 2∙( + ). The effect is sometimes called
“jacking” and it increases the risk for roll-over.
• Road leaning left/right (road banking) or driving with one side on a different level (e.g. out-
side road or on pavement) also influence the roll-over performance.
273
Lateral Dynamics
ℎ m*ay
h
Fy
𝐷𝑒𝑓𝑦
Fiz=0
Foz
(=roll-over w/2
threshold) wtyre/2
w
Figure 4-44: Steady-state roll-over model, with fore/aft symmetry. The measures 𝐷𝑒𝑓 ℎ∙
𝑛 mark effects additional to what is covered with a simple SSF approach.
4.3.12.3.2 Model with Sequential Lifts of Inner Wheels
A model which does not assume wheel lift at same lateral acceleration will be sketched. For each axle
that has lifted, the equations have to be changed. Instead of simply the constitutive equation
( 𝑙 ) ⁄ 𝑙𝑙 one need to assure 𝑙 0 . The axle will then position itself so that it
keeps 𝑙 + . That means both roll and vertical translation of the axle centre, why also
the vertical suspension compliance needs to be modelled. A new position variable has to be declared,
e.g. the lift distance of the inner wheel, 𝑓𝑡 𝑙 . This variable is constrained to 𝑓𝑡 𝑙 0 before lift, but
after lift the constraint is 𝑙 0 . So, the model is suitably implemented as a state event model with
the event “ ℎ𝑒𝑛 𝑙 𝑒 𝑜 𝑒𝑠 < 0”. If is swept from zero and upwards, the result will be something
like shown in Figure 4-42.
4.3.12.3.3 Using a Transient Model for Steady-State Roll-Over
Another work-around to avoid complex algebra is to run a fully transient model, including suspension,
and run it until a steady state cornering conditions occur. If then, the lateral acceleration is slowly in-
creased, one can identify when or if the roll-over threshold is reached. Lateral acceleration increase
can be through either increase of longitudinal speed or steer angle. It should be noted that the model
should reasonably be able to manage at least lift of one wheel from the ground. This way of verifying
steady state cornering roll-over requirements has the advantage that, if using tyre models with friction
saturation, the limitation discussed in 4.3.12.3.2 does not have to be checked separately.
274
Lateral Dynamics
275
Lateral Dynamics
velocities: forces:
𝑓 𝑣 𝑓 𝑓 𝑣 𝑓
𝑓 𝑣
𝑓 𝑣
𝑓 𝑓
𝑓 𝑓
𝑓
𝐿
𝒗
sin
Figure 4-46: One-track model for transient dynamics. Dashed show fictive forces. Compare to
Figure 4-17. Note that road is assumed to be level in 4.4, i.e. 0.
It is assumed that the longitudinal tyre forces and slips are small, so 𝑤 (𝑠 ) ≈ 𝑤 (𝑠 𝑤 ). It is also
assumed that lateral tyre forces are far below road friction saturation, 𝑤 (𝑠 𝑤 ) ≈ 𝑠 𝑤.
It can be difficult to understand the difference between acceleration [ ] and derivatives [ ] of
velocities [v v ]. Some explanations are proposed in 4.4.2.3.
The Mathematical model shown in Eq [4.46] in Modelica format. The subscript and refers to vehi-
cle coordinate system and wheel coordinate system, respectively. The model is developed from the
model for steady state in Eq [4.6], with the changes marked with yellow and underlined text.
//(Dynamic) Equilibrium:
m*ax = Ffxv + Frx; m*ay = Ffyv + Fry; J*der(wz) = Ffyv*lf - Fry*lr;
ax=der(vx)-wz*vy; ay=der(vy)+wz*vx;
//Compatibility:
vfxv = vx; vfyv = vy + lf*wz;
vrx = vx; vry = vy - lr*wz;
276
Lateral Dynamics
shown in Figure 4-47. The manoeuvre selected is same steering wheel function of time as in Figure
4-18, for better comparison of the different characteristics of the models.
exact.df_deg
10
0
𝑓 [deg] exact.y
-10 100
0 10 20 30
y [m]
exact.vx exact.w z_dps 90
smalAngleApprox.vx small_vy_w z_Approx.vx smalAngleApprox.w z_dps small_vy_w z_Approx.w z_dps small_vy_
40
80
30 70
60
20
50
10 40
30
0
20
-10 10
0
-20
x[m]
-10
time [s] 0 50 100
-30
exact.x
0 10 20 30
Figure 4-47: Simulation results of transient one-track models. Solid: Eq [4.46] or [4.47]. Dashed:
Eq [4.48], which employs the small angle approximation. Dotted: Eq[4.46]=[4.47] but without the
term .
Eq [4.46] is a complete model suitable for simulation. Removing [𝑥 𝑦 p ] and the 3 equations for
“path with orientation”, eliminating some variable and rewrite (no added approximations) gives:
Equilibrium: ∙( ∙ ) 𝑓 𝑤 ∙ cos( 𝑓 ) 𝑓 𝑤 ∙ sin( 𝑓 ) +
∙( + ∙ ) 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓) +
∙ ( 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓 )) ∙ 𝑓 ∙
Constitution: 𝑓 𝑤 𝑓 ∙ 𝑠𝑓 𝑤 ∙𝑠 𝑤
[4.47]
Compatibility: 𝑠𝑓 𝑤 𝑓 𝑤 | 𝑓 𝑤 | ∎ and 𝑠 𝑤
⁄ ( ∙ )⁄| |
Transformation from vehicle to wheel coordinate system on front axle:
𝑓 𝑤 ( + 𝑓 ∙ ) ∙ sin( 𝑓 ) + ∙ cos( 𝑓 ) ∎
𝑓 𝑤 ( + 𝑓 ∙ ) ∙ cos( 𝑓 ) ∙ sin( 𝑓 ) ∎
Typically, this model is used for simulation, where 𝑓 , 𝑓 𝑤 and are input variables. Suitable state
variables are , and . It is a non-linear model suitable for arbitrary transient manoeuvres and we
will come back to this in 4.4.3.5.
277
Lateral Dynamics
𝑣𝑦 +𝑙𝑓 𝜔𝑧 𝑣𝑦 𝑙𝑟 ∙𝜔𝑧
Compatibility: 𝑓 𝑓 + arctan(𝑠𝑓 𝑤) 𝑓 |𝑣 |
𝑠 𝑤 |𝑣 |
Now, we assume small angles: steering angle 𝑓 , wheel side slip angles ≈ arctan( 𝑤 ⁄ 𝑤 )
arctan(𝑠 𝑤 ), and body side slip angle over steered axle 𝑓 ≈ arctan( 𝑓 ⁄ ). These are “trigonometric
approximations” during the Mathematical modelling stage, motivated if not too sharp turning.
Equilibrium: ∙( ∙ ) 𝑓 𝑤 + 𝑓 𝑤 ∙ 𝑓
∙( + ∙ ) 𝑓 𝑤 ∙ 𝑓 + 𝑓 𝑤 +
∙ ( 𝑓 𝑤 ∙ 𝑓 + 𝑓 𝑤) ∙ 𝑓 ∙ [4.48]
Constitution: 𝑓 𝑤 𝑓 ∙ 𝑠𝑓 𝑤 ∙ 𝑠 𝑤
𝑣𝑦 +𝑙𝑓 𝜔𝑧 𝑣𝑦 𝑙𝑟 ∙𝜔𝑧
Compatibility: 𝑓 ≈ 𝑓 + 𝑠𝑓 𝑤 𝑓 ≈ |𝑣 |
𝑠 𝑤 |𝑣 |
(Note that 𝑤 ∙ can be employed as approximation already during tyre modelling. This
would not change the resulting Eq [4.48].)
Figure 4-47 shows a simulation with model in Eq [4.48], for comparison with the model without small
angle approximations, i.e. from Eq [4.47].
Eliminating 𝑓 𝑤 𝑠𝑓 𝑤 𝑠 𝑓 gives:
𝑣𝑦+𝑙𝑓 𝜔𝑧
∙( ∙ ) 𝑓 𝑤 + 𝑓 ∙( |𝑣 | 𝑓) ∙ 𝑓 +
𝑣𝑦 +𝑙𝑓 𝜔𝑧 𝑣𝑦 𝑙𝑟∙𝜔𝑧
∙( + ∙ ) 𝑓 𝑤 ∙ 𝑓 𝑓 ∙ ( |𝑣 | 𝑓) ∙ |𝑣 |
𝑣 +𝑙 𝜔 𝑣𝑦 𝑙𝑟 ∙𝜔𝑧
∙ ( 𝑓 𝑤 ∙ 𝑓 𝑓 ∙ ( 𝑦 |𝑣𝑓| 𝑧 𝑓 )) ∙ 𝑓 + ∙ |𝑣 | ∙
Yet another approximation which we can do is to assume that centripetal acceleration is directed
purely lateral in vehicle and hence remove the term ∙ . Figure 4-47 shows a simulation of Eq
[4.47] without the term ∙ for judging the importance of it; it has considerable influence on
over time. Also, the influence of the force 𝑓 𝑤 ∙ 𝑓 in longitudinal direction can be neglected. Eq [4.49]
shows the resulting model.
∙ 𝑓 𝑤 +
+ 𝑓 ∙
∙( + ∙ ) 𝑓 𝑤 ∙ 𝑓 𝑓 ∙( 𝑓) ∙
| | | | [4.49]
+ 𝑓 ∙
∙ ( 𝑓 𝑤∙ 𝑓 𝑓 ∙( 𝑓 )) ∙ 𝑓 + ∙ ∙
| | | |
For stationary oscillation steering, 0, so can be seen as a manoeuvre-parameter. This is often a
good approximation also when turning during mild acceleration and deceleration. But, instead of set-
ting 0, we keep it more generic: is considered as a known function of time (𝑡), i.e. and
are known functions of time, or is an input and a state. So, we can continue to select as
states, but inputs become 𝑓 𝑓 𝑤 :
The first scalar equation (the longitudinal equilibrium) in Eq [4.49] can be used to calculate what is
needed to keep the prescribed (𝑡): ∙ (𝑡) 𝑓 𝑤 . Note that neither aerodynamic, grade nor
rolling resistance are considered.
The two latter scalar equations can be written as a linear model on matrix form, with 2 states
and 2 inputs 𝑓 𝑓 𝑤 𝑓 , see Eq [4.50], where the output is chosen as 𝒚𝑻 [ ]
[ + ]. The (𝑡) is often a scalar constant .
278
Lateral Dynamics
𝑓
[ ] 𝑨∙[ ]+𝑩∙[ ]
𝑓𝑥 𝑓 𝒙 𝑨∙𝒙+𝑩∙
or {
𝑓 𝒚 𝑪∙𝒙+𝑫∙
[ ] 𝑪∙[ ]+𝑫∙[ ]
{ 𝑓𝑥 𝑓
𝑓+ 𝑟 𝑓 ∙ 𝑙𝑓 𝑟 ∙ 𝑙𝑟
+ ∙ (𝑡)
0 |𝑣 ( )| |𝑣 ( )|
[4.50]
Where 𝑨 [ ] ∙[ 2
+ 𝑟 ∙ 𝑙𝑟 2
]
0 𝑓 ∙𝑙𝑓 𝑟 ∙𝑙𝑟 𝑓 ∙ 𝑙𝑓
|𝑣 ( )| |𝑣 ( )|
0 𝑓
and 𝑩 [ ] ∙[ ]
0 𝑓 ∙ 𝑓 𝑓
0 0 0 0 0
and 𝑪 [ ]∙𝑨+[ (𝑡)]; and 𝑫 [ ] 𝑩;
0 0 0
The model above is approximated only for driving situation where tyre forces are far from saturation
and angles are small. If studying driving conditions which do not fulfil this, we can still approximate
with linearized models if variation from the driving condition is small. The method to linearize is to
use only the first terms in a Taylor series expansion of the model 𝒙 𝒇(𝒙 ) around [𝒙 ] [𝒙𝟎 𝟎 ]:
𝒅𝒇 𝒅𝒇
𝒙 𝒇(𝒙 ) ≈ 𝒇(𝒙𝟎 𝟎 ) + | (𝒙 𝒙𝟎 ) + | ( 𝟎)
𝒅𝒙
⏟ 𝒙𝟎 𝟎 𝒅
⏟ 𝒙𝟎 𝟎
𝑨𝒍𝒊𝒏 𝑩𝒍𝒊𝒏
𝒅𝒇 𝒅𝒇 𝒅𝒇 𝒅𝒇
(𝒇(𝒙𝟎 𝟎) | 𝒙𝟎 | 𝟎) + | 𝒙+ |
⏟ 𝒅𝒙 𝒙𝟎 𝟎
𝒅 𝒙𝟎 𝟎
𝒅𝒙
⏟ 𝒙𝟎 𝟎
𝒅
⏟ 𝒙𝟎 𝟎
𝒇𝟎 𝑨𝒍𝒊𝒏 𝑩𝒍𝒊𝒏
This can be used when studying small input offsets at a steady state cornering ( 𝑓 𝑤 𝑓 𝑓 𝑤 𝑓
and 𝑓 𝑓 ). It can also be used for finding eigen solutions.
The coefficients in front of ̈ 𝑓 and 𝑓 can be compared with coefficients in simpler damped
mass and spring systems, see 1.5.2.2.1. Note, for instance, that the coefficients are dependent of and
that 𝑓 and cannot be simply identified as either springs or dampers.
Fourier transform gives:
∙ ∙ 2 2 ℱ( ) +
279
Lateral Dynamics
2 2)
+ (( 𝑓 + )∙ +( 𝑓 ∙ 𝑓 + ∙ ∙ )∙| |∙𝑗 ℱ( )+
2
+( 𝑓 ∙ ∙(𝑓+ ) +( ∙ 𝑓 ∙ 𝑓) ∙ ∙ ∙ | |) ∙ ℱ( )
2
𝑓 ∙ 𝑓 ∙ ∙ ∙𝑗 ℱ( 𝑓 ) (𝑓+ )∙ 𝑓 ∙ ∙ | | ∙ ℱ( 𝑓 ) ⇒
ℱ( )
⇒ 𝐻𝛿𝑓 →𝜔𝑧 ( )
ℱ( 𝑓 )
2∙
𝑓 ∙𝑙𝑓 ∙𝑚∙𝑣 𝜔+(𝑙𝑓 +𝑙𝑟 )∙ 𝑓 ∙ 𝑟 ∙|𝑣 |
2 2 2 2
𝐽∙𝑚∙𝑣 𝜔2 (( 𝑓 + 𝑟 )∙𝐽+( 𝑓 ∙𝑙𝑓 + 𝑟 ∙𝑙𝑟 )∙𝑚)∙|𝑣 |∙ 𝜔 ( 𝑓 ∙ 𝑟 ∙(𝑙𝑓 +𝑙𝑟 ) +( 𝑟 ∙𝑙𝑟 𝑓 ∙𝑙𝑓 )∙𝑚∙𝑣 ∙|𝑣 |)
This gives the same transfer function as in 4.4.3.3.1, which is derived with matrix formula for the “1 st
order, 2 state model”. A single state, 2nd order model for can be derived similarly.
Compatibility:
bf 𝑓 + 𝑓
𝑓 𝑓 + 𝑠𝑓 ≈ 𝑓 + 𝑓 𝑓 ≈
af
b df 𝑠 ≈ ≈
br
Eliminate 𝑓 𝑓 𝑓 yields:
𝑓+ 𝑓 𝑓
𝑓 + + + ≈ 𝑓 𝑓
𝐿 𝑓 𝑓 𝑓
2
𝑓 +
2
+ + ≈ 𝑓 𝑓 𝑓
Figure 4-48: Less general derivation of the Linear One-Track Model, i.e. Eq [4.50].
One can add more axles to the derivation. E.g., a truck with double rear axle will get an added term
proportional to in the moment equilibrium.
See Equation [4.45]. This section provides some explanations to the difference between acceleration
[ ] and derivatives [ ] of velocities [ ].
4.4.2.3.1 Theoretical explanation
First, think of and as geometric vectors (acceleration and velocity are vectors in physics). Accelera-
tion is acceleration in inertial frame, i.e. accelerations over ground. Velocity is velocity in inertial
frame, i.e. velocity over ground. A driver or accelerometers, attached to the vehicle body, will experi-
ence [ ]. The ground, observed from the vehicle through a hole in the floor, would be experienced
as moving with [ ].
It is suitable to decompose and in vehicle directions, since most equations (tyres, steering, propul-
sion, braking, etc) are expressed in those directions. With 𝑢
⃗ and 𝑢
⃗ as unit vectors in vehicle
280
Lateral Dynamics
directions: 𝑢
⃗ + ⃗ and
𝑢 𝑢
⃗ + 𝑢⃗ , or shorter, [ ] and [ ], which are
mathematical vectors. Now, it is important to remember that the vehicle is rotating with .
The accelerations are used in equilibrium equations, where the fictive force is . Since
we express in components in vehicle coordinate system directions, the differentiation is not as sim-
ple as differentiation component by component, [ ]. Instead: [ ] [ ]+
[ ∙ ∙ ] . The terms proportional to are centrifugal accelerations.
This can be mathematically shown as differentiation of the geometrical vector :
( 𝑢
⃗ + ⃗ ) (
𝑢 𝑢
⃗ + ⃗ )+(
𝑢 𝑢
⃗ + ⃗ )
𝑢
( 𝑢
⃗ + ⃗ )+(
𝑢 𝑢
⃗ ⃗ )
𝑢 (⏟ ) 𝑢
⃗ + (⏟ + ) 𝑢
⃗
𝑦
The constitutive and compatibility equations are typically expressed in velocities, not accelerations.
For high index models, some of the compatibility equations (a.k.a. constraint equations) also needs to
be differentiated, see example in 4.5.2.2.
4.4.2.3.2 Practical Explanation
Another explanation is given in left part of Figure 4-49 and the following reasoning. We can express
the velocity in direction of the x axis at time t, at two time instants:
• Velocity at time 𝑡:
• Velocity at time 𝑡 + 𝑡: ( + ) ∙ cos( ) ( + ) ∙ sin( ) ( ∙ cos( ) +
∙ cos( )) ( ∙ sin( ) + ∙ sin( )) ≈ { 𝑠 }≈( + )
( ∙ + ∙ )≈{ ∙ 𝑠 }≈( + ) ∙
Using these two expressions, we can express as the change of that speed per time unit:
{(𝑣 + 𝑣 ) 𝑣𝑦 ∙ 𝜓} 𝑣 𝑣 𝑣𝑦 ∙ 𝜓 𝑣 𝜓
• Acceleration=Velocity change per time ≈
∙ ∙
Corresponding for the lateral direction gives the Equation [4.45]. In Equation [4.45], the term ∙ is
generally more important than the term ∙ . This is because is generally much larger than .
Comparison with mass on
𝑦 𝑙 𝑙
𝑡 𝑒 𝑡 + Δ𝑡: +Δ circular path, with centrifugal
𝑥 acceleration
𝑦 Δ
Δ Δ
𝑡 𝑒 𝑡: 𝒂𝒄𝒆𝒏𝒕𝒓𝒊𝒑𝒆𝒕𝒂𝒍
𝑦
Δ
𝑥
2
𝑅 2
𝑛 𝑝 𝑙
𝑥 𝑙 𝑙 𝑅
𝑅
Figure 4-49: How to understand the acceleration term for vehicle motion in ground plane.
Left: Two consecutive time instants. Right: Comparison with circular motion, to identify the
centripetal acceleration.
