Vehicle Dynamics - Rill PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 153

VEHICLE DYNAMICS

FACHHOCHSCHULE REGENSBURG
UNIVERSITY OF APPLIED SCIENCES
HOCHSCHULE FR
TECHNIK
WIRTSCHAFT
SOZIALES

LECTURE NOTES
Prof. Dr. Georg Rill

October 2005

download: http://homepages.fh-regensburg.de/%7Erig39165/

Contents

Contents
1

Introduction

1.1

Literature

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2

Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.1

Vehicle Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.2

Driver

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.3

Vehicle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.4

Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2.5

Environment

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3

Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3.1

Reference frames

1.3.2

Toe and camber angle

1.3.3

Design Position of Wheel Rotation Axis

1.3.4

Steering Geometry

. . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3.4.1

Kingpin . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3.4.2

Caster and Kingpin Angle . . . . . . . . . . . . . . . . . .

1.3.4.3

Caster, Steering Oset and Disturbing Force Lever

. . . .

Road

10

2.1

Modeling Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

10

2.2

2.3

Deterministic Proles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

11

2.2.1

Bumps and Potholes

. . . . . . . . . . . . . . . . . . . . . . . . . .

11

2.2.2

Sine Waves

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

2.3.1

Random Proles

Statistical Properties . . . . . . . . . . . . . . . . . . . . . . . . . .

12

2.3.2

Classication of Random Road Proles . . . . . . . . . . . . . . . .

15

2.3.3

Realizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16

2.3.3.1

Sinusoidal Approximation

. . . . . . . . . . . . . . . . . .

16

2.3.3.2

Shaping Filter

. . . . . . . . . . . . . . . . . . . . . . . .

17

2.3.3.3

Two-Dimensional Model . . . . . . . . . . . . . . . . . . .

18

Tire
3.1

3.2

3.3

3.4

19

3.1.1

Tire Development . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19

3.1.2

Tire Composites

19

3.1.3

Tire Forces and Torques

. . . . . . . . . . . . . . . . . . . . . . . .

20

3.1.4

Measuring Tire Forces and Torques . . . . . . . . . . . . . . . . . .

21

Contact Geometry

. . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

3.2.1

Basic Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23

3.2.2

Local Track Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

3.2.3

Tire Deection

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

26

3.2.4

Length of Contact Patch . . . . . . . . . . . . . . . . . . . . . . . .

28

3.2.5

Static Contact Point

. . . . . . . . . . . . . . . . . . . . . . . . . .

29

3.2.6

Contact Point Velocity . . . . . . . . . . . . . . . . . . . . . . . . .

31

3.2.7

Dynamic Rolling Radius

. . . . . . . . . . . . . . . . . . . . . . . .

32

Forces and Torques caused by Pressure Distribution . . . . . . . . . . . . .

34

3.3.1

Wheel Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

34

3.3.2

Tipping Torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

35

3.3.3

Rolling Resistance

. . . . . . . . . . . . . . . . . . . . . . . . . . .

36

. . . . . . . . . . . . . . . . . . . . . . . . . .

37

3.4.1

Friction Forces and Torques

Longitudinal Force and Longitudinal Slip . . . . . . . . . . . . . . .

37

3.4.2

Lateral Slip, Lateral Force and Self Aligning Torque . . . . . . . . .

40

3.4.3

Wheel Load Inuence . . . . . . . . . . . . . . . . . . . . . . . . . .

41

3.4.4

Two-Dimensional Tire Characteristics . . . . . . . . . . . . . . . . .

43

3.4.5

Dierent Friction Coecients

. . . . . . . . . . . . . . . . . . . . .

45

3.4.6

Self Aligning Torque

. . . . . . . . . . . . . . . . . . . . . . . . . .

46

3.4.7

Camber Inuence . . . . . . . . . . . . . . . . . . . . . . . . . . . .

48

3.4.8

Bore Torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

3.4.9

Typical Tire Characteristics

52

. . . . . . . . . . . . . . . . . . . . . .

Suspension System

54

4.1

Purpose and Components

. . . . . . . . . . . . . . . . . . . . . . . . . . .

54

4.2

Some Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

4.2.1

Multi Purpose Systems . . . . . . . . . . . . . . . . . . . . . . . . .

55

4.2.2

Specic Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

56

4.3

4.4

II

19
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Steering Systems

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

56

4.3.1

Requirements

4.3.2

Rack and Pinion Steering

56

4.3.3

Lever Arm Steering System

4.3.4
4.3.5

Bus Steer System . . . . . . . . . . . . . . . . . . . . . . . . . . . .

58

. . . . . . . . . . . . . . . . . . . . . . .

57

. . . . . . . . . . . . . . . . . . . . . .

57

Drag Link Steering System . . . . . . . . . . . . . . . . . . . . . . .

58

Standard Force Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . .

59

4.4.1

Springs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

59

4.4.2

Damper

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

60

4.4.3

Rubber Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . .

62

4.5

63

Testing and Evaluating Procedures

. . . . . . . . . . . . . . . . . .

63

4.5.2

Simple Spring Damper Combination

. . . . . . . . . . . . . . . . .

66

4.5.3

General Dynamic Force Model . . . . . . . . . . . . . . . . . . . . .

68

4.5.3.1

69

Hydro-Mount . . . . . . . . . . . . . . . . . . . . . . . . .

Vertical Dynamics

72

5.1

Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

72

5.2

Modelling Aspects

72

5.3

5.4

5.5

Dynamic Force Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . .


4.5.1

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.2.1

Full Vehicle Model

5.2.2

Twodimensional Models

. . . . . . . . . . . . . . . . . . . . . . . .

73

5.2.3

Simple Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

75

Basic Tuning

. . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.3.1

Natural Frequency and Damping Rate

5.3.2

Spring Rates

6.2

6.3

6.4

76

. . . . . . . . . . . . . . . .

76

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

78

5.3.2.1

Minimum Spring Rates . . . . . . . . . . . . . . . . . . . .

78

5.3.2.2

Nonlinear Springs . . . . . . . . . . . . . . . . . . . . . . .

79

5.3.3

Inuence of Damping . . . . . . . . . . . . . . . . . . . . . . . . . .

81

5.3.4

Optimal Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . .

81

5.3.4.1

Avoiding Overshoots . . . . . . . . . . . . . . . . . . . . .

81

5.3.4.2

Fast Approach to Steady State

. . . . . . . . . . . . . . .

82

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

87

Sky Hook Damper


5.4.1

Modelling Aspects

. . . . . . . . . . . . . . . . . . . . . . . . . . .

87

5.4.2

Eigenfrequencies and Damping Ratios . . . . . . . . . . . . . . . . .

88

5.4.3

Technical Realization . . . . . . . . . . . . . . . . . . . . . . . . . .

90

Nonlinear Force Elements

. . . . . . . . . . . . . . . . . . . . . . . . . . .

90

5.5.1

Quarter Car Model . . . . . . . . . . . . . . . . . . . . . . . . . . .

90

5.5.2

Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

92

Longitudinal Dynamics
6.1

72

94
. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

94

6.1.1

Dynamic Wheel Loads

Simple Vehicle Model . . . . . . . . . . . . . . . . . . . . . . . . . .

94

6.1.2

Inuence of Grade

. . . . . . . . . . . . . . . . . . . . . . . . . . .

95

6.1.3

Aerodynamic Forces

. . . . . . . . . . . . . . . . . . . . . . . . . .

96

Maximum Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97

6.2.1

Tilting Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97

6.2.2

Friction Limits

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

97

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

98

6.3.1

Single Axle Drive . . . . . . . . . . . . . . . . . . . . . . . . . . . .

98

6.3.2

Braking at Single Axle . . . . . . . . . . . . . . . . . . . . . . . . .