4.4.2.3.3 Model with Velocity Components in Ground Fixed Directions
An alternative mathematical model is to express the velocity in components in ground fix (inertial) di-
rections ( 𝑢
⃗ + ⃗ , subscript for ground) instead of vehicle fix directions (
𝑢 𝑣
𝑢
⃗ 𝑣 + 𝑣 ⃗ 𝑣 , subscript for vehicle). The fictive forces are till
𝑢 , but we should express in
[ ] instead of [ 𝑣 𝑣 ], which becomes: ( 𝑢
⃗ + 𝑢
⃗ ) 𝑢
⃗ + 𝑢
⃗ .
281
Lateral Dynamics
velocities: accelerations:
𝑣
𝑣 𝑣 𝑣 ≠ 𝑣
𝑣 𝑣 + 𝑣 ≠ 𝑣
Figure 4-50: Velocity and Acceleration Components in Vehicle fix and Ground Fixed Directions.
Then, we replace the 2 translation equilibrium equations in Eq [4.47]. We also add 2 transformation
equations and a compatibility in yaw. The remaining equations can be kept, but it is proposed to add
subscript to and . The two models, side-by-side becomes as in Figure 4-51.
Model in… …vehicle fix directions, Eq [4.47] … ground fixed directions
Constitution 𝑓 𝑤 𝑓 ∙ 𝑠𝑓 𝑤 𝑣 ∙𝑠
Force transfor-
mation, steered 𝑓 𝑣 𝑓 𝑤 𝑓 𝑤 ∙ 𝑓 𝑓 𝑣 𝑓 𝑤 𝑓 + 𝑓 𝑤
wheels to vehicle
From Eq [4.46]: 𝑥
Path with orienta- 𝑥 𝑣 𝑜𝑠( ) 𝑣 𝑠 𝑛( )
𝑦
tion: 𝑦 𝑣 cos( )+ 𝑣 𝑠 𝑛( )
Figure 4-51: Comparison between model with velocities in vehicle fixed and ground fixed
directions.
Using the velocity components in inertial directions brings yaw angle into the differential equations
for velocities. So, we have to add as state: we get 4 states ( ) instead of 3 ( 𝑣 𝑣 ).
The need of transformation manifests that vehicle directions is more natural directions for the consti-
tutive equations (tyre models). However, the equations for path becomes simpler with the ground
fixed directions. The two models gives, of course, the same result, e.g. (𝑡) 𝑣 (𝑡) cos( (𝑡))
𝑣 (𝑡) sin( (𝑡)), but the model in vehicle fixed directions is more frequently used.
A similar comparison between model in vehicle and ground fix and components can be done for
pitch rotation in 3.4.5 and 5.7.2 but the difference there is less important since pitch angle is much
more limited than yaw angle.
282
Lateral Dynamics
(Note similar notation for vehicle yaw velocity, , and steering angular frequency, .) Knowing ̂ and
, it is possible to calculate the responses ̂ , ̂ , and :
̂ 𝑣𝑦 ̂ 𝜔𝑧
[ ] [ ] ∙ cos ( ∙ 𝑡 [ ]) 𝑛 [ ] [ ̂ ] ∙ cos ( ∙ 𝑡 [ ])
̂ 𝜔 𝑧 𝑦
𝑗∙ ∙ 𝓕 ([ ]) 𝑨 ∙ 𝓕 ([ ]) + 𝑩 ∙ ℱ( 𝑓 )
{
𝓕 ([ ]) 𝑪 ∙ 𝓕 ([ ]) + 𝑫 ∙ ℱ( 𝑓 )
Solving for states and outputs, using ℱ ( ) 𝑗∙ ℱ ( );, gives:
[ ] ℱ (𝓕 ([ ])) ℱ ((𝑗 ∙ ∙𝑰 𝑨) ∙ 𝑩 ∙ ℱ( 𝑓 ))
[ ] ℱ (𝓕 ([ ])) ℱ (𝑪 ∙ 𝓕 ([ ]) + 𝑫 ∙ ℱ( 𝑓 ))
{
Expressed as transfer functions:
(𝑗 ∙ ∙𝑰 𝑨) ∙ 𝑩 ∙ ℱ( 𝑓 )
𝑯 𝑣𝑦 𝓕 ([ ]) (𝑗 ∙ ∙𝑰 𝑨) ∙𝑩
𝛿𝑓 →[𝜔 ] ℱ( 𝑓 ) ℱ( 𝑓 )
𝑧
[4.51]
𝑯 𝜔𝑧 𝓕 ([ ]) 𝑪∙𝑯 𝑣𝑦 +𝑫 𝑪 ∙ (𝑗 ∙ ∙𝑰 𝑨) ∙𝑩+𝑫
𝛿𝑓 →[
𝑦
] ℱ( 𝑓 ) 𝛿𝑓→[𝜔 ]
𝑧
283
Lateral Dynamics
We have derived the transfer functions. The subscript tells that the transfer function is for the vehicle
operation with excitation=input 𝑓 and response=output [ ]𝑇 and output [ ]𝑇 . The
transfer function has dimension 2x1 and is complex. It operates as follows:
̂
[
] |𝐻 𝑣𝑦 | ∙ ̂𝑓
̂ 𝛿𝑓 →[𝜔 ]
𝑧
Amplitudes:
̂
[ ] |𝐻 𝜔𝑧 | ∙ ̂𝑓
{ ̂ 𝛿𝑓→[
𝑦
]
[4.52]
arg( ) )
{arg( 𝑓 }
𝑣𝑦
[ ] [ ] [ ] ∙ arg( 𝑓 ) arg (𝐻 𝑣𝑦 )
𝜔𝑧 arg( ) 0 𝛿𝑓 →[𝜔 ]
𝑧
Phase delays:
arg( )
{arg( 𝑓 )}
𝜔𝑧
[ ] [ ] [ ] ∙ arg( 𝑓 ) arg (𝐻 𝜔𝑧 )
{ 𝑦 arg( ) 0 𝛿𝑓→[
𝑦
]
[ [ ] ]
We intend to solve the complex equation, and then find the solutions as real parts: Re[ ] and
Re[ ] . (Subscript c means complex.)
If only interested in the stationary solution, which is valid after possible initial value dependent transi-
ents are damped out, we can assume a general form for the solution.
̂ ∙𝜔∙ ̂ ∙𝜔∙
[ ] [ ]∙𝑒 ⇒ [ ] 𝑗∙ ∙[ ]∙𝑒 [4.54]
̂ ̂
284
Lateral Dynamics
[ ]
̂
⇒[ ]
̂ [4.55]
𝑓 + 𝑓 ∙ 𝑓 ∙
+ ∙
0
[ ]∙𝑗∙ + 2 2 ∙[ ]∙ 𝑓 ∙ ̂𝑓
0 𝑓 ∙ 𝑓 ∙ 𝑓 ∙ 𝑓 + ∙ 𝑓
( [ ])
𝐻 𝑣𝑦 ∙ ̂𝑓
𝛿𝑓→[𝜔 ]
𝑧
Then, we can assume we know ̂ and ̂ from Equation [4.55], and consequently we know and
from Eq [4.54]. We have derived the transfer function, 𝐻 𝑣𝑦 . The subscript tells that the trans-
𝛿𝑓 →[𝜔 ]
𝑧
fer function is for the vehicle operation with excitation=input 𝑓 and response=output [ ] case.
This transfer function has dimension 2x1 and it is complex. It operates as follows:
̂ | |
Amplitudes: [ ] [ ] |𝐻 𝑣𝑦 | ∙ | 𝑓 |
̂ | | 𝛿𝑓 →[𝜔 ]
𝑧
[4.56]
arg( )
Phase delays: [ ] [ ] ∙ arg( 𝑓 ) {arg( 𝑓 ) 0} arg (𝐻 𝑣𝑦 )
arg( ) 𝛿𝑓→[𝜔 ]
𝑧
285
Lateral Dynamics
100 8
wz gain [(rad/s)/rad]
vy gain [(m/s)/rad]
80
6
60
4
40
20 2
0 0
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
vx=150 km/h
vx=100 km/h
2 2
vx=50 km/h
1 vx=1 km/h
1.5
wz delay [rad]
vy delay [rad]
0
1
-1
0.5
-2
-3 0
-4 -0.5
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
frequency [Hz] frequency [Hz]
Figure 4-52: Frequency response to steer angle. Vehicle data: 000 𝑘 000 𝑘
2
𝑓 . .5 𝑓 8 400 𝑁/𝑟 78000 𝑁/𝑟 (𝐾 . 6 𝑟 /𝑀𝑁).
286
Lateral Dynamics
LateralLateral
Velocity Frequency
Velocity Response Response Yaw Velocity Frequency
Yaw Rate ResponseResponse
80 8
Cf+Cr=119 kN/rad
Cf+Cr=159 kN/rad
wz gain [(rad/s)/rad]
vy gain [(m/s)/rad]
60 Cf+Cr=199 kN/rad 6
40 4
20 2
0 0
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
2 2
1
1.5
wz delay [rad]
vy delay [rad]
-1 1
-2
0.5
-3
-4 0
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
frequency [Hz] frequency [Hz]
Figure 4-53: Frequency response to steer angle. Same vehicle data as in Figure 4-52, except
varying 𝑓 and but keeping understeering gradient 𝐾 constant. Vehicle speed = 100 km/h.
LateralLateral
Velocity Frequency
Velocity ResponseResponse Yaw Velocity
Yaw Frequency Response
Rate Response
400 40
Ku= -1.26 rad/kN
wz gain [(rad/s)/rad]
300 30
Ku= 0.01 rad/kN
Ku= 1.26 rad/kN
200 20
100 10
0 -2 -1 0 1 2
0 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
2 2
1
1.5
wz delay [rad]
vy delay [rad]
-1 1
-2
0.5
-3
-4 -2 -1 0 1 2
0 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
frequency [Hz] frequency [Hz]
Figure 4-54: Frequency response to steer angle. Same vehicle data as in Figure 4-52, except
varying understeering gradient Ku but keeping 𝑓 + constant. Vehicle speed = 100 km/h.
287
Lateral Dynamics
Lateral acceleration Frequency Response is plotted for different vehicle speeds in Figure 4-55.
Lateral AccelerationLateral Acceleration FrequencyLateral
Response Response
Acceleration Response
250 2
vx=150 km/h
vx=100 km/h 1.5
200 vx=50 km/h
vx=1 km/h 1
ay gain [(m/(s*s))/rad]
0.5
150
ay delay [rad]
100
-0.5
-1
50
-1.5
0 -2
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10
frequency [Hz] frequency [Hz]
Figure 4-55: Lateral acceleration frequency response to steer angle. Vehicle data:
2
000 𝑘 000 𝑘 𝑓 . .5 𝑓 8 400 𝑁/𝑟
78000 𝑁/𝑟 (𝐾 . 6 𝑟 /𝑀𝑁).
§ Influence of Tyre Relaxation
If relaxation is taken into account using 0 and 𝐿 50% 𝑜𝑓 ℎ𝑒𝑒 𝑟 𝑢 𝑓𝑒𝑟𝑒𝑛 𝑒 0.5 𝜋 𝑅 𝑅
0. , we get the following versions of the diagrams in Figure 4-52 .. Figure 4-55:
288
Lateral Dynamics
We see that we get similar frequency responses, but the gain peaks at 𝑓 ≈ 0. . .08 𝐻 becomes a bit
higher. This is expected, since we add a delay in the system. The delay influence generally more if the
vehicle is articulated, such as long heavy combination vehicles.
It should be noted that the Lateral Acceleration Frequency Response gain approaches zero asymptoti-
cally when frequency increases, which was not the case without relaxation. The approach to zero is
truer and more intuitive since, with relaxation, the steered axle will not have time to develop forces
when its steering angle oscillates with high frequency.
4.4.3.4 Other Frequency Responses to Oscillating Steering *
In principle, it is possible to study a lot of other responses, such as Path Curvature Frequency Re-
sponse, Side slip Frequency Response and Lateral Path Width Frequency Response etc. For combina-
tion-vehicles it is common to plot Rearward amplification (RA) as function of frequency. The compen-
dium identifies some alternative definitions: 𝑅𝐴𝜔 max( ̂ ⁄ ̂ ) where numbers the vehicle
units. If the manoeuvre is not stationary oscillations, but e.g. a single lane change, we instead propose
𝑅𝐴𝜔 max (max| |⁄max| |) . RA can also be defined for the worst frequency, e.g. 𝑅𝐴𝜔
max (max( ̂ ⁄ ̂ )) . (Some argue for alternative definitions with instead of . However, these
𝑓
are more difficult to agree on; which point, vertical and longitudinal, one should measure in.)
289
Lateral Dynamics
290
Lateral Dynamics
lock rear
neutral-steered
under-steered
lock front
under-steered
neutral-steered
over-steered
lock rear
lock front
neutral-steered
Figure 4-56: Stability analysis of passenger car with varying cornering stiffnesses.
Eigenvalue analysis on matrix (𝑨 + 𝑩 ∙ [0 𝑘]) for varying speed and driver gain gives:
𝑓 5 stable for
𝐿
driver with
k=-0.05
𝑓
0
𝐿
291
Lateral Dynamics
292
Lateral Dynamics
[m]
𝑊 + 0. 5
𝑊 + 0. 5
𝑊 + 0. 5
1
.
.
.5
driving direction
.
15 30 25 25 15+15
Figure 4-58: Passenger cars - Test track for a severe lane-change manoeuvre - Part 1: Double lane-
change. Reference (ISO 3888).
. 𝑊 + 0. 5
[m]
1
driving direction
.
12 13.5 11 12.5 12
Figure 4-59: Passenger cars - Test track for a severe lane-change
manoeuvre - Part 2: Obstacle avoidance. Reference (ISO 3888).
Figure 4-60: Cone track for one standardized test of Over-speeding into curve
293
Lateral Dynamics
manoeuvres, and also active control such as ESC interventions. Hence, we extend the model in three
ways:
• The constitutive relation is saturated, to reflect that each axle may reach friction limit, friction
coefficient times normal load on the axle. See max functions in Equation [4.59].
• To be able to do mentioned limitation, the longitudinal load transfer is modelled, but only in
the simplest possible way using stiff suspension models. Basically, it is the same model as given
in Equation [3.13].
• A yaw moment representing (left/right) unsymmetrical braking/propulsion. See the term
𝑀 in yaw equilibrium in Equation [4.59]. This is much better modelled in a two-track
model, where 𝑀 does not appear, but we instead can have different longitudinal tyre forces
on left and right side.
It should be noted that the model lacks lateral load transfer and the transients in longitudinal load
transfer and the reduced cornering stiffness and reduced max friction due to load transfer and utiliz-
ing friction for wheel longitudinal forces.
Equilibrium in road plane (longitudinal, lateral, yaw):
∙( ∙ ) 𝑓 𝑤 ∙ cos( 𝑓 ) 𝑓 𝑤 ∙ sin( 𝑓 ) +
∙( + ∙ ) 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓 ) +
∙ ( 𝑓 𝑤 ∙ sin( 𝑓 ) + 𝑓 𝑤 ∙ cos( 𝑓 )) ∙ 𝑓 ∙ +𝑀
Equilibrium out of road plane (vertical, pitch):
𝑓 + ∙ 0
𝑓 ∙ 𝑓 + ∙ ( 𝑓 𝑤 ∙ cos( 𝑓 ) 𝑓 𝑤 ∙ sin( 𝑓 ) + )∙ℎ 0
Constitution:
[4.59]
𝑓 𝑤 sign(𝑠𝑓 ) ∙ min( 𝑓 ∙ |𝑠𝑓 | ∙ 𝑓 )
sign(𝑠 ) ∙ min( ∙ |𝑠 | ∙ )
Compatibility, slip definitions:
𝑓 𝑤 ∙
𝑠𝑓 𝑛 𝑠
𝑓 𝑤
Compatibility, transformation from vehicle to wheel coordinate system:
𝑓 𝑤 ( + 𝑓 ∙ ) ∙ sin( 𝑓 ) + ∙ cos( 𝑓 )
𝑓 𝑤 ( + 𝑓 ∙ ) ∙ cos( 𝑓 ) ∙ sin( 𝑓 )
A Modelica model is given in Eq [4.60]. Changes compared to Eq [4.46] are marked as underlined code.
//Equilibrium, in road plane:
m*(der(vx)-wz*vy) = Ffxv + Frx;
m*(der(vy)+wz*vx) = Ffyv + Fry;
J*der(wz) = Ffyv*lf - Fry*lr + Mactz;
//Equilibrium, out of road plane:
Ffz + Frz - m*g = 0;
-Ffz*lf + Frz*lr -(Ffxv + Frx)*h = 0;
//Compatibility:
vfxv = vx; vfyv = vy + lf*wz;
vrx = vx; vry = vy - lr*wz;
//Lateral tyre force model:
Ffyw = -sign(sfy)*min(Cf*abs(sfy), mu*Ffz); sfy = vfyw/vfxw; [4.60]
Fry = -sign(sry)*min(Cr*abs(sry), mu*Frz); sry = vry/vrx;
//Transformation between vehicle and wheel coordinate systems:
Ffxv = Ffxw*cos(df) - Ffyw*sin(df);
Ffyv = Ffxw*sin(df) + Ffyw*cos(df);
vfxv = vfxw*cos(df) - vfyw*sin(df);
vfyv = vfxw*sin(df) + vfyw*cos(df);
//Shaft torques
Ffxw = +1000; // Front axle driven.
Frx = -100; // Rolling resistance on rear axle.
Mactz=0;
A simulation of this model is seen in Figure 4-61. Same steering input as used in Figure 4-47. 𝑀 is
zero. Cornering stiffnesses are chosen so that the vehicle is understeered in steady state. Road friction
294
Lateral Dynamics
coefficient is 1. We can see that the vehicle now gets unstable and spins out with rear to the right. This
is mainly because longitudinal load transfer unloads the rear axle, since the kept steer angle deceler-
ates the vehicle. In this manoeuvre, it would have been reasonable to model also that the rear corner-
ing stiffness decreases with the decreased rear normal load, and opposite on front. Such addition to
the model would make the vehicle spin out even more. On the other hand, the longitudinal load shift is
modelled to take place immediately. With a suspension model, this load shift would require some
more time, which would calm down the spin-out. In conclusion, the manoeuvre is violent enough to
trigger a spin-out, so a further elaboration with how to control 𝑀 could be of interest. However, it
is beyond the scope of this compendium.
df=df [deg]
y [m]
wz [deg/s]
x[m]
yz=pz [deg]
time [s]
x[m]
Figure 4-61: Simulation results of one-track model for transient dynamics. The vehicle drawn in
the path plot is not in proper scale, but the orientation is approximately correct.
The vehicle reaches zero speed already after 7 seconds, because the wide side slip decelerates the ve-
hicle a lot. The simulation is then stopped, because the model cannot handle zero speed. That vehicle
models become singular at zero speed is very usual, since the slip definition becomes singular due to
zero speed in the denominator. The large difference compared to Figure 4-47 is due to the new consti-
tutive equation used, which shows the importance of checking validity region for any model one uses.
A simplified version of the mathematical model in Eq [4.59] follows in Eq [4.61]; assuming constant
and small angles and small propulsion force and no 𝑀 .
295
Lateral Dynamics
𝑓𝑦
𝑟𝑦
𝑟𝑥 df
𝑓 𝑣
𝐿
𝑓 𝑣
2 arctan 2 ⁄ 2 part of free body free body diagram part of free body
arctan ⁄ diagram of trailer: of coupling: diagram of trailer:
296
Lateral Dynamics
(𝐿 )
( ) ∎
Model equations for the 2nd unit:
Equilibrium of 2nd unit (longitudinal, lateral, yaw around CoG):
2 ( 2 2 2 ) 2 + 2
2 ( 2 + 2 2 ) 2 + 2
2 2 2 2 + 2 (2 2 )
Constitution for axles on 2nd unit:
[4.63]
2 2 𝑠2
Compatibility, within 2nd unit:
𝑠2 2 ⁄| 2 |
2 2 2 2
2 2 + 2 ( 2 2 ) ∎
Model equations for the coupling (see also1.6.1.1):
Equilibrium of coupling (x, y in 1st unit’s coordinates):
+ cos(θ) 2 + sin(θ) 2 0
sin(θ) 2 + cos(θ) 2 0
Compatibility of coupling: [4.64]
+ cos(θ) 2 + sin(θ) 2 ∎
sin(θ) 2 + cos(θ) 2 ∎
θ 2
𝑇
Let us assume [ 𝑓 𝑓 𝑤 ] are known inputs. Eqs [4.62].. [4.64] is a DAE system in with
2
24 equations and 24 unknowns: 2 2 2 θ 𝑓 𝑤 2 𝑠 𝑓 𝑠 𝑠2
2 2 𝑓 𝑤 𝑓 𝑤 𝑓 𝑣 2 2 . The equations can be written as they are into
a Modelica tool, which can find a possible state selection (such as 2 ), an explicit form
and perform simulations. Note that not only velocities but also an (relative angular) position becomes
states, as opposed to a one-unit vehicle. The use of a Modelica tool is very motivated for articulated ve-
hicles, since the articulation points makes the DAE system of equation to be of “high index”, meaning
that the Modelica tool identifies the constraints equations, i.e. the equations which have to be differen-
tiated, cf. 3.3.2. In this case, it is the equations marked with ∎ in Eqs [4.62]..[4.64].
4.5.2.2.1 § Explicit Form Model without DAE Tool
Contribution by Toheed Ghandriz, Vehicle Dynamics at Chalmers
It will now be explained how the explicit form model can be found using manual equation manipula-
tions or simpler symbolic tools, unable to handle DAEs (e.g. Matlab Symbolic Toolbox). We have to dif-
ferentiate the equations marked with ∎ in Eqs [4.62]..[4.64]. It leads to:
( )
2 2 + 2 (2 2 )
( sin(θ θ 2 + cos(θ) 2 ) + (cos(θ) θ 2 + sin(θ) 2 )
)
( cos(θ) θ 2 sin(θ) 2 ) + ( sin(θ) θ 2 + cos(θ) 2 )
So, we regard it as an ODE with 24+4=28 equations: . States and inputs as above DAE. The states are
regarded as known and we can count to 28 unknown variables, with notations as in 1.5.1.1.4.2:
{𝒙 𝒚𝟏 𝒚𝟐 𝒚𝟐 }
{ 2 }
{{ 𝑓 𝑤 2 𝑠 𝑓 𝑠 𝑠2 2 2 𝑓 𝑤 𝑓 𝑤 𝑓 𝑣 2 }}
{ 2 2 2 }{ 2 2 2 }
The expressions for state derivatives 2 θ and output variables (remaining of the 28 varia-
bles) can be found through algebraic manipulations, but in this case the symbolic tools will generate
297
Lateral Dynamics
huge expressions (hundreds of thousands of tokens) or even fail. Therefore, we will reformulate model
and introduce approximations in 4.5.2.2.1.1.
A way to keep symbolic expressions smaller (still large, but not huge) without approximations is to
stay at the implicit form, and solve it with special time integration methods, see 1.5.1.1.4.2:
𝒓𝒆 𝒊𝒅 𝒂𝒍 ← 𝒇𝒊 𝒑𝒍𝒊𝒄𝒊𝒕 (𝒙 𝒙 𝒚𝟏 𝒚𝟐 𝒚𝟐 (𝑡) 𝑡) ℎ𝑒𝑟𝑒 𝒚𝟐 { 𝑦 𝑦 𝑦 𝑥}
298
Lateral Dynamics
2 ( 2 2 (2 2 )) ( cos( ) + sin( ) 2 ( 2 2 ))
𝑡 𝑡
(( ( )) cos( ) + sin( ) 2 (2 2 ))
𝑡
( ( )) cos( ) ( ( )) sin( ) +
sin( ) + cos( ) 2 (2 2 )≈
≈( ( )) ( ( )) +
+ + 2 ( 2 2 )
Now the constitution can be involved:
Constitution for axles (to eliminate lateral forces 𝑓 𝑤 2 ):
𝑓 𝑤 𝑓 𝑠 𝑓 𝑤
𝑠 𝑤
2 2 𝑠2
We now eliminate the slips 𝑠 𝑓 𝑤 𝑠 𝑤 𝑠2 and express them in state variables:
Compatibility (to eliminate 𝑠𝑓 𝑤 ):
arctan ( 𝑓 𝑣 | |) 𝑓 + arctan(𝑠 𝑓 𝑤) ⇒𝑠 𝑓 𝑤 ≈ 𝑓 𝑣 ⁄| | 𝑓
where 𝑓 𝑣 +
Compatibility (to eliminate 𝑠 ):
𝑠 ⁄| |
where (𝐿 )
Compatibility (to eliminate 𝑠2 ):
𝑠2 2 ⁄| 2 |
where 2 2 2 2
where 2 cos( ) + sin( ) ≈ +
where ( )
and 2 + cos( ) sin( ) ≈ +
If we perform the eliminations we get 5 equations:
𝒇𝑫𝑨𝑬 (𝒙 𝒙 ) 𝟎 ℎ𝑒𝑟𝑒 𝒙 [ 𝑥 𝑦 ] [ 𝑓 𝑓𝑥 𝑟𝑥 𝑥]
Using the exact (grey text) expressions, this is exactly same model as in Eqs [4.62]..[4.64].
Symbolically Linearized Explicit Form Model
We can reformulate to explicit form: 𝒙 𝒇𝑶𝑫𝑬 (𝒙 ) but this is the same huge expression, mentioned
above. Using the approximate (black text) expressions, the solution will be somewhat smaller, but
still very large. However, we can linearize around an operation point (𝒙𝟎 𝟎 ):
𝒙 ≈ 𝒇𝑶𝑫𝑬 (𝒙𝟎 𝟎 ) + 𝑨(𝒙𝟎 𝟎 ) (𝒙 𝒙𝟎 ) + 𝑩(𝒙𝟎 𝟎 ) ( 𝟎)
𝒇𝟎 + 𝑨𝟎 (𝒙 𝒙𝟎 ) + 𝑩𝟎 ( 𝟎)
(𝒇𝟎 𝑨𝟎 𝒙𝟎 𝑩𝟎 𝟎 ) + 𝑨𝟎 𝒙 + 𝑩𝟎 𝒌𝒙 + 𝑨𝟎 𝒙 + 𝑩𝟎
𝜕𝒇𝑶𝑫𝑬 𝒊 𝜕𝒇𝑶𝑫𝑬 𝒊
The elements of the matrices 𝑨𝟎 𝑩𝟎 is: 𝐴 𝜕𝒙𝒋
| 𝑛 𝐵
𝜕 𝒋
|
𝒙𝟎 𝟎 𝒙𝟎 𝟎
As example of relevant operating condition to linearize around, we take straight-line without signifi-
cant longitudinal tyre forces: 𝒙𝟎 [ 𝑥0 0 0 0 0]𝑇 𝟎 [0 0 0 0]𝑇 . Note that we do not assume
constant speed, i.e. we do not set 0, but is still a state variable.
Then we get symbolic expressions of only 𝑨𝟎 𝑩𝟎 of around 15 thousand tokens, which is manageable,
but not worth writing out in this compendium, except on this overviewing way:
0 0 0 0 0 0 ≠0 ≠0 ≠0
0 0 ≠0 ≠0 ≠0 ≠0 0 0 0
𝑨𝟎 0 0 ≠0 ≠0 ≠0 𝑩𝟎 ≠0 0 0 0
0 0 ≠0 ≠0 ≠0 ≠0 0 0 0
[0 0 0 ] [ 0 0 0 0 ]
299
Lateral Dynamics
0 𝑓 [deg]
-20
0 1 2 3 4 5
21.5
[ ⁄𝑠]
21.0
20.5
20.0
0 1 2 3 4 5
veh_Exact.w _1z veh_Exact.w _2z veh_SmallAngleApprox.w _1z veh_SmallAngleApprox.w _2z veh_BothApprox.w _1z veh_BothApprox.w _2z veh_L
12
10
-2
-4
-6
-8
𝑡 [𝑠]
-10
0 1 2 3 4 5
Figure 4-63: Tractor and semitrailer. Comparison of simulations with the 4 models (thick solid:
”Exact”, thick dotted: “Small Angles”, thin dashed: “With All Physical Approximations, but not
linearized”, thin solid: “(Symbolically) Linearized”).
A similar linearization is done for a longer combination vehicle (A-double, i.e. Tractor+Semi-
trailer+Fulltrailer) in https://research.chalmers.se/publication/192958.
4.5.2.2.2 § Model Library for Articulated Vehicles
The first model (Eqs [4.62].. [4.64]) for articulated vehicle is written so that each equation belongs to
either a unit or a coupling. This opens up for systematic treatment of combination vehicles with more
than one articulation point. There are basically two conceptual ways:
• Modular library from which parts can be graphically dragged, dropped and connected.
• Vectorised formulation, see Reference (Sundström, Jacobson, & Laine, 2014) and 4.5.2.2.2.2.
§ Modular Library
A modular library is shown in figure below. Eq [4.62] is declared in “Unit” and Eq [4.64] in “Coupling”.
300
Lateral Dynamics
Figure 4-64: § Drag, drop and connect library for heavy combination vehicles. The model
example shows a so-called A-double, Tractor+SemiTrailer+Dolly+SemiTrailer.
An example of lane change manoeuvre, defined as lateral acceleration on 1 st axle follows a single sine-
wave, is shown in Figure 4-65. The natural input is prescribed steer angle ( 𝑓 sin(𝑡 𝑒) ), but since
modelled in Modelica, it is as easy to prescribe something else, e.g. lateral acceleration on first axle
( 𝑓 sin(𝑡 𝑒) ).
301
Lateral Dynamics
When modelling combination vehicles, there is a value of being able to use same model for vehicles
with different number of units. The following figure shows the physical model for a unit . . 𝑛𝑢,
where 𝑛𝑢 is the number of units. Only one axle 𝑗 . . 𝑛 , where 𝑛 is the maximum number of axles
per unit. There are 𝑛𝑢 couplings. The inputs to a model can be, e.g., 𝑛𝑢 𝑛 steering angles and
𝑛𝑢 𝑛 wheel-longitudinal forces 𝑤 . The latter can be actuated as 𝑤
(𝑇 𝑝 + 𝑇 𝑘 )⁄𝑅𝑤 𝑙 .
unit unit
𝑥 𝑀 𝑀
axle j 𝑀
ℎ
ℎ
Coupling Eqs:
+ 0
𝑀 +𝑀 0
𝑦
𝑥
unit
rear coupling front coupling rear coupling
on unit on unit on unit
Subscripts:
unit number, . . 𝑛𝑢
𝑗 axle number, . . 𝑛
𝑢 unit
axle
Coupling Eqs:
wheel
R rear coupling
F front coupling + 0
Figure 4-66: § Physical model of a combination vehicle. Only one unit and one coupling drawn.
Notation “⋯
⃗⃗⃗ ” means 2D geometric vector in 𝑥𝑦-plane.
The Modelica model becomes rather compact, using Modelica’s vector, for and if notation:
model vehModel_LateralOneTrack
… //Vectorised parameters, e.g. nu=number of units, na=max number of axles per unit, m[nu], …
… //Vectorised variables, e.g. v_x[nu], F_az[nu,na], …
equation
//For each unit:
for i in 1:nu loop
//In-road-plane Equilibrium of units (longitudinal, lateral, yaw around CoG):
if DynOpCond==3 or DynOpCond==2 then
KinEnForce_x[i] =m[i]*((if DynOpCond==3 then der(v_x[i]) else 0) - w_z[i]*v_y[i]);
KinEnForce_y[i] =m[i]*((if DynOpCond==3 then der(v_y[i]) else 0) + w_z[i]*v_x[i]);
KinEnMoment_z[i]=J_z[i]*der(w_z[i]);
Else // if DynOpCond==1 then
KinEnForce_x[i] =0;
KinEnForce_y[i] =0;
KinEnMoment_z[i]=0;
end if;
KinEnForce_x[i] = sum(F_axu[i,:]) +F_cRx[i] +F_cFx[i] +m[i]*g*sin(phi_ry)*sin(phi_z[i])
-(if i==1 then HalfRhoACd*v_x[i]^2 else 0);
KinEnForce_y[i] = sum(F_ayu[i,:]) +F_cRy[i] +F_cFy[i] -m[i]*g*sin(phi_ry)*cos(phi_z[i]);
KinEnMoment_z[i]= sum(F_ayu[i,:].*(l_a[i,:]-l_CoG[i]*ones(na)))
+F_cRy[i]*(l_cR[i]-l_CoG[i]) +F_cFy[i]*(l_cF[i]-l_CoG[i]);
a_x[i]=der(v_x[i]) - w_z[i]*v_y[i];
a_y[i]=der(v_y[i]) + w_z[i]*v_x[i];
//Out-of-road-plane Equilibrium of units (vertical, pitch around point on ground under CoG):
0 = +sum(F_az[i,:]) +F_cRz[i] +F_cFz[i] -m[i]*g*cos(phi_ry);
0 = -sum(F_az[i,:].*(l_a[i,:]-l_CoG[i]*ones(na))) -F_cRz[i]*(l_cR[i]-l_CoG[i])
-F_cFz[i]*(l_cF[i]-l_CoG[i]) +F_cRx[i]*h_cR[i] +F_cFx[i]*h_cF[i]
302
Lateral Dynamics
// if WhlTyreMod == 2 then
//Wheel, Rotational Equilibrium:
J_w*der(w[i, j]) = (F_aa[i, j] - RRC*F_az[i, j] - F_axw[i, j])*R_w;
//Tyre Constitution:
(F_axw[i, j],F_ayw[i, j]) = F_xy_RotatingWhl(F_z=max(F_az[i, j], F_eps), w=w[i, j],
v_xw=v_axw[i, j],v_yw=v_ayw[i, j],R_w=R_w,mu_peak=mu_peak);
s_x[i, j] = (Rw[i, j] - v_axw[i, j])/max(abs(Rw[i, j]), v_eps);
s_yt[i, j] = v_ayw[i, j]/max(abs(Rw[i, j]), v_eps);
alpha[i, j] = atan2(v_ayw[i, j], v_axw[i, j]);
Note that the tyre model used takes the influence of vertical tyre load into account. This is important
for heavy vehicles which typically have large variation of load and position of load. It requires that also
the out-of-road-plane equilibria, vertical and pitch, are included. It also requires that different coupling
303
Lateral Dynamics
degrees of freedom and axle lifting are modelled. Therefore, the coupling also have the parameters
“Boolean VertFree_RearCoupl[nu]”, “Boolean PitchFree_RearCoupl[nu]” and “In-
teger AxleGroupNumber[nu,na]”. Based on these parameters, and an assumption about the
suspension30 control, e.g. same vertical forces within an axle group, the vertical forces can be solved for.
Adouble_Dolly.veh.v_x[1]
Adouble_FullTrailer.veh.v_x[1]
It cannot be
20 guaranteed to work for all combination vehicles, but it works e.g. for the “A-double with
Dolly” and10 “A-double
0 1 with2 FullTrailer” 3 in4 Figure51-60. The
6 change
7 is only
8 to change
9 VertFree_Rear-
10
Adouble_Dolly.delta_f
Adouble_FullTrailer.delta_f
Coupl[2]0 from false to true and PitchFree_RearCoupl[3] from true to false. This
gives a change
0 due1 to that2 vertical3 force is 4 taken5 slightly
6 different,
7 see 8figure below.
9 10
Adouble_Dolly.veh.w_z[1]
Adouble_Dolly.veh.w_z[2]
Adouble_Dolly.veh.w_z[3]
Adouble_Dolly.veh.w_z[4]
Adouble_FullTrailer.veh.w_z[1]
Adouble_FullTrailer.veh.w_z[2]
Adouble_FullTrailer.veh.w_z[3]
Adouble_FullTrailer.veh.w_z[4]
0.6
3
0.4 2
4 solid=A-double with Dolly
dashed=A-double with FullTrailer
0.2
0.0
-0.2
-0.4
-0.6 𝑡 /[𝑠]
0 1 2 3 4 5 6 7 8 9 10
Figure 4-67: § Difference between “A-double with Dolly” and “A-double with FullTrailer”, see
Figure 1-60 . Single sine steering during < 𝑡 < at 0 /𝑠. The drawbar coupling carries no
vertical force in a FullTrailer, but some (here ≈ 𝑘𝑁) in the other case. Lateral acceleration,
peaking around 7. .8 /𝑠 2, is unrealistic for roll-over, but reducing to realistic . .4 /𝑠 2 gives
hardly any visible difference.
The difference does not become so large because the pitch moment of the dolly is taken by either the
pitch-rigid drawbar or the turntable of the 2nd SemiTrailer. The dynamic longitudinal load transfer is
considered by adding the term m[i]*der(v_x[i])*h[i] in pitch equilibrium of the units. This
generates more simultaneous equations (algebraic loops). So, neglecting the dynamic longitudinal load
transfer is an option for a simpler explicit form model.
304
Lateral Dynamics
ℎ
𝑓 𝑣
𝑓 𝑣
𝑓
𝑓 𝑓
𝑓 𝑣
𝑓 𝑣
Figure 4-68: Model of cambering vehicle. (The stars marks point of moment equilibria in the
“mathematical model” derived from this “physical model”.)
//Equilibrium in road plane (x,y,rotz):
m*ax = Ffxv + Frx;
m*ay = Ffyv + Fry;
Jz*der(wz) = Ffyv*lf - Fry*lr - m*ax*h*px;
ax = der(vx) - wz*vy;
ay = der(vy) + wz*vx;
//Equilibrium out of road plane (z, rotx, roty):
m*g = Ffz + Frz;
Jx*der(wx) = m*g*h*px + m*ay*h;
0 = -Ffz*lf + Frz*lr - m*ax*h; [4.65]
wx = der(px);
//Constitution:
Ffyw = -CC*Ffz*sfy; Fry = -CC*Frz*sry;
//Compatibility, slip definition:
atan(sfy) = atan2(vy + lf*wz + h*wx, vx - h*px*wz) - d;
atan(sry) = atan2(vy - lr*wz + h*wx, vx - h*px*wz) - 0;
//Force coordinate transformation:
Ffxv = -Ffyw*sin(d);
Ffyv = +Ffyw*cos(d);
When entering a constant radius curve from straight driving one has to first steer out of the curve to
tilt the vehicle a suitable amount for the coming path curvature, 𝑅𝑝 . The suitable amount is hence
𝑙 ⁄ ⁄ 2⁄
(𝑅𝑝 ) . Then one steers with the turn and balances
(closed loop controls) to the desired roll angle. Systems like this, which has to be operated in opposite
direction initially is called “Non-minimum phase systems”. It is generally difficult to design a controller
for such systems. The two simulations shown in Figure 4-85 are done with the above model. Initial
speed is 0 /𝑠 and 0. Path radius, 𝑅𝑝 (𝑡) 𝑠𝑡𝑒𝑝 𝑓𝑢𝑛 𝑡 𝑜𝑛 𝑡 𝑡 0. , representing a sud-
denly curving road or path.
• One simulation (veh_driver) uses a closed loop driver model which first steers outwards
( < 0) and then continuously calculates the steer angle as a closed loop controller: 𝑛
( ) . It is not claimed that the driver model is representative for real drivers.