99

6.3.3

Optimal Distribution of Drive and Brake Forces . . . . . . . . . . . 100

6.3.4

Dierent Distributions of Brake Forces

6.3.5

Anti-Lock-Systems

Driving and Braking

Drive and Brake Pitch

. . . . . . . . . . . . . . . . 102

. . . . . . . . . . . . . . . . . . . . . . . . . . . 102

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

III

6.4.1

Vehicle Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

6.4.2

Equations of Motion

6.4.3

Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

6.4.4

Driving and Braking

6.4.5

Brake Pitch Pole

7.2

7.3

109

Kinematic Approach

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

7.1.1

Kinematic Tire Model

. . . . . . . . . . . . . . . . . . . . . . . . . 109

7.1.2

Ackermann Geometry

. . . . . . . . . . . . . . . . . . . . . . . . . 109

7.1.3

Space Requirement . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

7.1.4

Vehicle Model with Trailer . . . . . . . . . . . . . . . . . . . . . . . 112


7.1.4.1

Kinematics

7.1.4.2

Vehicle Motion

. . . . . . . . . . . . . . . . . . . . . . . . . . 112

7.1.4.3

Entering a Curve . . . . . . . . . . . . . . . . . . . . . . . 114

7.1.4.4

Trailer Motions . . . . . . . . . . . . . . . . . . . . . . . . 115

7.1.4.5

Course Calculations

. . . . . . . . . . . . . . . . . . . . . 116

Steady State Cornering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117


Cornering Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . 117

7.2.2

Overturning Limit

7.2.3

Roll Support and Camber Compensation . . . . . . . . . . . . . . . 121

7.2.4

Roll Center and Roll Axis

7.2.5

Wheel Loads

. . . . . . . . . . . . . . . . . . . . . . . . . . . 118
. . . . . . . . . . . . . . . . . . . . . . . 124

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

Simple Handling Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125


7.3.1

Modeling Concept

7.3.2

Kinematics

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

7.3.3

Tire Forces

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

7.3.4

Lateral Slips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

7.3.5

Equations of Motion

7.3.6

Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

7.3.8

. . . . . . . . . . . . . . . . . . . . . . . . . . . 125

. . . . . . . . . . . . . . . . . . . . . . . . . . 127

7.3.6.1

Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . 128

7.3.6.2

Low Speed Approximation . . . . . . . . . . . . . . . . . . 129

7.3.6.3

High Speed Approximation

. . . . . . . . . . . . . . . . . 129

Steady State Solution . . . . . . . . . . . . . . . . . . . . . . . . . . 130


7.3.7.1

Side Slip Angle and Yaw Velocity . . . . . . . . . . . . . . 130

7.3.7.2

Steering Tendency

7.3.7.3

Slip Angles

. . . . . . . . . . . . . . . . . . . . . . 132

. . . . . . . . . . . . . . . . . . . . . . . . . . 133

Inuence of Wheel Load on Cornering Stiness

Driving Behavior of Single Vehicles


8.1

. . . . . . . . . . . . . . . . . . . . . . . . 113

7.2.1

7.3.7

IV

. . . . . . . . . . . . . . . . . . . . . . . . . . 107

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

Lateral Dynamics
7.1

. . . . . . . . . . . . . . . . . . . . . . . . . . 105

. . . . . . . . . . . 133

136

Standard Driving Maneuvers . . . . . . . . . . . . . . . . . . . . . . . . . . 136


8.1.1

Steady State Cornering . . . . . . . . . . . . . . . . . . . . . . . . . 136

8.1.2

Step Steer Input

8.1.3

Driving Straight Ahead . . . . . . . . . . . . . . . . . . . . . . . . . 138

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

8.2

8.3

8.1.3.1

Random Road Prole

8.1.3.2

Steering Activity . . . . . . . . . . . . . . . . . . . . . . . 140

. . . . . . . . . . . . . . . . . . . . 138

Coach with dierent Loading Conditions

. . . . . . . . . . . . . . . . . . . 140

8.2.1

Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

8.2.2

Roll Steering

8.2.3

Steady State Cornering . . . . . . . . . . . . . . . . . . . . . . . . . 141

8.2.4

Step Steer Input

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

Dierent Rear Axle Concepts for a Passenger Car

. . . . . . . . . . . . . . 143

1 Introduction
1.1 Literature

ATZ: Automobiltechnische Zeitschrift


Fachbuchgruppe Fahrwerktechnik:

Jrnsen Reimpell, Hrsg. Vogel Buchverlag Wrzburg.

Dynamik der Kraftfahrzeuge: M. Mitschke, Bde. A,B,C; Springer-Verlag.


Mitschke,

M.;

Wallentowitz,

H.: Dynamik der Kraftfahrzeuge. 4. Auage.

Springer-Verlag Berlin Heidelberg 2004.

Popp, K.; Schiehlen, W.: Fahrzeugdynamik. Teubner Stuttgart 1993.

Simulation von Kraftfahrzeugen: G. Rill, Vieweg-Verlag 1994.

Radfhrungen der Straenfahrzeuge: W. Matschinsky, Springer-Verlag.

Blundell, M.; Harty, D.: The Multibody System Approach to Vehicle Dynamics.

Elsevier Butterworth-Heinemann Publications, 2004.

Fundamentals of Vehicle Dynamics: Th., D. Gillespie, SAE, Inc.

ISO-Standards: (International Organisation for Standardization.)

z.B.: ISO 4138 Steady State Circular Test Procedure.

Kraftfahrtechnisches

Handbuch:

Robert

Bosch

GmbH

(Hrsg.),

23.

Au.,

Vieweg-Verlag.

Proceedings:

VDI-Tagungen: z.B.: Berechnung im Automobilbau.


SAE-Congress:
FISITA:
IAVSD:

(Society of Automotive Engineers).

( Fd. Internat. des Socits d'Ingnieurs de Techniques de l'Automobile).


(International Assosiation for Vehicle System Dynamics).

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

1.2 Terminology
1.2.1 Vehicle Dynamics
The expression `Vehicle Dynamics' encompasses the interaction of:

driver

vehicle

load

environment

Vehicle dynamics mainly deals with:

the improvement of active safety and driving comfort

the reduction of road destruction

In vehicle dynamics are employed:

computer calculations

test rig measurements

eld tests

In the following the interactions between the single systems and the problems with computer calculations and/or measurements shall be discussed.

1.2.2 Driver
By various means the driver can interfere with the vehicle:

driver

steering wheel
accelerator pedal
brake pedal
clutch
gear shift

lateral dynamics

longitudinal dynamics

vehicle

The vehicle provides the driver with these information:

vehicle

vibrations:
sounds:

motor, aerodynamics, tires

instruments:

velocity, external temperature, ...

longitudinal, lateral, vertical

driver

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

The environment also inuences the driver:

environment

climate

track

trac density

driver

The driver's reaction is very complex. To achieve objective results, an `ideal' driver is
used in computer simulations, and in driving experiments automated drivers (e.g. steering
machines) are employed.
Transferring results to normal drivers is often dicult, if eld tests are made with test
drivers. Field tests with normal drivers have to be evaluated statistically. Of course, the
driver's security must have absolute priority in all tests.
Driving simulators provide an excellent means of analyzing the behavior of drivers even
in limit situations without danger.
It has been tried to analyze the interaction between driver and vehicle with complex driver
models for some years.

1.2.3 Vehicle
The following vehicles are listed in the ISO 3833 directive:

motorcycles

passenger cars

busses

trucks

agricultural tractors

passenger cars with trailer

truck trailer / semitrailer

road trains

For computer calculations these vehicles have to be depicted in mathematically describable


substitute systems. The generation of the equations of motion, the numeric solution, as
well as the acquisition of data require great expenses. In times of PCs and workstations
computing costs hardly matter anymore.
At an early stage of development, often only prototypes are available for eld and/or
laboratory tests. Results can be falsied by safety devices, e.g. jockey wheels on trucks.

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

1.2.4 Load
Trucks are conceived for taking up load. Thus, their driving behavior changes.


Load

mass, inertia, center of gravity


dynamic behaviour (liquid load)

In computer calculations problems occur at the determination of the inertias and the
modeling of liquid loads.
Even the loading and unloading process of experimental vehicles takes some eort. When
carrying out experiments with tank trucks, ammable liquids have to be substituted with
water. Thus, the results achieved cannot be simply transferred to real loads.

1.2.5 Environment
The environment inuences primarily the vehicle:


Environment

road:

irregularities, coecient of friction

air:

resistance, cross wind


vehicle

but also aects the driver:


environment

climate
visibility


driver

Through the interactions between vehicle and road, roads can quickly be destroyed.
The greatest diculty with eld tests and laboratory experiments is the virtual impossibility of reproducing environmental inuences.
The main problems with computer simulation are the description of random road irregularities and the interaction of tires and road as well as the calculation of aerodynamic
forces and torques.