• In the other simulation (veh_inverse), the roll angle is prescribed as 2 ⁄(𝑅
𝑝 ) .
To prescribe the roll angle, instead of steer angle, is a way to avoid the “Non-minimum phase”
difficulties. The system becomes a normal Minimum phase system if actuated with roll angle
instead of steer angle. A price one has to pay for this is that the model equations has to go
305
Lateral Dynamics
through a more advanced symbolic manipulation, e.g. differentiation, to solve for all variables,
including the steer angle . However, with a Modelica tool the symbolic manipulation is done
automatically. One can see this as a way to avoid a controller design and instead use an optimal
driver; optimal in the sense that it follows the path curvature with optimal yaw agility. The
road path curvature is a step function but has to be filtered twice (time constant 0.1 s is used in
Figure 4-85) to become differentiable. veh_driver.y
veh_inverse.y
10
The latter, optimal driver, negotiates the turn without overshot in yaw velocity, so it follows a sud-
denly curving path better. 8
veh_driver.d veh_inverse.d veh_driver.px veh_inverse.px
10
d= =steering angle [deg] 0
px= =roll angle [deg] 6
4
0 -10
2
-10 -20
0.0 0.4 0.8 1.2 0.0 1.6 0.4 2.0 0.8 1.2 1.6 2.0
0
veh_driver.Ffyv veh_inverse.Ffyv veh_driver.w z veh_inverse.w z
1000
1 -2
0
Ffyv= 𝑓 𝑣 = -4
0
vehicle-lateral force
-1000 on front tyre [N] wz= =yaw velocity [rad/s]
-6
-1
0.0 0.4 0.8 1.2 0.0 1.6 0.4 2.0 0.8 1.2 1.6 2.0 0
Figure 4-69: Simulations of entering a curve with a cambering vehicle. Blue solid curves without
dot-marker show a closed-loop driver model which actuates . Red curves with dot-marker shows
an optimal/inverse driver model which actuates .
Both driver models above only exemplify the low-level, roll-balancing, part of a driver. To run the
model in an environment with obstacles, one would also need a high-level, path selecting, part which
outputs desired, e.g., or . Additional driver model for longitudinal actuation is also needed.
It can be noted that the roll influences in two ways, compared to the non-cambering (roll-stiff) vehicles
previously modelled in the compendium:
• The roll motion itself is a dynamic motion, where the roll velocity becomes a state variable car-
rying kinetic energy.
• The roll influences the tyre slip, e.g. rear: 𝑠 ( +ℎ )⁄ ( ℎ ) . The
term ℎ can generally be neglected for non-cambering vehicles, but for a cambering vehicle,
such as a bike, it is essential. The term ℎ is only important at large roll angles, it is for
instance used as lever for longitudinal wheel forces in ESC-like control systems for motorbikes.
306
Lateral Dynamics
The free-body diagrams are given in Fig below, which should be compared to Figure 4-38.
Fsfl+Fsrl+ Fsfr+Fsrr+
zflr=zrlr=0
zfrr=zrrr=0 +Fdfl+Fdrl Pfz+Prz +Fdfr+Fdrr
Figure 4-70: § Model for transient load transfer due to lateral acceleration, using axle roll centres.
The constitutive equations for the compliances (or springs) are as follows, see Eq [4.31]. Note that the
anti-roll bars are not modelled. Note also that we differentiate, since we will later use the spring
forces as state variables in a simulation.
𝑓𝑙 𝑓𝑙 + 𝑓𝑤 ∙ ( 𝑓𝑙 𝑓𝑙 ) ⇒ 𝑓𝑙 𝑓𝑤 ∙ 𝑓𝑙 𝑓𝑤 ∙ 𝑓𝑙
𝑓 𝑓 + 𝑓𝑤 ∙ ( 𝑓 𝑓 ) ⇒ 𝑓 𝑓𝑤 ∙ 𝑓 𝑓𝑤 ∙ 𝑓
𝑙 𝑙 + 𝑤 ∙ ( 𝑙 𝑙 ) ⇒ 𝑙 𝑤 ∙ 𝑙 𝑤∙ 𝑙
+ 𝑤 ∙( ) ⇒ 𝑤 ∙ 𝑤 ∙
∙ ∙ ∙ ∙ 𝑓
ℎ𝑒𝑟𝑒 𝑓𝑙 𝑓 𝑛 𝑙
∙𝐿 ∙𝐿
The constitutive equations for the dampers have to be added:
𝑓𝑙 𝑓𝑤 ∙ 𝑓𝑙
𝑓 𝑓𝑤 ∙ 𝑓
𝑙 𝑤∙ 𝑙
𝑤 ∙
As comparable with Equation [4.32], we get the next equation to fulfil the equilibrium. The change
compared to Equation [4.32] is that we also have a roll and lateral inertia terms and 4 damper forces,
acting in parallel to each of the 4 spring forces. Actually, when setting up equations, we also under-
stand that a model for longitudinal load transfer is needed, which is why the simplest possible such,
which is the stiff suspension on in Equation [3.13].
In-road-plane: Equilibrium for vehicle (longitudinal, lateral, yaw):
∙ ∙( ∙ ) 𝑓 +
∙ ∙( + ∙ ) 𝑓 + [4.66]
∙ 𝑓 ∙ 𝑓 ∙
Out-of-road-plane: Equilibrium for vehicle (vertical, pitch, roll):
307
Lateral Dynamics
𝑓𝑙 + 𝑓 + + 𝑙 ∙ 0
( 𝑓𝑙 + 𝑓 ) ∙ 𝑓 + ( 𝑙 + )∙ ( 𝑓 + )∙ℎ 0
𝑤 𝑤
∙ ( 𝑓𝑙 + 𝑙 ) ∙ ( 𝑓 + )∙ +( 𝑓 + )∙ℎ
2 2
Equilibrium for each axle (pitch, around roll centre):
𝑤 𝑤
( 𝑓𝑙 ( 𝑓𝑙 + 𝑓𝑙 )) ∙ ( 𝑓 ( 𝑓 + 𝑓 )) ∙ + 𝑓 ∙ ℎ 𝑓 0
2 2
𝑤 𝑤
( 𝑙 ( 𝑙 + 𝑙 )) ∙2 ( ( + )) ∙ + ∙ℎ 0
2
Compatibility gives, keeping in mind that is the only non-zero out-of road plane velocity (i.e.
0):
𝑓𝑙 𝑓𝑙 + ∙ 𝑛 𝑓 𝑓 ∙
𝑙 𝑙 + ∙ 𝑛 ∙
Constitution for tyre forces versus slip in ground plane becomes per wheel, as opposed to per axle in
previous model:
𝑓 𝑓
𝑓 𝑤 sign(𝑠𝑓 ) ∙ (min ( ∙ |𝑠𝑓 | ∙ 𝑓𝑙 ) + min ( ∙ |𝑠𝑓 | ∙ 𝑓 ))
308
Lateral Dynamics
df=df [deg]
y [m]
wz [deg/s]
x[m]
yz=pz [deg]
x[m]
time [s]
Figure 4-71: § Simulation results of one-track model for transient dynamics with lateral load
transfer. The vehicle drawn in the path plot is not in proper scale, but the orientation is
approximately correct.
Even if the load transfer model does not influence the vehicle path a lot in this case, it may be im-
portant to include it to validity check the model through checking wheel lift. Wheel lift can be identi-
fied as negative vertical wheel forces, which are why we plot some vertical wheel forces, see Figure
4-72. In this case we see that we have no wheel lift (which would disqualify the simulation). In the
right part of the figure we can also see the separate contribution from spring ( ), damper ( ) and
linkage ( 𝑙 𝑛𝑘 𝑓𝑙 𝑓𝑙 𝑓𝑙 𝑓𝑙 ).
Frrz [N]
Fsrr [N]
Ffrz [N]
Frrz [N]
Flink,rr,z [N]
Fflz [N]
309
Lateral Dynamics
Figure 4-73: Top level of model with model tree structure. The Environment is a track test with
cone walls to go left and right around. Notation “irp” and “oorp” refers to in-road-plane and out-
of-road-plane, respectively.
As an initial overview, the states are presented. There are 21 states in total, and distributed:
• Driver: 0 states
• Vehicle:
o Vehicle Control & Actuation: 0 states
o Wheels, Tyres & Suspension: 12 states (4 wheels’ rotational speed, 4 Elastic parts of
vertical wheel forces, 4 Longitudinal tyre forces, 4 Lateral tyre forces)
o Vehicle Motion: 9 states (6 velocities and 3 positions)
• Environment: 0 states
The 4+4 tyre force states arise from modelled tyre relaxation, see 2.2.5.3.2.1.
310
Lateral Dynamics
9 states
12 states
311
Lateral Dynamics
propulsion system dynamics and control functions (ABS, ESC, TC, …). However, in this example
model it is only modelled very simple:
o Propulsion system outputs a fraction (determined by APed) of a certain maximum
power, distributed equally on front left and front right wheel. If brake pedal is applied,
the propulsion system outputs zero torque.
o Brake system outputs a fraction (determined by BPed) of a maximum brake torque (
/𝑅𝑤 ), distributed in a certain fix fraction between front and rear axle (70/30). The
distribution within each axle is equal on left and right wheel.
o There are no states modelled in the vehicle Control & Actuator submodel.
• Submodel “Wheels, Tyre & Suspension” models the part which pushes the tyres towards the
ground and consequently transforms the wheel torques and wheel steer angles, via the tyre, to
forces on the whole vehicle. F_xyv8 is the x and y forces in each of the 4 wheels, 2x4=8. F_z4
is the 4 vertical forces under each wheel. Submodel “Wheels, Tyre & Suspension” is further
explained in 4.5.3.2.5.
• Submodel “Vehicle Motion” models the motion of the whole vehicle in-road-plane and motion
of sprung body out-of-road-plane. The inertial effects (mass∙acceleration) of the unsprung
parts are considered for in-road-plane but not for out-of-road-plane. This submodel includes
integrators for the 3+3+3=9 states:
o Velocities in-road-plane, : v_irp3 (which is transformed to x- and y-veloci-
ties of each wheel and then fed back as v_irpv8)
o Position in-road-plane, 𝑥 𝑦 : pos_irp3 (which is only fed forward to “Environ-
ment”)
o Velocities out-of-road-plane, : v_oorp3 (which is transformed to z-velocities
of sprung body over each wheel and then fed back as v_z4)
4.5.3.2.4 Submodel Vehicle Control and Actuators
312
Lateral Dynamics
variables are present in this minimalistic example, but in a more advanced actuation model there
could typically be states such as: engine speed, gear (discrete state), delay states for brake system and
elastic forces in steering system.
4.5.3.2.5 Submodel Wheels, Tyres and Suspension
This submodel is shown in Figure 4-76.
• The 2 sub-models “Coord Transf to Vehicle” and “Coord Transf to Wheels” are straight-for-
ward coordinate transformations, see Eq [1.3].
• The sub-model “Wheels” is also relatively straight-forward. For each wheel, the rotational equi-
librium is used as model: ∙ 𝑇 +𝑇 ∙ 𝑅𝑤 sign( ) ∙ 𝑅𝑅 ∙ ∙ 𝑅𝑤 where 𝑅𝑤 is wheel
radius and 𝑅𝑅 is rolling resistance coeeficient. The submodel will hence contain the 4 states:
Rotational speeds of each wheel: w_w4.
• The sub-model “Springs, Dampers & Linkage” models the springs (incl. anti-roll-bars) and
dampers and the linkage. For each wheel:
o Four states: Elastic part of vertical tyre force under each wheel: F_s4
o The derivatives are governed by the differentiated constitution of the springs: Conceptually
but involving both wheel spring and anti-roll-bar.
o The force in damper is governed by the damper’s constitutive relation:
o The contact forces are calculated in submodel “Suspension Equilibrium” in Figure 4-77.
They are calculated from moment equilibrium of unsprung parts around a 3-dimensional
pivot axis. The pivot axis is defined by two points, the pivot point in longitudinal load trans-
fer (see Figure 3-31) and the pivot point in longitudinal load transfer (see Figure 4-37). The
scalar equilibrium equation for one wheel can be expressed, with vector (cross) and scalar
(dot) products, in 𝑣 𝑣 𝑇 and point coordinates. From this, can be solved. It
should be noted that the general relation should use a screw joint along the pivot axis, see
https://en.wikipedia.org/wiki/Screw_theory, which is why the 𝑒 of the screw appears in
equations.
313
Lateral Dynamics
o Unfortunately, the tyre forces and depend on . This could easily create algebraic
loops. However, since we also model relaxation, the tyre forces become state variables
which breaks such algebraic loops. Another way of getting rid of the algebraic loops could
have been to use “memory blocks”. “Memories” are such that value from last time instant is
used to calculate derivatives in present time instant. This is generally NOT a recommended
way of modelling.
There is one rotational equilibrium around pivot axis for the un- +
sprung part for each wheel (× means cross product, • means
scalar product): 𝑢 𝑓
𝑒 sin cos 0
𝑀𝑝 𝑣 𝑣 𝑣 • 𝑢𝑝 𝑣 shaft 𝑇
𝜋 axis
𝑟𝑙 𝑣 × 𝑣 𝑣 +
where 𝑀𝑝 𝑣 • 𝑢𝑝 𝑣
+𝑇 sin cos 0 𝑣 𝑣
where 𝑟𝑙 𝑣 𝑟𝑙 𝑟𝑙 𝑟𝑙
and 𝑢𝑝 𝑣 ⁄ 𝑟𝑝 𝑣 𝑟𝑝 𝑣 𝑟𝑝 𝑣 𝑟𝑝 𝑟𝑝 𝑟𝑝
If a linkage without screw effect on pivot axis, the 𝑒 is zero.
314
Lateral Dynamics
• state
30
0 7
6.5
5.5
tyre force in ground plane
(vertical tyre force is radius
-10 5
of circle)
4.5
4
wheel hub translational
3.5
velocity
3
cone
-20
2.5
-30
150 160 170 180 190 200 210 220
315
Lateral Dynamics
axle. In more detail: if < 𝑓 , we get 𝑛 𝑙 < , which means that will initially be directed
in same direction as 𝑓 , which means that rear axle rather helps than rotation. If smaller , the
the rear axle will help rather than .
316
Lateral Dynamics
Figure 4-81: Steering step response. Simulation with model from Equation [4.68].
4.5.4.1.1 Solution with Ansatz
Start from Equation [4.50] (assuming 𝑓 𝑤 0): [ ] 𝐀∙[ ]+𝐁∙ 𝑓
(0)
With steady state initial conditions: [ ] [ ] or [ ] 𝐀 ∙𝐁∙ 𝑓
(0)
̂ ̂ 𝜆1 ∙
] [ 2 ]] ∙ [𝑒 0 ]∙[ ] ⇒
∞
Make an ansatz: [ ] [ ] + [[
∞ ̂ ̂2 0 𝑒 𝜆2 ∙ 2
̂ ̂ 2 𝜆 ∙ 𝑒 𝜆1 ∙ 0
⇒ [ ] [[ ] [ ]] ∙ [ ]∙[ ]
̂ ̂2 0 𝜆2 ∙ 𝑒 𝜆2 ∙ 2
̂ ̂ 𝜆 ∙ 𝑒 𝜆1 ∙ 0
Insert: [[ ] [ 2 ]] ∙ [ ]∙[ ]
̂ ̂2 0 𝜆2 ∙ 𝑒 𝜆2 ∙ 2
̂ ̂ 𝜆1 ∙
] [ 2 ]] ∙ [𝑒 0 ] ∙ [ ]) + 𝐁 ∙
∞
𝐀 ∙ ([ ] + [[ 𝑓
∞ ̂ ̂2 0 𝑒 𝜆2 ∙ 2
317
Lateral Dynamics
̂ ̂ 2 𝜆1 ∙
]] ∙ [𝑒 0 ]∙[ ]
∞
[ ] [ ] + [[ ] [
{ ∞ ̂ ̂2 0 𝑒 2∙
𝜆 2
𝜆1 ∙
+ ∙ 𝜆 ∙ ̂ ∙𝑒 ∙ + 𝜆2 ∙ ̂ 2 ∙ 𝑒 𝜆2 ∙ ∙ 2 + ∙
∞ ̂ ̂ 𝜆 0 [4.68]
where: [ ] 𝑨 ∙𝑩∙ 𝑓 and [[[ ] [ 2 ]] [ ]] eig(𝑨)
∞ ̂ ̂2 0 𝜆2
̂ ̂ ∞
and [ ] [[ ] [ 2 ]] ∙ ([ ] [ ])
2 ̂ ̂2 ∞
Another way to express or compute this is the “matrix exponential”, mentioned in 1.5.1.1.5.
4.5.4.1.2 Solution with Laplace transform
By Mats Jonasson, Volvo Cars and Vehicle Dynamics at Chalmers
⇒ 𝑠 ℒ ([ ]) 𝟎 𝐀 ∙ ℒ ([ ]) + 𝐁 ∙ ℒ ( 𝑓 (𝑡)) ⇒
𝑓 𝑯(𝑠)
⇒ {ℒ(𝑠𝑡𝑒𝑝) 𝑠𝑡𝑒𝑝⁄𝑠} ⇒ ℒ ([ ]) ⏟(𝐀 𝑠 𝑰) 𝐁∙ ⇒ ℒ ([ ]) ∙ 𝑓 ⇒
𝑠 𝑠
𝑯( )
𝑯(𝑠)
⇒ [ ] ℒ ( ∙ 𝑓) ⇒
𝑠
inverse Laplace transform
𝑯( ) cos( 6 𝑡 + 7 )
⇒ [ ] ℒ ( )∙ 𝑓 ⋯{ e.g. using Matlab command } [ ] [ 3 ] 𝑒 5 [ ]
2 4 cos( 6 𝑡 + 8 )
ilaplace see .5.4.
where 2 ⋯ 8 are real parameters, expressed in model parameters 𝑓 𝑓 , initial condit-
ions and size of step in input variable 𝑓 . So, (𝑡 and
) (𝑡 can be computed and plotted, overlapping
)
in Figure 4-81. Note:
• The 𝑯(𝑠) is the Laplace transfer function. Each component of 𝑯(𝑠) is a surface over the com-
plex space [𝑅𝑒(𝑠) 𝐼 (𝑠)]. Compare with Fourier transfer function 𝑯( ) of which each compo-
nent is a one-dimensional curve over .
• Laplace transform can also handle other transient vehicle problems, such as engine torque step
or vertical road step. Laplace can also handle other tranisent disturbances, such as ramps.
• For stationary oscillations, Fourier transform is often enough and easier to use.
318
df sfy sry
0.0
Lateral Dynamics
-0.2
0 1 2 3 4 5
Ffyw Fry
8000
vx
28
4000
26 vx[m/s]
0 1 2 3 4 5 0
0 1 2 3 4 5
df sfy sry
df[rad]
0.0 beta_deg wz_dps ay
15
sry[1] wz[deg/s]
-0.2 sfy[1] 10
0 1 2 3 4 5
ay[m/(s*s)]
5
Ffyw Fry
8000 Ffyw[N]
0
4000 Fry[N]
-5 beta[deg]
0
-10
0 1 2 3
time[s] 4 5 0 1 2 time[s]
3 4 5
Function definition:
5 Steering effort at high speed is the steering wheel torque (or subjectively assessed effort)
needed to perform
0
a certain avoidance manoeuvre at high road friction.
At higher vehicle speeds, the steering effort is normally less of a problem since unless really high steer-
ing wheel rate. Hence,
-5
steering wheel torque in avoidance manoeuvres in e.g. 70 km/h can be a rele-
vant requirement.-10
In these situations, the subjective assessment of steering effort can also be the
0 1 2 3 4 5
measure. Then, steering effort is probably assessed based on both steering wheel rate and steering
wheel torque.