1.3 Definitions
1.3.1 Reference frames
A reference frame xed to the vehicle and a ground-xed reference frame are used to
describe the overall motions of the vehicle, Figure 1.1. The ground-xed reference frame

x0 , y0 , z0 serves as an inertial reference frame. Within the vehicle-xed


reference frame the xF -axis points forward, the yF -axis to the left, and the zF -axis upward.
with the axis

The wheel rotates around an axis which is xed to the wheel carrier. The reference frame

is xed to the wheel carrier. In design position its axes

the corresponding axis of vehicle-xed reference frame

F.

xC , yC

and

zC

are parallel to

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

zF
yF
xF
z0
zC

x0

yC
en

xC

y0

eyR

Figure 1.1: Frames used in vehicle dynamics

The momentary position of the wheel is xed by the wheel center and the orientation of
the wheel rim center plane which is dened by the unit vector

eyR

into the direction of

the wheel rotation axis.


Finally, the normal vector

en

describes the inclination of the local track plane.

1.3.2 Toe and camber angle


front

xF

left
wheel

yF

vehicle
center
plane

right
wheel

rear
Figure 1.2: Positive toe-in angle
According to the DIN 70 000 directive the angle

between the vehicle center plane in

longitudinal direction and the intersection line of the tire center plane with the track
plane is named toe or toe-in angle. It will be positive, if the front part of the wheel is
oriented towards the vehicle center plane, Figure 1.2. Toe-in reduces the tendency of the
wheels to shimmy.
The camber angle

en .

is the angle between the wheel center plane and the local track normal

It will be positive, if the upper part of the wheel is inclined outwards, Figure 1.3. A

cambered wheel causes a non symmetric tire wear.

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

en

en
left
wheel

right
wheel

zF

yF
road

Figure 1.3: Positive camber angle

1.3.3 Design Position of Wheel Rotation Axis


The unit vector

eyR

describes the wheel rotation axis. Its orientation with respect to the

wheel carrier xed reference frame can be dened by the angles 0 and 0 or 0 and 0 ,
Fig. 1.4. In design position the corresponding axes of the frames

and

are parallel.

zC = zF

xC = xF
0
0

0*

yC = yF

eyR

Figure 1.4: Design position of wheel rotation axis


Then, for the left wheel we get

eyR,F = eyR,C

tan 0
1

1
= q
tan2 0 + 1 + tan2 0
tan 0

or

eyR,F = eyR,C

(1.2)

yF -axis and the projection line of the wheel rotation


0 describes the angle between the yF -axis and the
0
projection line of the wheel rotation axis into the yF - zF -plane, whereas 0 is the angle
between the wheel rotation axis eyR and its projection into the xF - yF -plane. Kinematics

where

sin 0 cos 0
= cos 0 cos 0 ,
sin 0

(1.1)

is the angle between the

axis into the

xF - yF -plane,

the angle

FH Regensburg, University of Applied Sciences

and compliance test machines usually measure the angle


industry mostly uses this angle instead of

That is why, the automotive

en

points into the direction of

correspond with the toe angle and the

camber angle 0 . To specify the dierence between 0 and 0 the ratio between the third
the vertical axes

the angles

0 .

0 .

On a at and horizontal road where the track normal

zC = zF

Prof. Dr.-Ing. G. Rill

and second component of the unit vector


deliver

and

eyR

is considered. The Equations 1.1 and 1.2

sin 0
tan 0
=
1
cos 0 cos 0

Hence, for small angles

0  1

or

tan 0 =

tan 0
.
cos 0

the dierence between the angles

(1.3)
and

is hardly

noticeable.

1.3.4 Steering Geometry


1.3.4.1 Kingpin
At the steered front axle, the McPherson-damper strut axis, the double wishbone axis,
and the multi-link wheel suspension or the enhanced double wishbone axis are mostly
used in passenger cars, Fig. 1.5 and Fig. 1.6.

zR
B

yR
M

xR

kingpin axis A-B


Figure 1.5: Double wishbone wheel suspension
The wheel body rotates around the kingpin at steering motions. At the double wishbone

B , which determine the kingpin, are both xed to the wheel


A is still xed to the wheel body at the standard McPherson
wheel suspension, the ball joint B is now xed to the vehicle body. At a multi-link axle the

axis the ball joints

and

body. Whereas the ball joint

kingpin is no longer dened by real joints. Here, as well as with the enhanced McPherson
wheel suspension, the kingpin changes its position relative to the wheel body at wheel
travel and steering motions.

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

B
zR

zR

yR

yR
xR

xR

M
A

rotation axis

kingpin axis A-B

Figure 1.6: McPherson and multi-link wheel suspensions

1.3.4.2 Caster and Kingpin Angle

eS describes the direction of the kingpin axis. Within the vehicle xed
reference frame F it can be xed by two angles. The caster angle denotes the angle
between the zF -axis and the projection line of eS into the xF -, zF -plane. In a similar
way the projection of eS into the yF -, zF -plane delivers the kingpin inclination angle ,

The unit vector

Fig. 1.7.

eS

zF

zF

yF
xF

Figure 1.7: Kingpin and caster angle


At many axles the kingpin and caster angle can no longer be determined directly. Here, the
current rotation axis at steering motions, which can be taken from kinematic calculations
will deliver a virtual kingpin. The current values of the caster angle
inclination angle

and the kingpin

can be calculated from the components of the unit vector

eS

in the

direction of the kingpin, described in the vehicle xed reference frame

(1)

tan =

eS,F
(3)

eS,F
where

(1)

(2)

(3)

eS,F , eS,F , eS,F

xed reference frame

(2)

and

tan =

eS,F

are the components of the unit vector

F.

(1.4)

eS,F

expressed in the vehicle

(3)

eS,F

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

1.3.4.3 Caster, Steering Offset and Disturbing Force Lever


The contact point

P,

the local track normal

en

and the unit vectors

ex

and

ey

which

point into the direction of the longitudinal and lateral tire force result from the contact
geometry. The axle kinematics denes the kingpin line. In general, the point

where an

extension oft the kingpin line meets the road surface does not coincide with the contact
point

P,

Fig. 1.8. As both points are located on the local track plane, for the left wheel

the vector from

to

can be written as

rSP = c ex + s ey ,
where

names the caster and

positive, if

is the steering oset.

is located in front of and inwards of

(1.5)
Caster and steering oset will be

P.
kingpin line

C d
ey

en
P

ex
S

Figure 1.8: Caster and steering oset


The distance

d between the wheel center C

and the king pin line represents the disturbing

force lever. It is an important quantity in evaluating the overall steering behavior.

2 Road
2.1 Modeling Aspects
Sophisticated road models provide the road height
at each point

x, y ,

zR

and the local friction coecient

Fig. 2.1.
Road profile

z0
y0

Segments

z(x,y)
x0

(x,y)

Friction

Obstacle

Center Line L(s)

Figure 2.1: Sophisticated road model


The tire model is then responsible to calculate the local road inclination. By separating
the horizontal course description from the vertical layout and the surface properties of
the roadway almost arbitrary road layouts are possible.
Besides single obstacles or track grooves the irregularities of a road are of stochastic
nature. A vehicle driving over a random road prole mainly performs hub, pitch and roll
motions. The local inclination of the road prole also induces longitudinal and lateral
motions as well as yaw motions. On normal roads the latter motions have less inuence
on ride comfort and ride safety. To limit the eort of the stochastic description usually
simpler road models are used.
If the vehicle drives along a given path its momentary position can be described by the
path variable

s = s(t).

Hence, a fully two-dimensional road model can be reduced to a

parallel track model, Fig. 2.2.

10

FH Regensburg, University of Applied Sciences

z
x

Prof. Dr.-Ing. G. Rill

z2

z1

zR(x,y)
z1(s)

Figure 2.2: Parallel track road model

Now, the road heights on the left and right track are provided by two one-dimensional
functions

z1 = z1 (s)

z2 = z2 (s).

and

Within the parallel track model no information

about the local lateral road inclination is available. If this information is not provided by
additional functions the impact of a local lateral road inclination to vehicle motions is not
taken into account.
For basic studies the irregularities at the left and the right track can considered to be

z1 (s) z2 (s).

approximately the same,

z1 (x) = z2 (x)

Then, a single track road model with

zR (s) =

can be used. Now, the roll excitation of the vehicle is neglected too.