319
Lateral Dynamics
small negative 𝑠𝑓 𝑠𝑓
large negative 𝑠
⁄𝑠
/𝑟
understeering
𝑠
(transient)
large negative 𝑠𝑓
Figure 4-83: Phase portrait for constant and constant steer angle 𝑓 . Bottom: Using simple
model in Eq [4.50]. Only black solid trajectories credible, since they are completely within model
validity. Upper: Using a model with larger validity. From Mats Jonasson, VCC.
320
Lateral Dynamics
Figure 4-84: Rearward amplification, P is peak value of motion variable of interest. From
(Kharrazi , 2012).
§ An alternative definition of RA is via the frequency response: 𝑅𝐴 |𝐻𝛿→𝜔1 ⁄𝐻𝛿→𝜔 |, where 𝐻 de-
nots transfer functions, so 𝐻 (𝑓), where 𝑓 is the frequency of an harmonic variation of . So, either we
see 𝑅𝐴 𝑅𝐴(𝑓) or 𝑅𝐴 max(𝑅𝐴(𝑓)). The 𝑓 where max appears is typically 0.4..0.6 Hz. The frequency
f
response definition is better (independent of amplitude) for measuring how difficult it is for driver
to keep the lateral stability of the last unit when driving straight ahead with small lateral disturbances.
It is a relevant measure both for human driver and virtual driver, such as LKA, 4.6.2.3.
321
Lateral Dynamics
𝐷 2
Figure 4-85: Yaw damping, 𝐷 denotes damping ratio of the articulation joint. From (Kharrazi ,
2012).
322
Lateral Dynamics
Desired yaw velocity and side-slip is calculated using a reference model and a closed loop control on
the reference, see Figure 4-86. The reference model requires at least steer angle and longitudinal ve-
locity as input. The reference model can be either of steady state type (approximately as Eq [4.20] or
the [4.15]) or transient (approximately as Eq [4.50]). The vehicle modelled by the reference model
should rather be a desired vehicle than the controlled vehicle. Figure 4-86 does NOT show: Slide slip
control, Reduction of 𝑓 due to low friction detection, Coordination with Engine/Steering interven-
tions, Arbitration with Pedal/ACC/ABS braking. (For single unit vehicles, can be seen as a yaw
moment.)
or transient: ClosedLoop
Brake
+ Controller, 𝑇𝑠 Brake
𝑓 Coordinator 𝑇𝑠 Actuator
𝑓 conceptually: Brake
Actuator Controlled
𝑓 𝑓 (select wheel) 𝑇𝑠
Actuator Vehicle
𝑓 𝑓 𝑛 𝑠𝑓
𝑡 𝑛
𝑛 𝑠
𝑠𝑓 + 𝑓 ⁄ 𝑓
𝑠 ⁄
Environ-ent
323
Lateral Dynamics
Figure 4-87: ESC brake interventions when oversteer and understeer, on a tractor with trailer.
4.6.2.1.3 Over-Speed Control
Over-speed control is not always recognised as a separate concept, but as a part of under-steer control.
The actuation is that propulsion is reduced, or more than just inner rear wheels are braked. In this
text, we identify this as done to decrease speed, which has a positive effect later in the curve.
4.6.2.1.4 Wheel-Level Control
A pre-requisite for all controls mentioned above in 4.6.2.1 is that the wheel torque actuator primarily
responds to a torque request. However, one need to have another request channel to adjust the lateral
force margin; normally one uses a longitudinal slip request, 𝑠 in Figure 4-86 and Figure 4-88. The
slip request is generally used as a “safety net” to avoid lock-up the wheel too much; so, it is a “max |𝑠 |
request”. Typically, 𝑠 is 0. . . 0. , for braking, but at RSC interventions (see 4.6.2.2) the lateral grip
should be braked away, so a deeper slip request is then used, typically 50-70%.
TorqueRequest,
𝑇 Torque
controller
ESC & RSC piston ℎ ℎ
Actuator 𝑭𝒂𝒙 caliper
controllers Min 𝑭𝒂𝒙
𝑇𝑠
Control
𝑜 (or large 𝑠 )
Closed loop Wheel rotational
brake pads
wheel
attaches
LongSlipRequest,
𝑠 slip controller speed sensor rotor=disk
here
𝑠
hub 𝑠
vehicle state Calculate
estimator slip normal request RSC request
𝑠 0. . . 0. % 𝑠 0.6. . 0.5%
”Smart actuator”
Figure 4-88: Individual wheel control by friction brakes for ESC-type functions. What is a “smart
actuator” can depend on which function architecture that vehicle manufacturer and brake
supplier has agreed.
4.6.2.1.5 Other Intervention than Individual Wheel Brakes
Balancing with Propulsion per Axle
For vehicles with controllable distribution of propulsion torque between the axles, ESC can intervene
also with a request for redistribution of the propulsion torque. If over-steering, the propulsion should
be redistributed towards front and opposite for understeering.
Torque Vectoring
For vehicles with controllable distribution of propulsion torque between the left and right, ESC can in-
tervene also with a request for redistribution of the propulsion torque. If over-steering, the propulsion
should be redistributed towards inner side and opposite for understeering.
Steering Guidance
For vehicles with controllable steering wheel torque, ESC can intervene also with a request for addi-
tional steering wheel torque. The most obvious function is to guide driver to open up steering (coun-
ter-steer) when the vehicle over-steers. Such functions are on market in passenger cars today. Less ob-
vious is how to guide the driver when vehicle is under-steering.
4.6.2.1.6 ESC using Environment Information / ESC for the Virtual Driver
A prognosis of the future development of ESC like functions is that environment sensors can be used to
better predict what driver tries to do; presently ESC can only look at steering wheel angle.
Related to this, but still somewhat different, would be to utilize the automated driving development by
utilizing that a “virtual driver” can be much better predicted than a “manual driver”. So, a predictive
ESC control is more possible.
324
Lateral Dynamics
Two concepts:
• Keep in centre of lane
• Intervene when about
to leave lane
lateral position
in lane
Figure 4-89: Two concepts for Lane Keeping Aid.
325
Lateral Dynamics
context. In a way, AD is already reality since there are vehicles on the road which can have ACC and
LKA active at the same time. On the other hand, AD can be seen as very futuristic, since completely
driverless vehicle which can operate in all situations is far from mass-production.
It is not obvious if AD will mean higher or lower requirements on vehicle dynamics. Some (of many
more!) examples of changes, relevant for vehicle dynamics are:
• The vehicle control can better predict the next few seconds of a virtual driver (AD algorithms)
than of a (human) driver. This can facilitate loss-of-grip functions, such as ABS & ESC.
• There will be new requirements on vehicle response on requests from the virtual driver, in
parallel with requirements on response on human drivers pedal and steering wheel operation.
• There will be new requirements on vehicle relative motion, relative to surrounding road and
traffic, such as lane edges and other road users. These will be in synergy or conflict to corre-
sponding requirement for absolute motion response.
• The motion actuation will have to be more redundant, since driver is less likely to take back
control quickly. Emergency functions to reach safe stop will need to work with partly faulty
sensors and actuators. Failures needs to be designed according to (ISO 26262, 2011-2012)
• The maximum speed for which the vehicle is designed can possibly be lower, since reduced
transport efficiency could be accepted if driver can do something else or is not needed at all.
• Estimation of Road friction, Controlling to Safe stop, Self-Diagnose, etc.
326
Vertical Dynamics
5 VERTICAL DYNAMICS
5.1 Introduction
The vertical dynamics are needed since vehicles are operated on real roads, and real roads are not per-
fectly smooth. Also, vehicle can be operated off-road, where the ground unevenness is even larger.
The irregularities of the road can be categorized. A transient disturbance, such as a pothole or bump,
can be represented as a step input or ramp. Undulating surfaces like grooves across the road may be a
type of sinusoidal or other stationary oscillating (or periodic) input. More natural input like the ran-
dom surface texture of the road may be a random noise distribution. In all cases, the same mechanical
system must react when the vehicle travels over the road at varying speeds including doing manoeu-
vres in longitudinal and lateral directions.
The chapter is organised around the 3 complete vehicle functions: 5.5.1 Ride Comfort *, 5.5.2 Fatigue
Life *, and 5.5.3 Road Grip *. It is, to a larger extent than Chapters 3 and 4, organised with mathemati-
cal theory first followed by the vehicle functions. In Figure 5-1 shows the 3 main functions. It explains
the importance of the vehicle’s dynamic structure. The vehicle’s dynamic structure calls for a pretty
extensive theory base, described mainly in 5.2.
𝒛̈ Human
Vibrations of Ride
perception of
human
Road surface
Vehicle’s vibrations Comfort
𝒛 irregularities 𝒛𝒓 𝒛
speed Vehicle’s Stresses in 𝒛 𝒛
𝒓 Material Fatigue
𝒛𝒓 dynamic vehicle
Vehicle structure structure
fatigue Life
𝑭𝒓𝒛 created Contact
disturbances Compression 𝑭𝒓𝒛 Road
between tyres
of tyre Grip
and road
Figure 5-1: Different types of knowledge and functions in the area of vertical vehicle dynamics,
organised around the vehicle’s dynamic structure.
Models in this chapter focus the disturbance from vertical irregularities from the road, i.e. only the ver-
tical forces on the tyre from the road and not the forces in road plane. This enables the use of simple
models which are independent of exact wheel and axle suspension, such as pivot axes and roll centres.
Only the wheel stiffness rate (effective stiffness) and wheel damping rate (effective damping), see Fig-
ure 2-61, influence. This has the benefit that the chapter becomes relatively independent of previous
chapters, but it has the drawback that the presented models are not really suitable for studies of steep
road irregularities (which have longitudinal components) and sudden changes in wheel torque or tyre
side forces. Also, noise (>≈ 5𝐻 ) is not covered in this compendium.
327
Vertical Dynamics
[5.1]
𝑁
𝑀𝑢 𝑡 𝑝 𝑒 𝑓𝑟𝑒𝑞𝑢𝑒𝑛 𝑒𝑠 ∶ (𝜉 ) ∑ ̂ ∙ cos( ∙𝜉+ )
=
ℎ𝑒𝑟𝑒 𝜉 𝑠 𝑡ℎ𝑒 𝑛 𝑒𝑝𝑒𝑛 𝑒𝑛𝑡 𝑟 𝑒.
m
zz1 zz1 zz1 zz1
c
t t t t
F1
Step, as one example of Harmonic None-harmonic Random noise
transient (none- stationary stationary oscillation,
stationary oscillating) oscillation, single single frequency (or multiple
frequency random-frequency
other examples can be (or multiple-frequency harmonic
step down, square pulse, harmonic stationary stationary
ramp etc. oscillation) oscillation)
Figure 5-2: Different types of variables, both transient and stationary oscillating. The independent
variable 𝜉 can, typically, be either time or distance.
The most intuitive is probably to think of time as the independent variable, i.e. that the variation takes
place as function of time and that 𝜉 𝑡 in Equation [5.1]. However, for one specific road, the vertical
displacement varies with longitudinal position, rather than with time. This is why we can either do
analysis in time domain (𝜉 𝑡) and space domain (𝜉 𝑥).
Since the same oscillation can be described either as a function of 𝜉 ( (𝜉 )) or as a function of fre-
quency ( ̂ ̂ ( ) ), we can do analysis either in the independent variable domain (𝜉) or in fre-
quency domain ( ).
The four combinations of domains are shown in Figure 5-3.
𝒙 𝒗𝒙 𝒕 𝒇 𝒇⁄𝒗𝒙
𝝎 𝝎⁄𝒗𝒙
328
Vertical Dynamics
∙𝜋∙𝑓
[5.2]
ℎ𝑒𝑟𝑒 [𝑟 ⁄𝑠 ] 𝑛 𝑢 𝑟 (𝑡 𝑒) 𝑓𝑟𝑒𝑞𝑢𝑒𝑛 𝑦
𝑛 𝑓[ 𝑠 ⁄ 𝑜𝑠 𝑡 𝑜𝑛𝑠⁄𝑠] (𝑡 𝑒) 𝑓𝑟𝑒𝑞𝑢𝑒𝑛 𝑦
The time for one oscillation is called the period time. It is denoted 𝑇:
[5.3]
𝑇 ⁄𝑓 ∙ 𝜋⁄
[5.4]
𝑡 𝑛
∫ 2 ∙ 𝑡
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆( ) √
𝑡 𝑛
[5.5]
] = 2
̂ ∙ + 4∙ ̂ ∙( + 4∙ ) ̂2
→
𝑡𝑛 𝑡𝑛 →∞
∫ 2 ∙ 𝑡 | ̂|
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆( ) √ √𝑀𝑆( )
𝑡 𝑛 √
If the variable is written as a multiple frequency harmonic stationary oscillation:
𝑁 𝑁
𝑉 𝑟 𝑒: ∑ ∑ ̂ ∙ cos( ∙𝑡+ )
= =
2
∫ 2
∙ 𝑡 ∫ (∑𝑁= ) ∙ 𝑡
𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑀𝑆 ( )
𝑡 𝑛 𝑡 𝑛
2
∫ (∑𝑁= ̂ ∙ cos( ∙𝑡+ )) ∙ 𝑡
⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑡𝑛 →
[5.6]
𝑡 𝑛
2 𝑁 𝑁
∫ ∑𝑁= ̂ ∙ (cos( ∙𝑡+ ))2 ∙ 𝑡 ̂ 2
⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑡𝑛 → ∑ 𝑀𝑆( ) ∑
𝑡 𝑛
= =
𝑁 2 𝑁 𝑁
̂ 2
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆( ) √𝑀𝑆( ) √∑ √∑ 𝑀𝑆 ( ) √∑(𝑅𝑀𝑆( ))
= = =
329
Vertical Dynamics
be thought of as integrals of a “continuous amplitude curve”, ̂ , where the integration is done over a
small frequency interval, centred around a mid-frequency, :
𝜔𝑖+𝜔𝑖+1
2
̂
[5.7]
+
̂ ∫ ̂ ∙ ̂ ( )∙ ̂ ( )∙ ⇒ ̂ ( )
𝜔 +𝜔𝑖
𝜔= 𝑖 1
2
We realize that the unit of ̂ has to be same as for , but per [rad/s]. So, if z is a displacement in [m], ̂
has the unit [m/(rad/s)]. Now, ̂ is a way to understand the concept of a spectral density. A similar
value, but more used, is the Power Spectral Density, PSD (also called Mean Square Spectral Density).
𝑆𝐷( ) which is a continuous function, while ̂ is a discrete function. That means that 𝑆𝐷( ) is fully
determined by a certain measured or calculated variable (𝑡), while ̂ depends on which discretiza-
tion (which or which ) that is chosen.
[5.8]
𝑆𝐷( (𝑡) ) ( )
ℎ𝑒𝑟𝑒 filter 𝑠 𝑛 𝑝 𝑠𝑠 𝑓 𝑡𝑒𝑟 𝑒𝑛𝑡𝑒𝑟𝑒 𝑟𝑜𝑢𝑛 𝑛 𝑡ℎ 𝑛 𝑡ℎ
PSD can also be defined with band width in time frequency instead of angular frequency. Eq [5.8] is the
same but replacing with 𝑓.
When the variable to study (z) is known and the band width is known, one often writes simply 𝑆𝐷( )
or ( ). G has the same unit as 2 , but per [rad/s] or per [oscillations/s]. So, if z is a displacement in
[m], G has the unit [ 2⁄(𝑟 ⁄𝑠)] or [ 2⁄( ⁄𝑠) 2
∙ 𝑠].
RMS is square root of the area under the PSD curve:
𝑁 𝑁 ∞
[5.9]
𝑅𝑀𝑆 ( ) √∑ 𝑀𝑆( ) √∑ ( )∙ √ ∫ ( )∙
= = 𝜔=
( ) 2
∙ ( ) [5.10]
∞
∙𝜔∙
ℎ𝑒𝑟𝑒 ℱ 𝑠 𝑡ℎ𝑒 𝑜𝑢𝑟 𝑒𝑟 𝑜𝑝𝑒𝑟 𝑡𝑜𝑟: 𝑍( ) ℱ( (𝑡)) ∫𝑒 ∙ (𝑡) ∙ 𝑡
330
Vertical Dynamics
Since there can be different excitations and responses in a system, there are several transfer functions.
To distinguish between those, a subscripting of 𝐻 is often used: 𝐻 𝑛→ 𝑝 𝑛 , which would be
𝐻 𝑟→ 𝐻 𝑝𝑙 𝑚 𝑛 → 𝑝 𝑛 𝑚 𝑝𝑙 𝑚 𝑛 in the example above. Other examples of relevant
transfer functions in vertical vehicle dynamics are:
• 𝐻 𝑝𝑙 𝑚 𝑛 → 𝑝 𝑛 𝑚 𝑙 [( ⁄𝑠 2 )⁄ ], see 5.5
𝑛
• 𝐻 𝑝𝑙 𝑚 𝑛 → 𝑝 𝑛 𝑛 𝑓 𝑚 𝑛 [ ⁄ ] , see 5.5.2
• 𝐻 𝑝𝑙 𝑚 𝑛 → 𝑓 [𝑁 ] , see 5.5.2
⁄
When transfer function for one derivative is found, it is often easy to convert it to another:
𝐻 1→ 𝑗∙ ∙ 𝐻 1→ 2
[5.13]
2
2
𝐻 1→ ̈2 𝑗∙ ∙ 𝑗 ∙ ∙ 𝐻 1→ 2 ∙𝐻 1→ 2
𝐻 1→ 2 3
𝐻 1→ 2 𝐻 1 → 3
The usage of the transfer function is, primarily, to easily obtain the response from the excitation, as
shown in Equation [5.12]. Also, the transfer function can operate on the Power Spectral Density,
PSD=G, as shown in the following:
2 2
𝑀𝑆( (𝑡) ) ( ̂ ( )) ⁄ (|𝐻( )| ∙ ̂ ( )) ⁄
( )
[5.14]
2
2 ( ̂ ( )) ⁄ 2
|𝐻 𝑟 → ( )| ∙ |𝐻 𝑟 → ( )| ∙ 𝑟
( )
Using Equation [5.9], we can then express 𝑅𝑀𝑆( ) (sprung mass), from knowing 𝑟
( ) (road):
[5.15]
∞ 2
𝑅𝑀𝑆( ) √∫𝜔= |𝐻 𝑟 → ( )| ∙ 𝑟
( )∙
[5.16]
𝑥 ∙𝑡+𝑥
The offset (𝑥 ) is the phase (spatial) offset (𝑥 ) is the correspondence to the phase angle ( ).
The corresponding formulas as given in Equations [5.2]..[5.13] can be formulated when changing to
space domain, or spatial domain. It is generally a good idea to use a separate set of notations for the
spatial domain. Hence the formulas are repeated with new notations, which is basically what will be
done in present section.
In space domain, the frequency has the common understanding of “how often per distance”. Even so,
there are two relevant ways to measure frequency: spatial angular frequency and spatial frequency.
∙𝜋∙𝑓
[5.17]
ℎ𝑒𝑟𝑒 [𝑟 ⁄ ] 𝑛 𝑢 𝑟 𝑠𝑝 𝑡 𝑓𝑟𝑒𝑞𝑢𝑒𝑛 𝑦
𝑛 𝑓 [ ⁄ 𝑜𝑠 𝑡 𝑜𝑛𝑠⁄ ] 𝑠𝑝 𝑡 𝑓𝑟𝑒𝑞𝑢𝑒𝑛 𝑦
The correspondence to period time is wavelength, denoted 𝜆:
[5.18]
𝜆[ ] ⁄𝑓 ∙ 𝜋⁄
Now, the basic assumption in Equation [5.16] and definitions of frequencies gives:
331
Vertical Dynamics
[5.19]
∙ 𝑛 𝑓 ∙𝑓
The relation between the phase (spatial) offset (𝑥 ) and the phase angle ( ) is:
[5.20]
𝜆∙
𝑥
∙𝜋
[5.21]
𝑥 𝑛
domain. We subscript these with s for
space. ∫ 2 ∙ 𝑥
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆 ( ) √
𝑥 𝑛
Because 𝑥 is constant, the Mean Square and Root Mean Square will be the same in time and space
domain. If the variable is written as a single frequency harmonic stationary oscillation, these values
becomes as follows:
𝑉 𝑟 𝑒: ̂ ∙ cos( ∙ 𝑥 + 𝑥 )
̂2
[5.22]
𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑀𝑆 ( ) ⋯ 𝑀𝑆( )
| ̂|
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆 ( ) ⋯ 𝑅𝑀𝑆( )
√
If the variable is written as a multiple frequency harmonic stationary oscillation:
𝑁 𝑁
𝑉 𝑟 𝑒: ∑ ∑ ̂ ∙ cos( ∙𝑥+𝑥 )
= =
𝑁 2
̂
[5.23]
𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑀𝑆 ( ) ∑ 𝑀𝑆 ( )
=
𝑁
2
𝑅𝑜𝑜𝑡𝑀𝑒 𝑛𝑆𝑞𝑢 𝑟𝑒: 𝑅𝑀𝑆 ( ) √𝑀𝑆 ( ) √∑(𝑅𝑀𝑆( )) 𝑅𝑀𝑆 ( )
=
𝑆𝐷 ( (𝑥) 𝜆) 𝜆
Φ( )
where "filter" is a band pass filter centred around ω and with band width 𝑓
When the variable to study z is known and the band width is known, one often writes simply 𝑆𝐷 ( )
or Φ( ). The Φ has the same unit as 2 , but per [rad/m] or per [oscillations/m]. So, if z is a displace-
𝑚2 𝑚3 𝑚2
ment in [m], Φ has the unit [ ⁄𝑚
] or [ ⁄𝑚
3
].
332
Vertical Dynamics
[5.25]
(𝑥) ∑ ̂ ∙ cos( ∙𝑥+𝑥 )
=
[5.26]
(𝑥) ̂ ∙ cos( ∙ 𝑥 + 𝑥 )
rad⁄m
m2
ery bad road ∶ 𝛷 00 ∙ 0 6 [ ]
rad⁄m
The waviness is normally in the range of .. [ ]
where smooth roads have larger waviness than bad roads.
The decreasing amplitude for higher (spatial) frequencies (i.e. for smaller wavelength) can be ex-
plained by that height variation over a short distance requires large gradients. On micro-level, in the
granular level in the asphalt, there can of course be steep slopes on each small stone in the asphalt.
333
Vertical Dynamics
These are of less interest in vehicle vertical dynamics, since the wheel dimensions filter out wave
length << tyre contact length, see Figure 2-58. Reference (ISO 8608) uses road waviness, w=2 for all
roads. Figure 5-5 is based on measurements on real roads, which shows that waviness actually varies
with road severity, 𝛷 . A certain road can be described with:
• 𝛺 ⋯ 𝛺𝑁
• ̂ ⋯ ̂𝑁
• 𝑥 ⋯ 𝑥 𝑁
Rough
The road condition is assessed as Rough if Cross Country
the road surface is of poor quality or if the The road condition is assessed as Cross
road is not properly maintained. Up to 5% of Country if a considerable amount of
the total distance may be covered on driving occurs in severe off-road
extremely poor roads or off-road. conditions.
Figure 5-4: Four typical road types. From (AB Volvo, 2011).
-2
Φ Φ ⁄
PHI=PHI0.*((OMEGA/OMEGA0).^(-waviness))
10
very rough very bad road
-3
rough bad road
10 smooth very good road
⁄𝑚
]
-4
𝑚2
10
𝑆𝐷 PHI /[m*m/(rad/m)]
in [
-5
10
Φ
-6
10
-7
10
m2 𝟔
very rough 𝟎 𝟏𝟎𝟎 𝟏𝟎 𝒘 𝟐 𝟏;
ad⁄m
m2
-8 rough 𝟎 𝟏𝟎 𝟏𝟎 𝟔 ad⁄m 𝒘 𝟐. 𝟏
10 m2
smooth 𝟎 𝟏 𝟏𝟎 𝟔 ad⁄m 𝒘 𝟏
𝟎 𝟏 rad⁄m for all;
-9
10
-1 0 1
10 10 in [𝑟 ⁄ ]→ 10
OMEGA /[rad/m]
𝜆 50 0 0 5
Figure 5-5: PSD spectra for the three typical roads in Figure 5-4.
334
Vertical Dynamics
N=20
0 N=100
-0.05
0 1 2 3 4 5 6 7 8 9 10
x/[m]
rough
RoadQuality=2
0.05
N=10
zr /[m]
N=20
0 N=100
-0.05
0 1 2 3 4 5 6 7 8 9 10
x/[m]
smooth
RoadQuality=3
0.05
N=10
zr /[m]
N=20
0 N=100
-0.05
0 1 2 3 4 5 6 7 8 9 10
x/[m]
Figure 5-6: Road profiles, (𝑥), for the three typical roads in Figure 5-4.
{𝑢𝑠𝑒: 𝛷 ∙ ( ) }
𝛺 𝛺
𝑤
𝛷 ∙𝛺 𝑤
𝛷 ( ) 𝛷 𝑤 𝑤
∙ ∙ ∙
𝛺 𝑤∙ 𝛺 𝑤
𝛺 𝑤
2 2 𝑤 𝑤
( ) |𝐻 𝑟→
( )| ∙ ( ) |𝐻 𝑟 → ( )| ∙ 𝑤 ∙ ∙
𝑟
𝛺
Then we can use Equation [5.9] to obtain the RMS of the response :
335
Vertical Dynamics
𝑁 𝑁
𝛷 𝑤
2
𝑤
𝑅𝑀𝑆 ( ) √∑ ( )∙ √ 𝑤 ∙ ∙ ∑|𝐻 𝑟 → ( )| ∙ ∙
𝛺
= =
[5.31]
or
∞ ∞
𝛷 𝑤
2
𝑤
𝑅𝑀𝑆( ) √ ∫ ( )∙ √ 𝑤 ∙ ∙ ∫ |𝐻 𝑟 → ( )| ∙ ∙
𝛺
𝜔= 𝜔=
+ 𝑥
𝑬𝒒 𝒊𝒍𝒊𝒃𝒓𝒊 : ∙ ̈ + ∙
[5.32]
𝑪𝒐 𝒑𝒂𝒕𝒊𝒃𝒊𝒍𝒊𝒕𝒚: 𝑛
𝑬𝒙𝒄𝒊𝒕𝒂𝒕𝒊𝒐𝒏: (𝑡)
336
Vertical Dynamics
∙π
(𝑥) ̂ ∙ cos( ∙ 𝑥 + 𝑥 ) ̂ ∙ cos ( ∙𝑥+𝑥 )
{ 𝜆 }⇒
𝑥 ∙𝑡
𝐴𝑠𝑠𝑢 𝑒 𝑥 0
[5.33]
∙π∙
⇒ (𝑡) ̂ ∙ cos ( ∙ 𝑡) ̂ ∙ cos(ω ∙ 𝑡) ⇒
𝜆
∙π∙ ∙π∙
⇒ (𝑡) ∙ ̂ ∙ sin ( ∙ 𝑡) ω ∙ ̂ ∙ sin(ω ∙ 𝑡) ⇒
𝜆 𝜆
2
∙π∙ ∙π∙
⇒ ̈ (𝑡) ( ) ∙ ̂ ∙ cos ( ∙ 𝑡) ω2 ∙ ̂ ∙ cos(ω ∙ 𝑡)
𝜆 𝜆
Insertion in the model in Equation [5.32] (with eliminated ) gives directly the solution:
[5.34]
{
(𝑡) (𝑡) ̂ ∙ cos (ω ∙ 𝑡) 𝑛 ̈ (𝑡) ̈ (𝑡) ̂ ∙ cos(ω ∙ 𝑡)
ℎ𝑒𝑟𝑒 ̂ 2
∙ω ∙ ̂ 𝑛 ̂ 2
ω ∙ ̂
We can identify the magnitude of the transfer functions 𝐻. The negative sign in Equation [5.35] means
180 degrees phase shift:
ℱ( )
𝐻 𝑟→ {𝐻 𝑟 → } { (𝑡) (𝑡)} +𝑗∙0
ℱ( )
[5.35]
𝐻 𝑟→ 𝑟 {𝐻 𝑟 → 𝑟 𝐻 𝑟→ 𝐻 → 𝐻 𝑟→ } 0+𝑗∙0
𝐻 𝑟→ ̈ {𝐻 𝑟 → ̈ (𝑗 ∙ )2 ∙ 𝐻 𝑟 → ω2 ∙ 𝐻 𝑟 → } ω2 + 𝑗 ∙ 0
𝐻 𝑟 → 𝑟𝑧 {𝐻 𝑟 → 𝑟𝑧 ∙ 𝐻 𝑟→ ̈ } ∙ ω2 + 𝑗 ∙ 0
The motivation to choose exactly those transfer functions is revealed later, in 5.5, 5.5.2 and 5.5.2. For
now, we simply conclude that various transfer functions can be identified and plotted. The plots are
found in Figure 5-8. Numerical values for m and 𝜆 are chosen.
5.4.1.1.1 Example Analysis
An example of how to use Figure 5-8 is: A certain road has amplitude of 1 cm ( ̂ 0.0 ). The vehi-
cle drives on it with a longitudinal velocity of 50 km/h ( ≈ 4 /𝑠 ̂ ≈ .8𝐻 ):
• |𝐻 𝑟 → ̈ ( )| ≈ 05 ⇒ | ̂| 05 ∙ ̂𝑟 05 ∙ 0.0 .05 ⁄𝑠 2 . From this we can calcu-
late 𝑅𝑀𝑆( ̈ 𝑠 ) | .05| ⁄√ ≈ . 6 ⁄𝑠 2 . The RMS value of acceleration will later be related to
ride comfort, see 5.5.
• |𝐻 𝑟 → 𝑢 ( )| 0 is the transfer function to deformation of suspension, which later
will be related to fatigue life, see 5.5.2. The model in 5.4.1 is not good for measuring fa-
tigue, since the ̂ ̂ is intrinsically zero because of no compliance between un-
sprung and sprung mass.
• |𝐻 𝑟 → 𝑟 ( )| ≈ 487000 ⇒ Δ̂ 𝑟 487000 ∙ ̂𝑟 487000 ∙ 0.0 4870 𝑁 . If Δ̂ > ∙
≈ 6000 𝑁, the model is outside its validity region, because it would require pulling forces
between tyre and road. If changing to ̂ 0. , this limit is defined by |𝐻 𝑟 → 𝑟 ( )| >≈
6
.6 05 [𝑁⁄ ], which is used to examplify the validity limit in Figure 5-8;
.
model becomes invalid for >≈ 00 ⁄𝑠. The variation in tyre road contact force will be re-
lated to road grip, see 5.5.2.
The phases for the studied variables can be found in Equation [5.35]. With this model, the phases be-
come constant and ±90 𝑒 .
337
Vertical Dynamics
7
Driving at road with wave-length, lambda = 5 [m].
10
invalid if ̂ 0. [ ]
6
10
𝐻 𝑟→ 𝑟𝑧
.6 05
5
10
4
10
H_zr_zs (==1)
abs(H)
3
10 H_zr_zr-zs (==0)
H_zr_derderzs
2 H_zr_Frz
10
m = ms+mu = 1600[kg];
1
10 f=1 Hz at vx=5[m/s]; f=10 Hz at vx=50[m/s]
0
10
-1
10
0 1 10 Hz 2 3
10 1 Hz 10 10 10
vx [m/s]
Figure 5-8: Transfer functions from model in Figure 5-7, excited with single frequencies.
[5.36]
𝑤
0
∙ 6 ∞ 2
√ ∙ 2.5 ∙ ∫𝜔= |𝐻 𝑟 → ( )| ∙ 2.5 ∙
For now, we simply note that it is possible to calculate this (scalar) RMS value for each vehicle speed
over the assumed road. In corresponding way, an RMS value can be calculated for any of the oscillating
variables, such as ̈ , and . We will come back to Equation [5.36] in 5.5.1.2.
z
̈
+ x
c d
py
̈
̈ ̈
Figure 5-9: One-dimensional model with 1 dynamic degree of freedom
338
Vertical Dynamics
[5.37]
Constitution: ∙( )+ ∙( )+ ∙
Compatibility:
Excitation: (𝑡)
[5.38]
⏟ ( + )∙ ∙ ( (𝑡) ) + ∙ ( (𝑡) )+ ∙ ̈ (𝑡)
Δ 𝑟𝑧
Note that since we measure and from the static equilibrium, the static load, ∙ , disappears
when constitution is inserted in equilibrium. The Δ denotes the variation from static contact force
between road and tyre.
Assume that the road has only one (spatial) frequency, i.e. one wavelength. Then the excitation is as in
Equation [5.26], in which we assume 𝑥 0. So, we can insert (𝑡) ̂ ∙ cos(ω ∙ 𝑡) ⇒ (𝑡)
̂ ∙ sin(ω ∙ 𝑡) ⇒ ̈ (𝑡) 2
̂ ∙ cos(ω ∙ 𝑡) in Equation [5.38] and solve it for (𝑡) and Δ (𝑡) with
trigonometry or Fourier transform.
In 4.4.3.1.1, we applied Fourier transform on the linear explicit form model. To show a slightly other
way, we do not rewrite to explicit form, but apply Fourier transform on Eq [5.38] directly:
∙ ( ω2 ∙ ℱ( ))
[5.39]
∙ (ℱ( ) ℱ( )) + ∙𝑗∙ ∙ (ℱ( ) ℱ( ))
ℱ( 𝑟 ) ∙ (ℱ( 𝑟 ) ℱ( 𝑠 )) + ∙𝑗∙ ∙ (ℱ( 𝑟 ) ℱ( 𝑠 )) 𝑢 ω ∙ ℱ( 𝑟 )
[5.40]
ℱ( ) ℱ( )
𝐻 𝑟 → 𝑟𝑧 ( +𝑗∙ ω2 ) ( + 𝑗 ∙ )∙
ℱ( ) ℱ( )
( +𝑗∙ 2
ω ) ( +𝑗∙ ) ∙ 𝐻 𝑟→
We can elaborate further with Eq [5.40]:
̂ + ∙ ∙ω | 1| √ + ∙ω
Amplitude: |𝐻 𝑟 → | |( | {| 1 | }
̂𝑟 𝑚∙ω2 )+ ∙ ∙ω 2 | 2| √( ∙ω ) + ∙ω
[5.41]
+ ∙ ∙ω
Phase: ( ) ( ) arg ( 𝑚∙ω2 + + ∙ ∙ω
) {arg ( ) 1
arg( ) arg( 2 )}
2
∙ ∙
arg( + 𝑗 ∙ ∙ ω) arg( ∙ ω2 + + 𝑗 ∙ ∙ ω) ⋯ arctan (
∙ ∙ + ∙
)
Equation [5.13] now allows us to get the magnitudes of the other transfer functions as well:
𝐻 𝑟→ from Equation [5.40]
𝐻 𝑟→ 𝑟
𝐻 𝑟→ 𝑟 𝐻 𝑟→ 𝐻 𝑟→
2
𝐻 𝑟→ ̈ ∙ 𝐻 𝑟→
̈ + ̈ +
[5.42]
𝐻 𝑟→ { }
𝑟𝑧 ̈ + ∙( )+ ∙( )
𝐻 𝑟→ ̈𝑟 + ∙ (𝐻 𝑟 → 𝑟
𝐻 𝑟→ ) + ∙ (𝐻 𝑟 → 𝑟
𝐻 𝑟→ )
(𝑗 ∙ )2 + ( + 𝑗 ∙ ∙ ) ∙ (𝐻 𝑟 → 𝑟
𝐻 𝑟→ )
2
+( +𝑗∙ ∙ )∙( 𝐻 𝑟→ )
The motivation to choose exactly those transfer functions is revealed later, in 5.5. Some of those mag-
nitudes are easily expressed in reel (non-complex) mathematics using Equation [5.41]:
339
Vertical Dynamics
+ ∙ω
|𝐻 𝑟 → | √
( ∙ω ) + ∙ω
[5.43]
2
+ ∙ω
|𝐻 𝑟 → ̈ | ∙√
( ∙ω ) + ∙ω
1 Hz 10 Hz 100 Hz
/[ /𝑠]
Figure 5-10: Transfer functions for amplitudes from model in Figure 5-9, excited with single
frequencies. Thin lines are without damping. Notation: 𝐻 → is denoted H_a_b.
340
Vertical Dynamics
/[ /𝑠]
Figure 5-11: Transfer functions for phase delays from model in Figure 5-9, excited with single
frequencies.
We compare these numbers with the corresponding numbers for the simpler model in 5.4.1. The com-
fort is better. The fatigue life and road grip have become more realistic.
Figure 5-10 also shows the curves for the undamped system (d=0). The highest peaks appear at ap-
proximately 5. .6 m/s. This corresponds to the speed where the natural (=undamped) eigen fre-
quency appears ( 𝜆∙𝑓 𝜆∙ ⁄( ∙ 𝜋) 𝜆 ∙ √ ⁄ ⁄( ∙ 𝜋) ≈ 5.5 /𝑠).
Figure 5-11 shows the phase angles for the different responses.
5.4.2.1.2 § Solution with Trigonometry
The purpose with 5.4.2.1.2 is that it might help some readers to better/intuitively understand what the
previously used, more efficient, Fourier transform method does.