2.2 Deterministic Profiles


2.2.1 Bumps and Potholes
Bumps and Potholes on the road are single obstacles of nearly arbitrary shape. Already
with simple rectangular cleats the dynamic reaction of a vehicle or a single tire to a
sudden impact can be investigated. If the shape of the obstacle is approximated by a
smooth function, like a cosine wave, then, discontinuities will be avoided. Usually the
obstacles are described in local reference frames, Fig. 2.3.

H
B

y
L

Figure 2.3: Rectangular cleat and cosine-shaped bump


Then, the rectangular cleat is simply dened by

(
z(x, y) =

H
0

if
else

0<x<L

and

12 B < y < 12 B

(2.1)

11

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

and the cosine-shaped bump is given by

z(x, y) =

1
2

where

H, B

if negative

x 
H 1 cos 2
L
0


0<x<L

if

and

21 B < y < 12 B

(2.2)

else

L denote height, width and length of the obstacle. Potholes are obtained
values for the height (H < 0) are used.
and

2.2.2 Sine Waves


Using the parallel track road model, a periodic excitation can be realized by

z2 (s) = A sin ( s ) ,

z1 (s) = A sin ( s) ,

(2.3)

s is the path variable, A denotes the amplitude, the wave number, and the angle
describes a phase lag between the left and the right track. The special cases = 0 and
= represent the in-phase excitation with z1 = z2 and the out of phase excitation with
z1 = z2 .
where

ds/dt = v0 , the momentary position of the


initial position s = 0 at t = 0 was assumed. By

If the vehicle runs with constant velocity


vehicle is given by

s = v0 t,

where the

introducing the wavelength

(2.4)

2
2
v0
s=
v0 t = 2 t = t .
L
L
L

(2.5)

L =
the term

can be written as

s =

Hence, in the time domain the excitation frequency is given by

f = /(2) = v0 /L.

0.5 Hz to 15 Hz . This
v0 /L 0.5 Hz and v0 /L 15 Hz .

For most of the vehicles the rigid body vibrations are in between
range is covered by waves which satisfy the conditions

L = 4 m, the rigid body vibration of a vehicle are excited


min
if the velocity of the vehicle will be varied from v0
= 0.5 Hz 4 m = 2 m/s = 7.2 km/h
max
= 15 Hz 4 m = 60 m/s = 216 km/h. Hence, to achieve an excitation in the
to v0
For a given wavelength, lets say

whole frequency range with moderate vehicle velocities proles with dierent varying
wavelengths are needed.

2.3 Random Profiles


2.3.1 Statistical Properties
Road proles t the category of stationary Gaussian random processes. Hence, the irregularities of a road can be described either by the prole itself
properties, Fig. 2.4.

12

zR = zR (s) or by its statistical

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

zR
0.15

Gaussian
density
function

Realization

0.10

[m]
0.05

0
-0.05

-0.10
-0.15
-200

Histogram
s

-100

-150

-50

50

200

[m] 150

100

Figure 2.4: Road prole and statistical properties

By choosing an appropriate reference frame, a vanishing mean value

1
m = E {zR (s)} = lim
X X

ZX/2
zR (s) ds = 0

(2.6)

X/2
can be achieved, where

E {} denotes the expectation operator. Then, the Gaussian density

function which corresponds with the histogram is given by

p(zR ) =

where the deviation or the eective value

zR2

e 2 2 ,

(2.7)

is obtained from the variance of the process

zR = zR (s)
ZX/2


1
2 = E zR2 (s) = lim
X X


zR (s)2 ds .

(2.8)

X/2
Alteration of

eects the shape of the density function. In particular, the points of

inexion occur at

The probability of a value

P () =

|z| <

is given by

Z+ z 2

e 2 2 dz .

P () = 0.683, P (2) = 0.955,


|z| 3 is 0.3%.

In particular, one gets the values:


Hence, the probability of a value

(2.9)

and

P (3) = 0.997.

In extension to the variance of a random process the auto-correlation function is dened


by

1
R() = E {zR (s) zR (s+)} = lim
X X

ZX/2
zR (s) zR (s+) ds .

(2.10)

X/2

13

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

R() = R(),

The auto-correlation function is symmetric,

and it plays an important

part in the stochastic analysis. In any normal random process, as

increases the link

zR (s) and zR (s+) diminishes. For large values of the two values are practically
unrelated. Hence, R( ) will tend to 0. In fact, R() is always less R(0), which
2
coincides with the variance of the process. If a periodic term is present in the process
it will show up in R().
between

Usually, road proles are characterized in the frequency domain. Here, the auto-correlation
function

S()

R()

is replaced by the power spectral density (psd)

S().

In general,

R()

and

are related to each other by the Fourier transformation

1
S() =
2

R() e

1
R() =
2

and

where

S() ei d ,

(2.11)

i is the imaginary unit, and in rad/m denotes the wave number. To avoid negative

wave numbers, usually a one-sided psd is dened. With

() = 2 S() ,
the relationship

if

() = 0 ,

and

ei = cos() i sin(),

if

<0,

and the symmetry property

(2.12)

R() = R()

Eq. (2.11) results in

2
() =

Z
R() cos () d

and

Z
R() = () cos () d .

(2.13)

Now, the variance is obtained from

Z
= R( = 0) = () d .
2

(2.14)

0
In reality the psd

()

will be given in a nite interval

1 N ,

Fig. 2.5. Then,

(i)

Figure 2.5: Power spectral density in a nite interval


Eq. (2.14) can be approximated by a sum, which for

N
X
i=1

14

(i ) 4

with

equal intervals will result in

4 =

N 1
.
N

(2.15)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

2.3.2 Classification of Random Road Profiles


Road elevation proles can be measured point by point or by high-speed prolometers.
The power spectral densities of roads show a characteristic drop in magnitude with the
wave number, Fig. 2.6a. This simply reects the fact that the irregularities of the road may
amount to several meters over the length of hundreds of meters, whereas those measured
over the length of one meter are normally only some centimeter in amplitude.
Random road proles can be approximated by a psd in the form of


() = (0 )
where,

= 2/L

in

rad/m

w

describes the value of the psd at a the reference wave

Power spectral density [m2/(rad/m)]

(2.16)

0 = (0 ) in m2 /(rad/m)
number 0 = 1 rad/m. The drop

denotes the wave number and

in magnitude is modeled by the waviness

10-3

w.

a) Measurements (country road)

b) Range of road classes (ISO 8608)


0=256106

10-4
10-5
Class E
0=1106

10-6
10-7
10-8
10-9 -2
10

Class A
10-1
100
101
Wave number [rad/m]

102

10-2

10-1
100
101
Wave number [rad/m]

102

Figure 2.6: Road power spectral densities: a) Measurements, b) Classication


According to the international directive ISO 8608 typical road proles can be grouped
into classes from A to E. By setting the waviness to w = 2 each class is simply dened by
6
its reference value 0 . Class A with 0 = 1 10
m2 /(rad/m) characterizes very smooth
6
highways, whereas Class E with 0 = 256 10
m2 /(rad/m) represents rather rough
roads, Fig. 2.6b.

15

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

2.3.3 Realizations
2.3.3.1 Sinusoidal Approximation
A random prole of a single track can be approximated by a superposition of
sine waves

zR (s) =

N
X

Ai sin (i s i ) ,

(2.17)

i=1

Ai and its
i , i = 1(1)N

i .

where each sine wave is determined by its amplitude

wave number

dierent sets of uniformly distributed phase angles

in the range between

and

By

dierent proles can be generated which are similar in the general appearance

but dierent in details.


The variance of the sinusoidal representation is then given by

1
2 = lim
X X

ZX/2 X
N
X/2

For

i=j

and for

i 6= j

!
Ai sin (i s i )

i=1

N
X

!
Aj sin (j s j )

ds .

(2.18)

j=1

dierent types of integrals are obtained. The ones for

i=j

can

be solved immediately

Z
Jii =

A2i




1
A2i
i si sin 2 (i si ) .
sin (i si ) ds =
2i
2
2

(2.19)

Using the trigonometric relationship

sin x sin y =
the integrals for

i 6= j

1
1
cos(xy) cos(x+y)
2
2

(2.20)

can be solved too

Z
Jij =

Ai sin (i si ) Aj sin (j sj ) ds

1
= A i Aj
2
=

1
cos (ij s ij ) ds Ai Aj
2

Z
cos (i+j s i+j ) ds

(2.21)

1 Ai Aj
1 A i Aj
sin (ij s ij ) +
sin (i+j s i+j )
2 ij
2 i+j

where the abbreviations

ij = ij

and

ij = ij

terms in Eqs. (2.19) and (2.21) are limited to values of

were used. The sine and cosine

1.