One way to solve the mathematical model in Eq [5.37] is to assume a real solution, insert the assump-
tion and find expressions for the coefficients in the assumption. Assume such solution (and that 𝑥 0
in Equation [5.26]): ̂ cos( ∙ 𝑡). Insertion in 1st scalar equation in Eq [5.37] gives:
∙ ̈ + ∙ + ∙ ̂ ∙ ( ∙ cos( ∙ 𝑡) ∙ ∙ sin( ∙ 𝑡))
Assumed solution:
̂ ∙ cos( ∙ 𝑡 ) ⇒
⇒ {𝑢𝑠𝑒: cos( ) cos ∙ cos + sin ∙ sin } ⇒
⇒ ̂ ∙ [cos( ∙ 𝑡) ∙ cos + sin( ∙ 𝑡) ∙ sin ] ⇒
⇒ ̂ ∙ ∙ [ sin( ∙ 𝑡) ∙ cos + cos( ∙ 𝑡) ∙ sin ] ⇒
⇒ ̈ ̂ ∙ 2 ∙ [cos( ∙ 𝑡) ∙ cos + sin( ∙ 𝑡) ∙ sin ]
Insertion:
∙ ̂ ∙ 2 ∙ [cos( ∙ 𝑡) ∙ cos + sin( ∙ 𝑡) ∙ sin ] +
+ ∙ ̂ ∙ ∙ [ sin( ∙ 𝑡) ∙ cos + cos( ∙ 𝑡) ∙ sin ] +
+ ∙ ̂ ∙ [cos( ∙ 𝑡) ∙ cos + sin( ∙ 𝑡) ∙ sin ]
̂ ∙ ( ∙ cos( ∙ 𝑡) ∙ ∙ sin( ∙ 𝑡)) ⇒
𝒄𝒐 𝒕𝒆𝒓 : ∙ ̂ ∙ 2 ∙ cos + ∙ ̂ ∙ ∙ sin + ∙ ̂ ∙ cos ̂ ∙
⇒ { ⇒
𝒊𝒏 𝒕𝒆𝒓 : ∙ ̂ ∙ 2 ∙ sin ∙ ̂ ∙ ∙ cos + ∙ ̂ ∙ sin ̂ ∙ ∙
341
Vertical Dynamics
∙ ∙ 3
arctan ( 2
)
∙ ∙ 2+ 2 ∙ 2
⇒
̂ ∙
2
|𝐻 𝑟→ 𝑠
|
{ ̂ ( ∙ ) ∙ sin + ∙ ∙ cos
∙π∙ 𝑥
ℎ𝑒𝑟𝑒
𝜆
We have identified |𝐻 𝑟 → ( )|, which can be compared to |𝐻 𝑟 → ( )| in Eq [5.35]. The other trans-
fer functions in Eq [5.35] are more difficult to derive using the method with trigonometry.
zr
Frz
Figure 5-12: One-dimensional model with two dynamic degrees of freedom
The corresponding mathematical model becomes as follows:
Equilibrium:
∙ ̈ ∙
∙ ̈ ∙
[5.44]
342
Vertical Dynamics
0 ̈ 0 0
[ ]∙[ ]+[ ]∙[ ]+[ + ]∙[ ] [ ]∙ +[ ]∙
0 ̈ +
⇒
̈
⇒ 𝑴 ∙ [ ] + 𝑫 ∙ [ ] + 𝑪 ∙ [ ] 𝑫 𝒓 ∙ + 𝑪𝒓 ∙ ⇒
̈
[5.45]
ℱ( ) ℱ( ) ℱ( )
⇒ 𝑴∙( 2 ∙[ ]) + 𝑫 ∙ (𝑗 ∙ ∙ [ ]) + 𝑪 ∙ [ ]
ℱ( ) ℱ( ) ℱ( )
𝑫𝒓 ∙ (𝑗 ∙ ∙ ℱ ( )) + 𝑪𝒓 ∙ ℱ ( ) ⇒
ℱ( )
⇒ ( 2 ∙ 𝑴 + 𝑗 ∙ ∙ 𝑫 + 𝑪) ∙ [ ] (𝑗 ∙ ∙ 𝑫𝒓 + 𝑪𝒓 ) ∙ ℱ ( )
ℱ( )
[5.46]
𝐻 → ℱ( ) 2
[ 𝑟 ] [ ]∙ ( ∙𝑴+𝑗∙ ∙ 𝑫 + 𝑪) ∙ (𝑗 ∙ ∙ 𝑫 𝒓 + 𝑪𝒓 )
𝐻 𝑟→ 𝑢 ℱ( ) ℱ( )
This format is very compact, since it includes both transfer functions for amplitude and phase. For nu-
merical analyses, the expression in Eq [5.46] is explicit enough, since there are tools, e.g. Matlab, which
do numerical matrix inversion and complex mathematics. Symbolic solution is rather lengthy, but one
can use symbolic tools, e.g. Mathematica or Matlab Symbolic Toolbox.
Expression in real (without phase information) can be derived, see Eq [5.47].
2
√( ∙ ∙ ∙ 2 )2 +( ∙( ∙ + ∙ ))
|𝐻 𝑟 → ̈ | 2
∙
√𝐴2 + 𝐵2
∙ √( ∙ 2 )2 + ( ∙ 3 )2
[5.47]
|𝐻 𝑟 → 𝑢
|
√𝐴2 + 𝐵2
√( ∙ ∙ 4+ 2 ∙( + )∙ )2 + ( 3 ∙( + )∙ )2
|𝐻 𝑟 → 𝑟 𝑢
|
√𝐴2 + 𝐵2
4 2
𝐴 ∙ ∙ ∙( ∙ + ∙ + ∙ + ∙ )+ ∙
3
𝐵 ∙( ∙ + ∙ + ∙ ) ∙( ∙ + ∙ )
With 0, the solutions (with phase information, i.e. complex) becomes as follows:
( ∙ ω2 + 𝑗 ∙ ∙ω+ + )∙
( + 𝑗 ∙ ∙ ω)2
∙ ω2 + 𝑗 ∙ ∙ω+ +
ℱ( ) ∙ ω2 + 𝑗 ∙ ∙ ω +
𝐻 𝑟→
ℱ( ) +𝑗∙ ∙ω
[5.48]
ℱ( )
𝐻 𝑟→ 𝑢
ℱ( ) ( + 𝑗 ∙ ∙ ω)2
∙ ω2 + 𝑗 ∙ ∙ω+ +
∙ ω2 + 𝑗 ∙ ∙ ω +
∙π∙ 𝑥
where 𝜆
Equation [5.13] now allows us to get the magnitudes of the other transfer functions as well:
343
Vertical Dynamics
𝐻 𝑟→ {use Eq [5.48]}
𝐻 𝑟→ 𝑢 {use Eq [5.48]}
𝐻 𝑟→ 𝑟 𝑢
𝐻 𝑟→ 𝑟 𝐻 𝑟→ 𝑢
𝐻 𝑟→ 𝑢
;
𝐻 𝑟→ 𝑢 𝐻 𝑟→ 𝑢 𝐻 𝑟→ 𝑠
;
[5.49]
𝐻 𝑟→ ̈ ∙ 𝐻 𝑟→ 𝑠 ;
𝐻 𝑟 →Δ 𝑧
{Δ ∙( )+ ∙( )}
∙ (𝐻 𝑟 → 𝑢 𝐻 𝑟 → ) + ∙ 𝑗 ∙ ∙ (𝐻 𝑟 → 𝑢 𝐻 𝑟 → )
( +𝑗∙ ∙ ) ∙ (𝐻 𝑟 → 𝑢 𝐻 𝑟 → );
𝐻 𝑟 →Δ 𝑟𝑧
{Δ ∙( )} ∙ (𝐻 𝑟 → 𝑟 𝐻 𝑟 → 𝑢 ) ∙( 𝐻 𝑟→ 𝑢
);
The transfer functions in Equation [5.49] are plotted in Figure 5-13.
5.4.3.1.1 Example Analysis
If we use Figure 5-13 as the example in 5.4.1:
• Ride comfort related: |𝐻 𝑟 → ̈ ( )| ≈ ⇒ | ̂| ∙ ̂𝑟 ∙ 0.0 . ⁄𝑠 2 .
From this we can calculate 𝑅𝑀𝑆( ̈ 𝑠 ) | . | ⁄√ ≈ 0.8697 ⁄𝑠 2 .
• Fatigue life related: |𝐻 𝑟 → 𝑢 ( )| ≈ . 4 ⇒ | ̂ ̂ | . 4 ∙ ̂𝑟 . 4 ∙ 0.0
0.0 4 . 4 .
• Road grip related: |𝐻 𝑟 →Δ 𝑟 ( )| ≈ 77470 ⇒ | 𝑟 | 77470 ∙ ̂𝑟 77470 ∙ 0.0
775 𝑁 .
7
Driving at road with wave-length, lambda = 5 [m].
10
6
10
5
10
ms = 1415[kg]; cs = 68 [kN/m]; ds = 8[kN/(m/s)];
4
10 mu = 185[kg]; ct = 676 [kN/m];
f=1 Hz at vx=5[m/s]; f=10 Hz at vx=50[m/s]
abs(H)
3
10
2
10
1
10 H_zr_zs
H_zr_zu-zs
0
10 H_zr_derderzs
H_zr_Frz
-1
10
0 1 10 Hz 2 3
10 10 10 10
1 Hz vx [m/s]
Figure 5-13: Transfer functions for amplitudes from model in Figure 5-12, excited with single
frequencies.
This analysis can be compared with the analysis in 5.4.2.1.1. Ride comfort and fatigue does not change
a lot, but road grip does. This indicates that the more advanced model is only needed for road grip
evaluation.
Figure 5-14 shows the phase angles for the different responses.
344
Vertical Dynamics
200
150
100
50
arg(H)
0
-50
-100
H_zr_zs
H_zr_zu-zs
-150
H_zr_derderzs
H_zr_Frz
-200
0 1 2 3
10 10 10 10
vx [m/s]
Figure 5-14: Transfer functions for phase delays. Same model and data as in Figure 5-13.
Figure 5-15, shows the amplitude gains for the corresponding un-damped system. Natural frequencies
are around 5 m/s and 50 m/s. These two speeds correspond to frequencies /𝜆, i.e. approxi-
mately 1 Hz and 10 Hz. The 1Hz frequency is an oscillation mode where the both masses move in
phase with each other, the so called “heave mode” or “bounce mode”. The 10 Hz frequency comes from
the mode where the masses are in counter-phase, the so called “wheel hop mode”. In the wheel hop
mode, the sprung mass is almost not moving at all. We will come back to these modes in 5.4.4.
7
10
6
10
5
10
4
10 H_zr_zs
H_zr_zu-zs
abs(H)
3
10 H_zr_derderzs
H_zr_Frz
2
10
1
10
0
10
-1
10
0 1 2 3
10 10 10 10
vx [m/s]
Figure 5-15: Un-damped transfer functions. Same model and data as in Figure 5-13, except 0.
345
Vertical Dynamics
2 𝑝 | 𝑝 | . But, regardless of the exact curve form, we can conclude that a traditional damper
can only dissipate energy. A step towards controllable suspension is to use a semi-active damper,
which has varying hydraulic orifices. It can still only dissipate energy, but the curve shape is variable.
If adding energy via a hydraulic pump, one can take one more step, called a “active damper”. In princi-
ple, an active damper can generate both positive and negative force (subscript for active
damper) for any sign on 𝑝 . A common way to design control algorithms for such dampers is to use
the to mimick a virtual “Skyhook damper” connected between vehicle unsprung parts and “the
sky”, i.e. a fix inertial system.
Ctrl
̈
Figure 5-16: § One-dimensional model with two dynamic degrees of freedom and with the
traditional dampers replaced by a Skyhook controlled active damper (ideal sensing, computation
and actuation). Tyre damping neglected.
The corresponding mathematical model becomes as follows:
Equilibrium:
∙ ̈ + ∙
∙ ̈ ∙
Constitution (displacements counted from static equilibrium):
∙( )+ ∙
∙( )+( + )∙
Control and actuation: ∙ ∙
Excitation: (𝑡 )
The same can be formulated with matrices and Fourier transforms:
0 ̈ 0 0
[ ]∙[ ]+[ ]∙[ ]+[ + ]∙[ ] [ ]∙
0 ̈ 0
Eq [5.46] is still valid with same matrix names, but the definition of matrix 𝑫 is different. No transfer
functions are plotted here, but in 5.6.5.
346
Vertical Dynamics
Modes Models
(Arrows marks displacement amplitudes) heave (or bounce) mode wheel hop mode
𝒄 𝒅
𝒄 𝒅
[5.50]
1 1
⁄( + ) ad + ad
√ 6.6 .05 H √ 6 .4 0. H
𝑛 𝑚 𝑊 𝑙𝐻 𝑝 𝑚𝑢
347
Vertical Dynamics
why the different diagrams cannot be directly compared to each other. The SAE has suggested that fre-
quencies from 4 to 8 Hz are the most sensitive and the accepted accelerations for these are no higher
than 0.025 g (RMS).
The curves in Figure 5-18 mostly represent an extended exposure to the vibration. As one can expect, a
human can endure exposure to more severe conditions for short periods of time. The SAE limits pre-
sented are indicative of 8 hours of continuous exposure. Curves for different exposure times can also
be obtained from ISO, (ISO 2631). The ISO curves are from the first version of ISO 2631 and were later
modified, see Figure 5-19.
RMS not
accele- OK
ration
OK
4 Hz 8 Hz Frequency
Figure 5-18: Various Human Tolerance Curves to Vertical Vibration, (Gillespie, 1992)
348
Vertical Dynamics
0
/2
𝑊
𝑊𝑘
𝑚
-5
0
𝑊
0𝑛
10
𝑊𝑘 𝑊𝑘
-10
𝑊𝑘
⇒ 𝑊𝑘
𝑊𝑘 (dB)
𝑊
W in dB
[1] ]
-1 -15
Wk [
10
(number= 0 log Gain
𝑊
𝑊
-20
-2
10 -25 Wd (horizontal)
Wk (vertical)
-30
-3
10 0 1
-1
10 10
0 1
10 10
2
10
3 10 10
f [Hz] Frequency (Hz)
Figure 5-20: Human Sensitivity Filter Function. From (ISO 2631). Right: Asymptotic
approximation
349
Vertical Dynamics
Figure 5-21: Human Filter Function for vertical vibrations. Table from (ISO 2631).
With formulas from earlier in this chapter we can calculate an RMS value of a signal with multiple fre-
quencies, see Equation [5.6]. Consequently, we can calculate RMS of multiple frequency acceleration.
Since humans are sensitive to acceleration, it would give one measure of human discomfort. However,
to get a measure which is useful for comparing accelerations with different frequency content, the
measure has to take the human filter function into account. The Weighted RMS Acceleration, a w, in the
following formula is such measure:
𝑁 2 𝑁 2
̂̈ (𝑊𝑘 ( ) ∙ ̂̈ )
𝑤 𝑤 ( ̈ (𝑡)) 𝑢𝑠𝑒: 𝑅𝑀𝑆( ̈ (𝑡)) √∑ √∑ or
= =
[5.51]
{ }
∞ ∞
2
𝑤 𝑤( ̈ (𝑡)) {𝑢𝑠𝑒: 𝑅𝑀𝑆( ̈ (𝑡)) √ ∫ ̈( )∙ } √ ∫ (𝑊𝑘 ( )) ∙ ̈( )∙
𝜔= 𝜔=
Equation [5.51] is written for a case with only vertical vibrations, hence 𝑊𝑘 and ̈ . If vibrations in sev-
eral directions, a total 𝑤 can still be calculated, see (ISO 2631).
In (ISO 2631) one can also find the following equation, which
[5.52]
350
Vertical Dynamics
mass” gives much different comfort value than the two others, so the simplest is not goof to estimate
comfort. However, the two other models give approximately same result, which indicates that the me-
dium model, “stiff tyre, no unsprung mass”, is enough for comfort evaluation. This is no general truth
but an indication that the most advanced model, “two masses, elastic tyre”, is not needed for comfort on
normal roads. The advanced model is more needed for road grip.
We can also see that the comfort decreases, the faster the vehicle drives. If we read out at which speed
we reach 𝑤 ⁄𝑠 2 (which is a reasonable value for long time exposure) we get around ≈
70 ⁄𝑠 ≈ 50 𝑘 /ℎ on this road type (“Rough”) with the medium (and advanced) model. With the
simplest model, we get ≈ ⁄𝑠 ≈ . .4 𝑘 /ℎ.
∞
Eq 𝛷 2 𝐻 𝑟→ ̈
𝑅𝑀𝑆( ̈ ) { } √ ∙ 𝑤 ∙ ∫ |𝐻 𝑟 → ̈ ( )| ∙ 𝑤 ∙ { }
[5. ] 𝑤 2
𝛺 ∙ 𝐻 𝑟→
𝜔=
∞
𝛷 𝑤 2
2
𝑤
√ 𝑤 ∙ ∙ ∫ | ∙ 𝐻 𝑟 → ( )| ∙ ∙
𝛺
[5.53]
𝜔=
∞
𝛷 𝑤
2
4 𝑤
Eq
√ 𝑤 ∙ ∙ ∫ |𝐻 𝑟 → ( )| ∙ ∙ ⇒{ }⇒
𝛺 [5.5 ]
𝜔=
∞
𝛷 𝑤
2 2
4 𝑤
⇒ 𝑤 √ 𝑤 ∙ ∙ ∫ (𝑊𝑘 ( )) ∙ |𝐻 𝑟 → ( )| ∙ ∙
𝛺
𝜔=
351
Vertical Dynamics
2
Ride Comfort. For road type "Rough"
10
1
10
Weigthed RMS value, aw [m/(s*s)]
0
10
-1
10
-2
10
-3
10
-2 -1 0 1 2
10 10 10 10 10
vx [m/s]
Figure 5-22: Weighted RMS values for road type “Smooth” from Figure 5-5. The 3 curves show 3
different models: Simplest (from 1.5.1), Medium (from 1.5.2) and Most advanced (from 1.5.3).
352
Vertical Dynamics
[5.54]
√𝛺 𝑤 ∙ 𝑥 ∙ ∫ 0|𝐻 𝑟→ 𝑢 𝑠
( )| ∙ ∙ {
𝐻 𝐻 𝑟→
𝑠
}
0 𝑟→ 𝑢 𝑠
𝛷0 2
√𝛺 𝑤 ∙ 𝑥 ∙ ∫ 0|𝐻 𝑟→ 𝑢 ( ) 𝐻 𝑟→ 𝑠 ( )| ∙ ∙
0
Equation is written for application to a known road spectra (𝛷 ) and vehicle dynamic structure
(𝐻 𝑟 → 𝑢 𝐻 𝑟 → 𝑠 ), but the first expression (𝑅𝑀𝑆( 𝑢 (𝑡) 𝑠 ( 𝑡))) is applicable on a measured or simu-
lated time domain solution.
𝑤
0
{𝑢𝑠𝑒: 𝐻 𝑟 → 𝑢 𝑠 𝑡 ∙( 𝐻 𝑟 → 𝑢 )}
𝛷0 2
√𝛺 𝑤 ∙ 𝑥 ∙ ∫ 0| 𝑡 ∙( 𝐻 𝑟→ 𝑢
)| ∙ ∙
0
Equation is written for application to a known road spectrum (𝛷 ) and vehicle dynamic structure
(𝐻 𝑟 → 𝑢 ), but the first expression (𝑅𝑀𝑆(Δ 𝑟 𝑡 ) is applicable on a measured or simulated time do-
( ))
main solution.
353
Vertical Dynamics
• An area of functions that encompasses the vertical dynamics is Noise, Vibration, and Harsh-
ness – NVH. It is similar to ride comfort, but the frequencies are higher, stretching up to sound
which is heard by humans.
• Ground clearance (static and dynamic) between vehicle body and ground. Typically, im-
portant for off-road situations.
• Longitudinal comfort, due to drive line oscillations and/or vertical road displacements. Espe-
cially critical when driver cabin is separately suspended to the body. This is the case for heavy
trucks.
• Disturbances in steering wheel feel, due to one-sided bumps. Especially critical for rigid
steered axles. This is often the design of the front axle in heavy trucks.
• There are of course an infinite number of combined manoeuvres, in which functions with re-
quirements can be found. Examples can be bump during strong cornering (possibly destabi-
lizing vehicle) or one-sided bump (exciting both heave=bounce, pitch and roll modes). When
studying such transients, the vertical dynamics is not enough to capture the comfort, but one
often need to involve also longitudinal dynamics; the linkage with ant-dive/anti-squat geome-
try from Chapter 3 becomes important as well as tyre vertical (radial) deflection characteris-
tics.
• Energy is dissipated in suspension dampers, which influence energy consumption for the ve-
hicle. This energy loss is much related, but not same as, to (tyre) rolling resistance. Suspension
characteristics do influence this energy loss, but it is normally negligible, unless driving very
fast on very uneven road.
354
Vertical Dynamics
3
10
2
10
1
10
amplification /[(m/(s*s))/(m)]
0
10
-1
10
-2
10
-3
10
1
355
Vertical Dynamics
Regarding Figure 5-24 and Figure 5-25 we see that there is a large influence of the acceleration gain at
low frequencies with little change at the wheel hop and higher frequencies. The suspension stiffness
and damping were seen to have little influence on the ride comfort / road grip response around 10 Hz.