Hence, Eq. (2.18) simply

results in

N
N
N
1 X   X/2
1 X   X/2
1X 2
Jii X/2 + lim
Jij X/2 =
= lim
A .
X X
X X
2 i=1 i
i=1
i,j=1
|
{z
}
|
{z
}
N
2
0
X A
i
i
2
i
i=1
2

16

(2.22)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

On the other hand, the variance of a sinusoidal approximation to a random road prole
is given by Eq. (2.15). So, a road prole
given psd

()

described by Eq. (2.17) will have a

if the amplitudes are generated according to

Ai =
and the wave numbers

2 (i ) 4 ,

are chosen to lie at

i = 1(1)N ,
equal intervals

(2.23)

4.

Road profile z=z(s)

0.10

[m]

zR = zR (s)

0.05
0
-0.05
-0.10
0

10

20

30

40

50

60

70

80 [m] 90

100

Figure 2.7: Realization of a country road

0 = 10 106 m2 /(rad/m) is shown in


Fig. 2.7. According to Eq. (2.17) the prole z = z(s) was generated by N = 200 sine waves
in the frequency range from 1 = 0.0628 rad/m to N = 62.83 rad/m. The amplitudes
Ai , i = 1(1)N were calculated by Eq. (2.23) and the MATLABr function rand was used
to produce uniformly distributed random phase angles in the range between 0 and 2 .

A realization of the country road with a psd of

2.3.3.2 Shaping Filter


The white noise process produced by random number generators has a uniform spectral
density, and is therefore not suitable to describe real road proles. But, if the white noise
process is used as input to a shaping lter more appropriate spectral densities will be
obtained. A simple rst order shaping lter for the road prole

where

is a constant, and

d
zR (s) = zR (s) + w(s) ,
ds
w(s) is a white noise process with

zR

reads as
(2.24)

the spectral density

w .

Then, the spectral density of the road prole is obtained from

R = H() W H T () =
where

is the wave number, and

H()

1
W
1
W
= 2
,
+ i
i
+ 2

(2.25)

is the frequency response function of the shaping

lter.
By setting

W = 10 106 m2 /(rad/m) and = 0.01 rad/m the measured psd of a typical

country road can be approximated very well, Fig. 2.8.


The shape lter approach is also suitable for modeling parallel tracks. Here, the crosscorrelation between the irregularities of the left and right track have to be taken into
account too.

17

FH Regensburg, University of Applied Sciences

Power spectral density [m2/(rad/m)]

Vehicle Dynamics

10-3
Measurements
Shaping filter

10-4
10-5
10-6
10-7
10-8
10-9 -2
10

10-1
100
101
Wave number [rad/m]

102

Figure 2.8: Shaping lter as approximation to measured psd

2.3.3.3 Two-Dimensional Model


The generation of fully two-dimensional road proles

zR = zR (x, y)

via a sinusoidal ap-

proximation is very laborious. Because a shaping lter is a dynamic system, the resulting road prole realizations are not reproducible. By adding band-limited white noise
processes and taking the momentary position

x, y

as seed for the random number gener-

ator a reproducible road prole can be generated.

z
x

1
0
-1
4

0
-2
-4

10

15

20

25

30

35

40

45

50

Figure 2.9: Two-dimensional road prole


By assuming the same statistical properties in longitudinal and lateral direction twodimensional proles, like the one in Fig. 2.9, can be obtained.

18

3 Tire
3.1 Introduction
3.1.1 Tire Development
Some important mile stones in the development of tires are shown in Table 3.1.
1839
1845

Charles Goodyear: vulcanization


Robert William Thompson: rst pneumatic tire
(several thin inated tubes inside a leather cover)

1888

John Boyd Dunlop: patent for bicycle (pneumatic) tires

1893

The Dunlop Pneumatic and Tyre Co. GmbH, Hanau, Germany

1895

Andr and Edouard Michelin: pneumatic tires for Peugeot


Paris-Bordeaux-Paris (720 Miles):

50 tire deations,
22 complete inner tube changes

1899

Continental: long-lived tires (approx. 500 Kilometer)

1904

Carbon added: black tires.

1908

Frank Seiberling: grooved tires with improved road traction

1922

Dunlop: steel cord thread in the tire bead

1943

Continental: patent for tubeless tires

1946

Radial Tire
Table 3.1: Milestones in tire development

Of course the tire development did not stop in 1946, but modern tires are still based on
this achievements.

3.1.2 Tire Composites


Tires are very complex. They combine dozens of components that must be formed, assembled and cured together. And their ultimate success depends on their ability to blend
all of the separate components into a cohesive product that satises the driver's needs. A
modern tire is a mixture of steel, fabric, and rubber. The main composites of a passenger
car tire with an overall mass of

8.5 kg

are listed in Table 3.2.

19

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Reinforcements: steel, rayon, nylon

16%

Rubber: natural/synthetic

38%

Compounds: carbon, silica, chalk, ...

30%

Softener: oil, resin

10%

Vulcanization: sulfur, zinc oxide, ...

4%

Miscellaneous

2%

Table 3.2: Tire composites: 195/65 R 15 ContiEcoContact, data from

www.felge.de

3.1.3 Tire Forces and Torques


In any point of contact between the tire and the road surface normal and friction forces
are transmitted. According to the tire's prole design the contact patch forms a not
necessarily coherent area, Figure 3.1.

140 mm

180 mm

Figure 3.1: Tire footprint of a passenger car at normal loading condition:


Continental 205/55 R16 90 H,

2.5

bar,

Fz = 4700 N

The eect of the contact forces can be fully described by a resulting force vector applied
at a specic point of the contact patch and a torque vector. The vectors are described in a
track-xed reference frame. The
to the

z -axis

z -axis

is normal to the track, the

and perpendicular to the wheel rotation axis

right-handed reference frame also xes the

eyR .

x-axis

is perpendicular

Then, the demand for a

y -axis.

The components of the contact force vector are named according to the direction of the
axes, Figure 3.2.
A non symmetric distribution of the forces in the contact patch causes torques around
the

and

axes. A cambered tire generates a tilting torque

Tx .

The torque

the rolling resistance of the tire. In particular, the torque around the

z -axis

Ty

includes

is important

in vehicle dynamics. It consists of two parts,

Tz = TB + TS .

20

(3.1)

FH Regensburg, University of Applied Sciences

Fx
Fy
Fz

longitudinal force

Tx
Ty
Tz

tilting torque

Prof. Dr.-Ing. G. Rill

lateral force
vertical force or wheel load

eyR
Fy

rolling resistance torque

Tx
Fx

self aligning and bore torque

Tz

Ty

Fz

Figure 3.2: Contact forces and torques

The rotation of the tire around the


torque

MS

z -axis

causes the bore torque

MB .

The self aligning

takes into account that ,in general, the resulting lateral force is not acting in

the center of the contact patch.

3.1.4 Measuring Tire Forces and Torques


To measure tire forces and torques on the road a special test trailer is needed, Figure
3.4. Here, the measurements are performed under real operating conditions. Arbitrary

Test trailer

tire
exact contact
real road

compensation wheel

test wheel

Figure 3.3: Layout of a tire test trailer


surfaces like asphalt or concrete and dierent environmental conditions like dry, wet or
icy are possible. Measurements with test trailers are quite cumbersome and in general
they are restricted to passenger car tires.
Indoor measurements of tire forces and torques can be performed on drums or on a at
bed, Figure 3.4.