0
10
-1
10
-2
10
-3
10
-1 0
10 10
356
Vertical Dynamics
2
10
0
10
-1
10
-2
10
-3
10
-1
10
2
Com
10
0
10
-1
10
-2
10
-3
10
-1
10
357
Vertical Dynamics
In figure below, the suspension dampers are replaced by ideal active dampers (ideal sensing, control
algorithm computation and actuation). One can compare with how variation of (traditional) damper in
Figure 5-25. The active dampers improve comfort and road grip, except for at wheel hop frequency
around 10 Hz. The fatigue is generally worse. The transfer functions are plotted using the model in
5.4.3.2.
358
Vertical Dynamics
m*az m*g
z
zl zr J*der(wx)
px
zrl zrr
zrl
zrr Flz Frrz
x x
Figure 5-29: Heave and roll model. Anti-roll bar not drawn but can be included in equations in
matrix 𝑪.
No equations are formulated for this model in this compendium, but a model will typically show two
different modes, the heave and roll. Heave Eigen frequency is typically 1-1.5 Hz for a passenger car, as
mentioned before. The roll frequency is similar or somewhat higher.
If modelling unsprung masses without inertia, we still get 2 state variables, heave and roll . Using
same mathematical form of equations as in Eq [5.45] we get this model (subscripts 𝑟 for “road left”
and 𝑟𝑟 for “road right”):
𝑴 ∙ 𝒛̈ + 𝑫 ∙ 𝒛 + 𝑪 ∙ 𝒛 𝑫 𝒓 ∙ 𝒛 𝒓 + 𝑪𝒓 ∙ 𝒛 𝒓
𝑙
ℎ𝑒𝑟𝑒 𝒛 [ ] 𝑛 𝒛𝒓 [ ]
[5.56]
The disturbances from the road are two independent ones, so the
transfer functions will be a × matrix:
ℱ( ) ℱ( 𝑙 ) 𝐻 → 𝐻 𝑟𝑟 → ℱ( 𝑙 )
[ ] 𝑯 [ ] [ 𝑟 ] [ ]
ℱ( ) ℱ( ) 𝐻 𝑟 →𝜑 𝐻 𝑟𝑟 →𝜑 ℱ( )
Note that the restoring matrix might need to include both elastic restoring (wheel springs and anti-
roll-bars) and (roll) pendulum effects, see 4.3.10.2 and Reference (Mägi, The Significance of System
Pre-Load at Modal Analysis of Low-Resonant Mechanical Systems, 1988). For high-loaded trucks, the
pendulum effect is really relevant, while it often can be omitted for a low sportscar.
359
Vertical Dynamics
ℎ 𝑓 𝑓
Motion node
for Heave mode
𝑓
𝑓
𝑓
Figure 5-30: Heave and pitch physical model. Figure 5-31: Oscillation modes of a
Heave and Pitch model.
Low heave
excitation
Low pitch
excitation
360
Vertical Dynamics
Constitution: 𝑓 𝑓 + 𝑓 ( 𝑓 𝑓 ) 𝑛 𝑓 𝑓 ( 𝑓 𝑓 ) 𝑛
+ ( ) 𝑛 ( )
Above model is formulated on first order form ( ⋯ and (⋯ ) ) to show an alternative to
second order differential equation form ( ̈ ⋯ ) used in 5.4.
We need to express in differentiated variables. Then we can either assume [ ] [ ] (ve-
locity components in ground fixed directions) or [ ] [ 𝑣 𝑣 ] (vehicle fixed directions). Both
are correct, in similar way as for the yaw rotation, see 4.4.2.3.3. If ground fix: [ ] [ ] . If
ground fix: [ 𝑣 𝑣 ] [ 𝑣 + 𝑣 ≈ 𝑣 𝑣 𝑣 ] . The intention for 5.7.2 is to study con-
stant speed over ground, so we know 0 . Therefore it is easiest to use ground fix directions. We
can compare with the pitching model in 3.4.5, which is typically used for longitudinal acceleration and
braking. Then it is most natural to use vehicle fix and solve 𝑣 and 𝑣 as state variables in an ode.
We formulate the matrix form of the model in ground fix direction :
0 0 0
0 0 0
0 0 ⁄𝑓 0 𝑓
[⏟0 0 0 ⁄ ] [ ]
𝑴𝑪
𝑓 𝑓 𝑓 𝑓
2 2
𝑓 𝑓 𝑓 𝑓 𝑓 + 0 𝑓 𝑓 𝑓
[ ]+[ ]+[ ] [ ]
+𝑓 0 0 𝑓 0 0 ⏟
[ 0 0 ] ⏟ ⏟ 0 ⏟0 𝒛𝒗𝒓
⏟
𝒛𝒗𝑭 𝒈 𝑫𝒓
𝑫
361
Vertical Dynamics
𝐻 𝑟𝑓 →𝑣𝑧
𝐻 𝑟𝑓 →𝜔𝑦
⇒ 𝓕(𝒛𝒗𝑭𝟎 ) (𝑗 𝑴𝑪 𝑫) 𝑫𝒓 𝑗 𝒅𝝀 ℱ( 𝑓) ℱ( 𝑓) 𝑯𝒛𝒓𝒇 →𝒗𝒛 𝒄𝒐𝒓𝒓 ℱ( 𝑓)
𝐻 𝑟𝑓 → 𝑓
[ 𝐻 𝑟𝑓 → 𝑟 ]
All 𝐻 depend on (or 𝜆) and .
5.7.2.2.2 Uncorrelated
We now assume that front and rear are excited “uncorrelated”. This is wrong if driving on a road
where rear axle follows front axle, but it is correct for a vehicle with independent excitation under
each axle, which can be achieved e.g. in a shake rig.
ℱ( 𝑓 ) ℱ( 𝑓) ℱ( 𝑓)
𝓕(𝒛𝒗𝒓 ) [ ] [ ] 𝑗 [ ]
ℱ( ) ℱ( ) ℱ( )
𝐻 𝑟𝑓 →𝑣𝑧 𝑛 𝐻 𝑟𝑟 →𝑣𝑧 𝑛
ℱ( 𝑓)
𝐻 𝑟𝑓 →𝜔𝑦 𝑛 𝐻 𝑟𝑟 →𝜔𝑦 𝑛 ℱ( 𝑓)
𝓕(𝒛𝒗𝑭𝟎 ) (𝑗 𝑴𝑪 𝑫) 𝑫𝒓 𝑗 [ ] [ ]
ℱ( ) 𝐻 𝑟𝑓 → 𝑓 𝑛 𝐻 𝑟𝑟 → 𝑓 𝑛 ℱ( )
][ 𝐻 𝑟𝑓 → 𝑟 𝑛 𝐻 𝑟𝑟 → 𝑟 𝑛
If 𝑓 and 𝑓 have the same amplitude(frequency) content, and is called , we can write:
𝐻 𝑟𝑓 →𝑣𝑧 𝑛 𝐻 𝑟𝑟 →𝑣𝑧 𝑛
𝐻 𝑟𝑓 →𝜔𝑦 𝑛 𝐻 𝑟𝑟 →𝜔𝑦 𝑛
𝓕(𝒛𝒗𝑭𝟎 ) [ ] ℱ ( ) 𝑯𝒛𝒓 →𝒗𝒛 𝒏𝒄𝒐𝒓𝒓 ℱ ( )
𝐻 𝑟𝑓 → 𝑓 𝑛 𝐻 𝑟𝑟 → 𝑓 𝑛
[ 𝐻 𝑟𝑓 → 𝑟 𝑛 𝐻 𝑟𝑟 → 𝑟 𝑛 ]
The elements in 𝐻 depend on (or 𝜆) and .
ℎ𝑒𝑟𝑒 𝒛 [ ]
𝑓𝑙 0
𝑓 0 𝑓𝑙
𝑛 𝒛𝒓 [ ] [ ( )
cos + 𝑗 sin ( ) 0
] [ ]
𝑙 𝑓
[5.57]
0 cos( ) + 𝑗 sin( )
ℎ𝑒𝑟𝑒 𝜋 𝐿⁄
The disturbances from the road are two independent ones, so the transfer
functions will be a × matrix:
ℱ( ) 𝐻 𝑟𝑓 → 𝐻 𝑟𝑓𝑟 →
ℱ( 𝑓𝑙 ) ℱ( 𝑙 )
[ℱ( ) ] 𝑯 [ ] [𝐻 𝑟𝑓 →𝜑 𝐻 𝑟𝑓𝑟 →𝜑 ] [ ]
ℱ( 𝑓 ) ℱ( )
ℱ( ) 𝐻 𝑟𝑓 →𝜑𝑦 𝐻 𝑟𝑓𝑟 →𝜑𝑦
362
Vertical Dynamics
Figure 5-33: Response of Vehicle for Front and Rear Axle Impulses, (Gillespie, 1992)
Models for studying transient vertical dynamics can, in general be categorized as the stationary oscilla-
tion models, 1D, 2D and 3D. But they cannot generally be linear, so they require simulation, not fre-
quency analysis. One typically need to add inertia of unsprung parts and vertical elasticities in each
tyre. And “trivial linkage” suspension is generally not enough if sharp road unevenness, but instead
one might identify the pivot axis in space for each wheel linkage.
A 3D model according to these concepts gets the states 𝒛 containing 𝑓𝑙 𝑓 𝑙 𝑛 if
modelled with a second order differential equation (𝒇 𝒛̈ 𝒛 𝒛 𝑡
( ) 0 ). If modelled with first order
differential equations (𝒇(𝒛 𝒛 𝑡) 0 ) and the concept of using forces in elasticities as states, see
1.5.2.1.2, the states will instead contain 𝑓𝑙 𝑓 𝑙 𝑓𝑙 𝑓 𝑙 𝑛 ,
where is vertical velocity of unsprung mass in wheel 𝑗 and is elastic part of vertical force un-
der wheel 𝑗. The inputs (disturbance) 𝒛𝒓 will contain 𝑓𝑙 𝑓 𝑙 𝑛 .
363
Vertical Dynamics
the stronger propulsion, brake and steering control functions. The compendium does not go deep into
this area, but a sample is seen in 5.4.3.2.
364
Bibliography
BIBLIOGRAPHY
AB Volvo. (2011). Global Transport Application.
Andreasson, J. (2007). On Generic Road Vehicle Motion Modelling and Control. Stockholm: KTH.
Bakker, E. N. (1987). Tyre Modelling for use in Vehicle Dynamics Studies.
Barnard, R. (2010). Road Vehicle Aerodynamic Design. Mechaero Publishing.
Barnard, R. H. (u.d.). Road Vehicle Aerodynamic Design. Mechaero Publishing.
Besselink, I. J. (2011). Experiences with the TYDEX standard tyre interface and file format. Vehicle
System Dynamics. doi:https://doi.org/10.1080/00423110500109299
Boerboom, M. (2012). Electric Vehicle Blended Braking maximizing energy recovery while maintaining
vehicle stability and maneuverability. Göteborg, Sweden: Chalmers University of Technology.
Brian Bole, e. a. (2012). Energy Management Control of a Hybrid Electric Vehicle with Two-Mode
Electrically Variable Transmission. Los Angeles: 2012 Electric Vehicle Symposium, EVS26.
doi:10.13140/2.1.2827.6480
Bruzelius, F. (2004). Linear Parameter-Varying Systems - an approach to gain scheduling (Vol. Doctoral
thesis). Chalmers University of Technology.
Clark, S. (. (1971). Mechanics of Pneumatic Tires, Monograph 122. National Bureau of Standards, USA.
Cooper Tire & Rubber Co. (den 16 September 2007). Passenger radial tire cutaway. Online Art. (Cooper
Tire & Rubber Co.) Hämtat från http://deantires.com/us/en/information/info-
construction.asp
DIRECTIVE 2002/44/EC. (2002). DIRECTIVE 2002/44/EC OF THE EUROPEAN PARLIAMENT AND OF
THE COUNCIL.
Drenth, E. F. (1993). Brake Stability of Front Wheel Driven Cars at High Speed. Delft, Netherlands: Delft
University of Technology.
Dugoff, H., Fancher, P., & Segel, L. (1969). Tire performance characteristics affecting vehicle response to
steering and braking control inputs. Final report. Washington, US. Hämtat från
http://hdl.handle.net/2027.42/1387
Encyclopædia Britannica Online. (2007, September 16). belted tire: tire designs." Online Art.
Encyclopædia Britannica Online. Retrieved from http://www.britannica.com/eb/art-7786
Gillespie, T. (1992). Fundamentals of Vehicle Dynamics. Society of Automotive Engineers.
Grosch, K., & Schallamach, A. (1961, September–October). Tyre wear at controlled slip. Wear, Volume
4(Issue 5), pp. Pages 356–371.
Happian-Smith, J. (2002). An Introduction to Modern Vehicle Design. Butterwoth-Heinemann, ISBN 0-
7506-5044-3.
Hedman, A. (1994). Intermittent Engine Operation -- A Way to Reduce Vehicle Fuel Consumption. Beijing:
FISITA, XXV FISITA Congress, 17-21 October 1994.
Hirschberg, W., Rill, G., & Weinfurter, H. (2007). Tire model TMeasy. Vehicle System Dynamics, 45:1, pp.
101 — 119. doi:10.1080/00423110701776284
Hucho, W.-H. (1998). Aerodynamics of Road Vehicles. SAE International.
ISO 11026. (u.d.). ISO 11026 Heavy commercial vehicles and buses - Test method for roll stability -
Closing-curve test. International Organization for Standardization, Genève, Switzerland.
ISO 14791. (u.d.). ISO 14791 Road vehicles - Heavy commercial vehicle combinations and articulated
buses - Lateral stability test methods. International Organization for Standardization, Genève,
Switzerland.
ISO 14792. (u.d.). ISO 14792 Road vehicles - Heavy commercial vehicles and buses - Steady-state circular
tests. International Organization for Standardization, Genève, Switzerland.
365
Bibliography
ISO 14793. (u.d.). ISO 14793 Road vehicles – Heavy commercial vehicles and buses – Lateral transient
response test methods. International Organization for Standardization, Genève, Switzerland.
ISO 14794. (2011). ISO 14794 Heavy commercial vehicles and buses - Braking in a turn - Open-loop test
methods. International Organization for Standardization, Genève, Switzerland.
ISO. (2006). Passenger cars – Braking in a turn – Open-loop test method. ISO.
ISO. (2011). ISO 14794 Heavy commercial vehicles and buses - Braking in a turn - Open-loop test
methods. ISO.
ISO 26262. (2011-2012). Road vehicles – Functional safety, 1-10. International Organization for
Standardization, Genève, Switzerland.
ISO 2631. (n.d.). ISO 2631 – Evaluation of Human Exposure to Whole-Body Vibration. International
Organization for Standardization, Genève, Switzerland.
ISO 3888. (u.d.). ISO 3888 Passenenger Cars -- Test track for severe lane change manouvre -- Part 2:
Obstacle Avoidance. International Organization for Standardization, Genève, Switzerland.
ISO 4130. (u.d.). ISO 4130: Road vehicles -- Three-dimensional reference system and fidicial marks --
Definitions, 1979.
ISO 4138. (u.d.). ISO 4138 Passenger cars – Steady-state circular driving behaviour - Open-loop test
methods. International Organization for Standardization, Genève, Switzerland.
ISO 7401. (u.d.). ISO 7401 Lateral transient response test methods - Open-loop test methods.
International Organization for Standardization, Genève, Switzerland.
ISO 7975. (2006). ISO 7975 Passenger cars – Braking in a turn – Open-loop test method. International
Organization for Standardization, Genève, Switzerland.
ISO 80000-3. (2013). ISO 80000 – Quantities and units – Part 3: Space and time. International
Organization for Standardization, Genève, Switzerland.
ISO 8608. (u.d.). ISO 8608 Mechanical vibration - Road surface profiles - Reporting of measured data.
International Organization for Standardization, Genève, Switzerland.
ISO 8855. (n.d.). ISO 8855 – Road vehicle – Vehicle dynamics and road holding ability – Vocabulary.
International Organization for Standardization, Genève, Switzerland.
ISO19377. (2017). ISO 19377 Heavy commercial vehicles and buses - Emergency braking on a defined
path - Test method for trajectory measurement. ISO.
ISO28580. (2018). Passenger car, truck and bus tyre rolling resitance measurement method -- single
point test and correlation of measurement results. ISO.
Jacobson, B. (1993). Gear Shifting with Retained Power Transfer. Göteborg, Sweden: Chalmers
Univeristy of Technology. Hämtat från https://research.chalmers.se/en/publication/1485
Jacobson, B. e. (2014). Vehicle Dynamics Compendium for Course MMF062. Göteborg, Sweden: Chalmers
University of Technology.
Jacobson, B., Sundström, P., Kharrazi, S., Fröjd, N., & Islam, M. (2017). An Open Assessment Tool for
Performance Based Standards of Long Combination Vehicles. Hämtat från
https://research.chalmers.se/en/publication/251269
Jonasson, M. (2009). Exploiting individual wheel actuators to enhance vehicle dynamics and safety in
electric vehicles (Vol. Doctoral thesis). Stockholm: Royal Insitute of Technology.
Jonsson, A., & Olsson, E. (2016). A Methodology for Identification of Magic Formula Tire Model
Parameters from In-Vehicle Measurements. Chalmers University of Technology. Hämtat från
http://studentarbeten.chalmers.se/publication/239258-a-methodology-for-identification-of-
magic-formula-tire-model-parameters-from-in-vehicle-measurements
Kati, M. S. (2013). Definitions of Performance Based Characteristics for Long Heavy Vehicle
Combinations. Signals and Systems. Chalmers University of Technology. Hämtat från
http://publications.lib.chalmers.se/records/fulltext/176464/176464.pdf
Kharrazi , S. (2012). Steering Based Lateral Performance Control of Long Heavy Vehicle Combinations.
Göteborg, Sweden: Chalmers University of Technology.
366
Bibliography
367
Bibliography
368
Modules in the Course
MMF062 Vehicle Dynamics
369
Modules in the Course
MMF062 Vehicle Dynamics
• Construct FBDs for mechanical systems, so they support formulating equations.
• Know that fictive forces can be drawn in FBD (and in compendium they are
drawn with dashed arrows).
• Understand that “Equilibrium with fictive forces, 0 ” is an alterna- • 1.5.2 Me-
tive way to formulate “Equations of motion, ”. chanical/Ma-
• Identify origin of equations: Equilibrium, Compatibility, Constitution and “Algo- chine Engi-
rithms”. neering
• Understand that one can select “Force in elastic part” as state variables, as an • Figure 1-22
alternative to “Position of inertial body”. • Figure 1-23
• Understand the difference between the operating conditions static/low speed,
steady state, stationary oscillating and transient.
• Set up Mathematical models for small systems, given the Physical model.
• Recognize the terms algebraic loop and high index as difficulties that can appear • Figure 1-16
in modelling. • Figure 1-17
370
Modules in the Course
MMF062 Vehicle Dynamics
371
Modules in the Course
MMF062 Vehicle Dynamics
372
Modules in the Course
MMF062 Vehicle Dynamics
373
Modules in the Course
MMF062 Vehicle Dynamics
374
Modules in the Course
MMF062 Vehicle Dynamics
375
Modules in the Course
MMF062 Vehicle Dynamics
376
Modules in the Course
MMF062 Vehicle Dynamics
377
Modules in the Course
MMF062 Vehicle Dynamics
378
Modules in the Course
MMF062 Vehicle Dynamics
379
Modules in the Course
MMF062 Vehicle Dynamics
380
Modules in the Course
MMF062 Vehicle Dynamics
381
Modules in the Course
MMF062 Vehicle Dynamics
382