21

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

tire

too small
contact area

safety walk
coating

tire

safety walk coating

perfect contact

rotation
drum

tire

too large contact area

Figure 3.4: Drum and at bed tire test rig

On drum test rigs the tire is placed either inside or outside of the drum. In both cases
the shape of the contact area between tire and drum is not correct. That is why, one
can not rely on the measured self aligning torque. Due its simple and robust design, wide
applications including measurements of truck tires are possible.
The at bed tire test rig is more sophisticated. Here, the contact patch is as at as on the
road. But, the safety walk coating which is attached to the steel bed does not generate
the same friction conditions as on a real road surface.
Radial 205/50 R15, FN= 3500 N, dry asphalt
4000

Longitud force Fx [N]

3000
2000

Driving

1000
0
-1000

Braking

-2000
-3000
-4000
-40

-30

-20

-10

10

20

30

40

Longitudinal slip [%]

Figure 3.5: Typical results of tire measurements

Tire forces and torques are measured in quasi-static operating conditions. Hence, the measurements for increasing and decreasing the sliding conditions usually result in dierent
graphs, Figure 3.5. In general, the mean values are taken as steady state results.

22

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

3.2 Contact Geometry


3.2.1 Basic Approach
x0 -, y0 - z0 -system

The current position of a wheel in relation to the xed


wheel center

and the unit vector

eyR

is given by the

in the direction of the wheel rotation axis, Figure

3.6.

rim
centre
plane e zR

tire

e yR

rMP

e yR

en

wheel
carrier

ex
b
P0 a
ey P

en
P0
z0

P*

local road plane

x0
road: z = z ( x , y )

x0
y0

y0

z0

Figure 3.6: Contact geometry

The irregularities of the track can be described by an arbitrary function of two spatial
coordinates

z = z(x, y).
At an uneven track the contact point

(3.2)

can not be calculated directly. At rst, one can

get an estimated value with the vector

rM P = r0 ezB ,
where

r0

is the undeformed tire radius, and

ezB

(3.3)

is the unit vector in the

z -direction of the

body xed reference frame.


The position of this rst guess

z0

with respect to the earth xed reference frame

x0 , y0 ,

is determined by

x
= y ,
z

r0P ,0 = r0M,0 + rM P ,0

(3.4)

23

Vehicle Dynamics

where the vector

FH Regensburg, University of Applied Sciences

r0M

describes the position of the rim center

P0

does not lie on the track. The corresponding track point

r0P0 ,0 =

M.

Usually, the point

follows from

(3.5)

z (x , y )
where Eq. (3.2) was used to calculate the appropriate road height. In the point
track normal

en

P0

the

is calculated, now. Then the unit vectors in the tire's circumferential

direction and lateral direction can be determined. One gets

ex =
where

ex

eyR en
| eyR en |

ey = en ex ,

and

(3.6)

eyR denotes the unit vector into the direction of the wheel rotation axis. Calculating
eyR not always being perpendicular to the track. The tire

demands a normalization, as

camber angle

= arcsin eTyR en

(3.7)

describes the inclination of the wheel rotation axis against the track normal.
The vector from the rim center

P0

to the track point

is split into three parts now

rM P0 = rS ezR + a ex + b ey ,
where

rS

denotes the loaded or static tire radius,

a, b

(3.8)

are distances measured in circum-

ferential and lateral direction, and the radial direction is given by the unit vector

ezR = ex eyR
which is perpendicular to

ex

and

eyR .

(3.9)

A scalar multiplication of Eq. (3.8) with

en

results

in

eTn rM P0 = rS eTn ezR + a eTn ex + b eTn ey .


As the unit vectors

ex

and

ey

are perpendicular to

en

(3.10)

Eq. (3.10) simplies to

eTn rM P0 = rS eTn ezR .

(3.11)

Hence, the static tire radius is given by

rS
The contact point

eTn rM P0
= T
.
en ezR

(3.12)

given by the vector

rM P = rS ezR

(3.13)

lies within the rim center plane. The transition from the point
takes place according to Eq. (3.8) by the terms

24

a ex

and

b ey

P0

to the contact point

perpendicular to the track

FH Regensburg, University of Applied Sciences

normal

en .

Prof. Dr.-Ing. G. Rill

The track normal, however, was calculated in the point

track the point

P0 .

With an uneven

no longer lies on the track and can therefor no longer considered as

contact point.

P = P now the Eqs. (3.5)


P and P0 is suciently small.

With the newly estimated value


until the dierence between

to (3.13) can be repeated

Tire models which can be simulated within acceptable time assume that the contact patch
is suciently at. At an ordinary passenger car tire, the contact area has at normal load
approximately the size of

1520cm. It makes no sense to calculate a ctitious contact point

to fractions of millimeters, when later on the real track will be approximated by a plane
in the range of centimeters. If the track in the contact area is replaced by a local plane,
no further iterative improvements will be necessary for the contact point calculation.

3.2.2 Local Track Plane


A plane is given by three points. In order to get a good approximation to the local track
unevenness four point will be used to determine the local track normal. Using the initial
guess in Eq. (3.3) and the wheel rotation axis
estimated by

ex =

eyr

the circumferential direction can be

eyR ezB
.
| eyR ezB |

(3.14)

Similar to Eq. (3.4) four points are generated now

xi

= yi ,
zi

i = 1(1)4 .

(3.15)

In order to sample the contact patch as good as possible the tire width

b and the unloaded

r0Qi ,0 = r0M,0 + rM Qi ,0

tire radius

r0

are used to place the points via

x r0 ex

r0 ezB ,

rM Q2 = x r0 ex

r0 ezB ,

rM Q1 =

y b eyR r0 ezB ,

rM Q3 =

(3.16)

y b eyR r0 ezB

rM Q4 =

in the front, in the rear, in the left, and in the right of the contact patch.
According to Eq. (3.5) the corresponding points on the track are given by

r0Qi ,0

xi

=
yi
,

z (xi , yi )

i = 1(1)4 .

(3.17)

The calculation of the track normal is straight forward now, Figure 3.7. The vectors

25

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

eyR
Q1*

rMP*

en

Q1

Q3*
Q3

P*

rQ2Q1

P
Q4*
Q4

Q2*

rQ3Q4
Q2

Figure 3.7: Local track plane

rQ2 Q1 = r0Q1 r0Q2

and

rQ4 Q3 = r0Q3 r0Q4

dene the local track inclination in longitudinal

and lateral direction. Hence, the local track normal is dened by

en =
The unit vectors

ex , ey

rQ2 Q1 rQ4 Q3
.
| rQ2 Q1 rQ4 Q3 |

(3.18)

in longitudinal and lateral direction are calculated from Eq. (3.6).

The mean value of the track points

r0P0 ,0 =
serves as rst improvement of
point

1
(r0Q1 ,0 + r0Q2 ,0 + r0Q3 ,0 + r0Q4 ,0 )
4

the contact point, P P0 . Finally,

(3.19)
the corresponding

in the rim center plane is obtained by Eqs. (3.12) and (3.13). On rough roads

the point

not always is located on the track. But, together with the local track normal

it represents the local road plane very well. As in reality, sharp bends and discontinuities
which will occur at step- or ramp-sized obstacles are smoothed by this approach.

3.2.3 Tire Deflection


For a vanishing camber angle

=0

the deected zone has a rectangular shape, Figure

3.8. Its area is given by

A0 = 4z b ,
where

(3.20)

is the width of the tire, and the tire deection is obtained by

4z = r0 rS .

(3.21)

Here, the width of the tire simply equals the width of the contact zone,

wC = b.

On a cambered tire the deected zone of the tire cross section depends on the contact
situation. The magnitude of the tire ank radii

rSL = rs +

26

b
tan
2

and

rSR = rs

b
tan
2

(3.22)

FH Regensburg, University of Applied Sciences

=/ 0

=0
eyR

eyR

rS

Prof. Dr.-Ing. G. Rill

rSL

r0

en

eyR

rSR

rS

en

rS

r0

en

r0

z
b

wC = b

rSR

b*
wC

wC

full contact

partial contact

Figure 3.8: Tire deection

determines the shape of the deected zone.

rSL r0

The tire will be in full contact to the road if

and

rSR r0

hold. Then, the

deected zone has a trapezoidal shape with an area of

A =

1
(r0 rSR + r0 rSL ) b = (r0 rS ) b .
2

Equalizing the cross sections

A0 = A

(3.23)

results in

4z = r0 rS .
Hence, at full contact the tire camber angle

(3.24)

has no inuence to the vertical tire force.

But, due to

wC =

b
cos

(3.25)

the width of the contact area increases with the tire camber angle.
The deected zone will change to a triangular shape if one of the ank radii exceeds the
undeected tire radius. Assuming

rSL > r0

obtained by

A =

and

rSR < r0

1
(r0 rSR ) b ,
2

the area of the deected zone is

(3.26)

where the width of the deected zone follows from

b =

r0 rSR
.
tan

(3.27)

Now, Eq. (3.26) reads as

1 (r0 rSR )2
A =
.
2 tan

(3.28)

27

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Equalizing the cross sections

A0 = A

results in

1 r0 rS + 2b tan
4z =
2
b tan
where Eq. (3.22) was used to express the ank radius
tire width

and the camber angle

wC =

2

rSR

(3.29)

by the static tire radius

| tan |

r0 rS + 2b tan
r0 rSR
b
=
=
,
cos
tan cos
sin

and

| sin |

the

Now, the width of the contact area is given by

(3.30)

where the Eqs. (3.27) and (3.22) where used to simplify the expression. If
are replaced by

rS ,

tan

and

sin

then, the Eqs. (3.29) and (3.30) will hold for positive

and negative camber angles.

3.2.4 Length of Contact Patch


To approximate the length of the contact patch the tire deformation is split into two
parts, Figure 3.9. By

4zF

and

4zB

the average tire ank and the belt deformation are

measured. Hence, for a tire with full contact to the road

4z = 4zF + 4zB = r0 rS

(3.31)

will hold.

undeformed
belt

Belt
Fz
Rim

zF

rS r0

r0
L/2

zB

zB

Figure 3.9: Length of contact patch

Assuming both deections being equal will lead to

4zF 4zB

1
4z .
2

(3.32)

Approximating the belt deection by truncating a circle with the radius of the undeformed
tire results in

28

 2
L
+ (r0 4zB )2 = r02 .
2

(3.33)

FH Regensburg, University of Applied Sciences

In normal driving situations the belt deections are small,

Prof. Dr.-Ing. G. Rill

4zB  r0 .

Hence, Eq. (3.33)

can be simplied and nally results in

L2
= 2 r0 4zB
4

L =

or

p
8 r0 4zB .

(3.34)

Inspecting the passenger car tire footprint in Figure 3.1 leads to a contact patch length of

L 140mm. For this tire the radial stiness and the inated radius are specied with cR =
265 000 N/m and r0 = 316.9 mm. The overall tire deection can be estimated by 4z =
Fz /cR . At the load of Fz = 4700N the deection amounts to 4z = 4700N / 265 000N/m =
0.0177 m. Then, by approximating the belt deformation
by the half of the tire deection,
p
the length of the contact patch will become L =
8 0.3169 m 0.0177/2 m = 0.1498 m =
150 mm which corresponds quite well with the length of the tire footprint.

3.2.5 Static Contact Point


Assuming that the pressure distribution on a cambered tire with full road contact corresponds with the trapezoidal shape of the deected tire area, the acting point of the
resulting vertical tire force
static contact point

Q,

FZ

will be shifted from the geometric contact point

to the

Figure 3.10.

rS
en
P

ey
r0-rSL

Fz
A

r0-rSR

y
wC

Figure 3.10: Lateral deviation of contact point at full contact


The center of the trapezoidal area determines the lateral deviation

yQ .

By splitting the

area into a rectangular and a triangular section we will obtain

yQ =

y  A  + y 4 A4
.
A

(3.35)

The minus sign takes into account that for positive camber angles the acting point will
move to the right whereas the unit vector

ey

dening the lateral direction points to the

left. The area of the whole cross section results from

A =

1
(r0 rSL + r0 rSR ) wC ,
2

(3.36)

29

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

where the width of the contact area

wC

is given by Eq. (3.25). Using the Eqs. (3.22) and

(3.24) the expression can be simplied to

A = 4z wC .

(3.37)

As the center of the rectangular section is located on the center line which runs through the
geometric contact point,

y = 0

will hold. The distance from the center of the triangular

section to the center line is given by

y4 =

1
1
1
wC wC = wC .
2
3
6

(3.38)

Finally, the area of the triangular section is dened by

A4 =

1
1
1
(r0 rSR (r0 rSL )) wC =
(rSL rSR )) wC =
(b tan ) wC ,
2
2
2

(3.39)

where Eq. (3.22) was used to simplify the expression. Now, Eq. (3.35) can be written as

yQ =

1
6

wC

1
2

b tan wC
b tan
b2 tan
=
wC =
.
4z wC
12 4z
12 4z cos

(3.40)

If the cambered tire has only a partial contact to the road then, according to the deection
area a triangular pressure distribution will be assumed, Figure 3.11.

b/2

en
ey

Q
Fz

y
wC

Figure 3.11: Lateral deviation of contact point at partial contact


Now, the location of the static contact point


yQ =
where the width of the contact area

wC

is given by

1
b
wC
3
2 cos


,

(3.41)

is determined by Eq. (3.30) and the term

describes the distance from the geometric contact point

b/(2 cos )

to the outer corner of the

contact patch. The plus sign holds for positive and the minus sign for negative camber
angles.
The static contact point

described by the vector

r0Q = r0P + yQ ey
represents the contact patch much better than the geometric contact point

30

(3.42)

P.

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

3.2.6 Contact Point Velocity


To calculate the tire forces and torques which are generated by friction the contact point
velocity will be needed. The static contact point

given by Eq. (3.42) can be expressed

as follows

r0Q = r0M + rM Q ,
where

(3.43)

rM Q describes the position of


Q relative to the wheel center M . The absolute velocity of the contact

denotes the wheel center and hence, the vector

static contact point

point will be obtained from

v0Q,0 = r0Q,0 = r0M,0 + rM Q,0 ,


where

r0M,0 = v0M,0

(3.44)

denotes the absolute velocity of the wheel center. The vector

rM Q

takes part on all those motions of the wheel carrier which do not contain elements of the
wheel rotation and it In addition, it contains the tire deection

4z

normal to the road.

Hence, its time derivative can be calculated from

rM Q,0 = 0R,0
rM Q,0 + 4z en,0 ,
where

0R

(3.45)

is the angular velocity of the wheel rim without any component in the direction

of the wheel rotation axis,

4z

denotes the change of the tire deection, and

en

describes

the road normal. Now, Eq. (3.44) reads as

v0Q,0 = v0M,0 + 0R,0


rM Q,0 + 4z en,0 .
As the point

lies on the track,

v0Q,0

(3.46)

must not contain any component normal to the

track

eTn,0 v0P,0 = 0
As

en,0

or

is a unit vector,

eTn,0 v0M,0 + 0R,0


rM Q,0 + 4z eTn,0 en,0 = 0 .

eTn,0 en,0 = 1

(3.47)

will hold, and then, the time derivative of the tire

deformation is simply given by

4z = eTn,0 v0M,0 + 0R,0


rM Q,0 .

(3.48)

Finally, the components of the contact point velocity in longitudinal and lateral direction
are obtained from

rM Q,0
vx = eTx,0 v0Q,0 = eTx,0 v0M,0 + 0R,0

(3.49)

and

rM Q,0 ,
vy = eTy,0 v0P,0 = eTy,0 v0M,0 + 0R,0
where the relationships

(3.50)

eTx,0 en,0 = 0 and eTy,0 en,0 = 0 were used to simplify the expressions.

31

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

deflected tire

rigid wheel

r0 r
S

rD

vt

x
Figure 3.12: Dynamic rolling radius

3.2.7 Dynamic Rolling Radius


At an angular rotation of

4, assuming the tread particles stick to the track, the deected


x, Figure 3.12.

tire moves on a distance of


With

r0

as unloaded and

rS = r0 4r

as loaded or static tire radius

r0 sin 4 = x

(3.51)

r0 cos 4 = rS

(3.52)

and

hold.
If the motion of a tire is compared to the rolling of a rigid wheel, then, its radius
have to be chosen so that at an angular rotation of

rD

will

the tire moves the distance

r0 sin 4 = x = rD 4 .

(3.53)

Hence, the dynamic tire radius is given by

rD =
For

4 0

r0 sin 4
.
4

one obtains the trivial solution

(3.54)

rD = r0 .

At small, yet nite angular rotations the sine-function can be approximated by the rst
terms of its Taylor-Expansion. Then, Eq. (3.54) reads as

rD

4 16 43
= r0
= r0
4

1
1 42
6


.

(3.55)

With the according approximation for the cosine-function

rS
1
= cos 4 = 1 42
r0
2

32

or

4 = 2

rS
1
r0


(3.56)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

one nally gets


r D = r0
Due to

rS = rS (Fz )

1
1
3

the ctive radius

rS
1
r0

rD


=

2
1
r0 + rS .
3
3

Fz . Therefore,
velocity , then

depends on the wheel load

called dynamic tire radius. If the tire rotates with the angular

(3.57)

v t = rD

it is

(3.58)

will denote the average velocity at which the tread particles are transported through the
contact patch.

[mm]

10
0

-10

r -r
D

-20

Fz [kN]

Measurements
TMeasy tire model

Figure 3.13: Dynamic tire radius


In extension to Eq. (3.57), the dynamic tire radius is approximated in the tire model
TMeasy by


FzS
= r0 + (1 ) r0
c
|
{z 0 }
rS


rD

(3.59)

where the static tire radius rS = r0 4r has been approximated by using the linearized
S
tire deformation 4r = Fz /c0 . The parameter is modeled as a function of the wheel
load Fz



= N + ( 2N N )
N and 2N
Fz = 2FzN .

where

Fz
1
FzN

denote the values for the pay load

Fz = FzN

(3.60)

and the doubled pay load

With the TMeasy parameters for a passenger car tire

vertical tire stiffness at fz=fz0


vertical tire stiffness at fz=2*fz0

[N/m],
[N/m],

coefficient for dynamic tire radius fz=fz0 [-],


coefficient for dynamic tire radius fz=2*fz0 [-],

190000.
206000.
0.375
0.750

the approximation of measured tire data can be done very well, Figure 3.13.

33

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

3.3 Forces and Torques caused by Pressure Distribution


3.3.1 Wheel Load
The vertical tire force

Fz

can be calculated as a function of the normal tire deection

and the deection velocity

4z

4z
Fz = Fz (4z, 4z)
.

(3.61)

Because the tire can only apply pressure forces to the road the normal force is restricted
to

Fz 0.

In a rst approximation

Fz

is separated into a static and a dynamic part

Fz = FzS + FzD .

(3.62)

The static part is described as a nonlinear function of the normal tire deection

FzS = a1 4z + a2 (4z)2 .
The constants

a1

and

a2

(3.63)

may be calculated from the radial stiness at nominal and double

payload

cN


d FzS
=
d 4z FzS =FzN

c2N

and


d FzS
=
.
d 4z FzS =2FzN

(3.64)

The derivative of Eq. (3.63) results in

d FzS
= a1 + 2 a2 4z .
d 4z

(3.65)

From Eq. (3.63) one gets

4z =

a1

a21 + 4a2 FzS


.
2a2

(3.66)

Because the tire deection is always positive, the minus sign in front of the square root
has no physical meaning, and can be omitted therefore. Hence, Eq. (3.65) can be written
as

d FzS
= a1 + 2 a2
d 4z

a1 +

a21 + 4a2 FzS


2a2

!
=

a21 + 4a2 FzS .

(3.67)

Combining Eqs. (3.64) and (3.67 results in

cN
c2N

a21 + 4a2 FzN


q
a21 + 4a2 2FzN
=

or

c2N

or

c22N

and

a2 =

= a21 + 4a2 FzN ,


=

a21

(3.68)

8a2 FzN

nally leading to

a1 =

34

2 c2N c22N

c22N c2N
.
4 FzN

(3.69)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

Results for a passenger car and a truck tire are shown in Figure 3.14. The parabolic
approximation in Eq. (3.63) ts very well to the measurements. The radial tire stiness

Fz = 3 200 N can be specied with c0 =


Fz = 35 000 N and the stiness c0 = 1 250 000N/m of a truck

of the passenger car tire at the payload of

190 000N/m.

The Payload

tire are signicantly larger.

Passenger Car Tire: 205/50 R15

80

60

40
20

2
0

Truck Tire: X31580 R22.5

100

Fz [kN]

Fz [kN]

10

10

20
30
z [mm]

40

50

Figure 3.14: Tire radial stiness:

20

40
60
z [mm]

80

Measurements,  Approximation

The dynamic part is roughly approximated by

FzD = dR 4z ,
where

dR

(3.70)

is a constant describing the radial tire damping, and the derivative of the tire

deformation

4z

is given by Eq. (3.48).

3.3.2 Tipping Torque


The lateral shift of the vertical tire force
static contact point

Fz

from the geometric contact point

is equivalent to a force applied in

Mx = Fz y
acting around a longitudinal axis in

P,

to the

and the tipping torque


(3.71)

Figure 3.15.

Note: Figure 3.15 shows a negative tipping torque. Because a positive camber angle moves
the contact point into the negative

y -direction and hence, will generate a negative tipping

torque.
As long as the cambered tire has full contact to the road the lateral displacement

is

given by Eq. (3.40). Then, Eq. (3.71) reads as

Mx = Fz

b2 tan
.
12 4z cos

(3.72)

35

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

en
ey

en

ey

P Q
y

Fz

Fz

Tx

Figure 3.15: Tipping torque at full contact

If the wheel load is approximated by its linearized static part


camber angles

||  1

Mx = cN 4z
where the term

1
12

c N b2

Fz cN 4z

and small

are assumed, then, Eq. (3.72) simplies to

1
b2
=
c N b2 ,
12 4z
12

(3.73)

can be regarded as the tipping stiness of the tire.

en
ey

Q
P

Fz

Figure 3.16: Cambered tire with partial contact


The use of the tipping torque instead of shifting the contact point is limited to those cases
where the tire has full or nearly full contact to the road. If the cambered tire has only
partly contact to the road, the geometric contact point
the contact area whereas the static contact point
3.16. In the following the static contact

may even be located outside

is still a real contact point, Figure

will be used as the contact point, because it

represents the contact area more precisely than the geometric contact point

P.

3.3.3 Rolling Resistance


If a non-rotating tire has contact to a at ground the pressure distribution in the contact
patch will be symmetric from the front to the rear, Figure 3.17. The resulting vertical
force

Fz

is applied in the center

torque around the

of the contact patch and hence, will not generate a

y -axis.

If the tire rotates tread particles will be stued into the front of the contact area which
causes a slight pressure increase, Figure 3.17. Now, the resulting vertical force is applied
in front of the contact point and generates the rolling resistance torque

ty = Fz xR sign() ,

36

(3.74)

FH Regensburg, University of Applied Sciences

rotating

en
ex non-rotating

xR

ex

Prof. Dr.-Ing. G. Rill

en
C

Fz

Fz

Figure 3.17: Pressure distribution at a non-rotation and rotation tire

where

sign() assures
xR from C to
r0

distance
radius

that

ty

always acts against the wheel angular velocity

the working point of

fR =

Fz

The

usually is related to the unloaded tire

xR
.
r0

(3.75)

The dimensionless rolling resistance coecient slightly increases with the traveling velocity

of the vehicle

fR = fR (v) .
20 km/h < v < 200 km/h, the rolling
tires is in the range of 0.01 < fR < 0.02.

Under normal operating conditions,


coecient for typical passenger car

(3.76)
resistance

The rolling resistance hardly inuences the handling properties of a vehicle. But it plays
a major part in fuel consumption.

3.4 Friction Forces and Torques


3.4.1 Longitudinal Force and Longitudinal Slip
To get a certain insight into the mechanism generating tire forces in longitudinal direction,
we consider a tire on a at bed test rig. The rim rotates with the angular velocity
the at bed runs with the velocity

vx .

and

The distance between the rim center and the at

bed is controlled to the loaded tire radius corresponding to the wheel load

Fz , Figure 3.18.

t = 0 the contact patch. If we assume adhesion between


the particle and the track, then the top of the particle will run with the bed velocity vx
and the bottom with the average transport velocity vt = rD . Depending on the velocity
dierence 4v = rD vx the tread particle is deected in longitudinal direction

A tread particle enters at the time

u = (rD vx ) t .

(3.77)

The time a particle spends in the contact patch can be calculated by

T =

L
,
rD ||

(3.78)

37

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

rD
vx

rD

vx
L

u max

Figure 3.18: Tire on at bed test rig

where

denotes the contact length, and

T >0

is assured by

||.

The maximum deection occurs when the tread particle leaves the contact patch at the
time

t=T
umax = (rD vx ) T = (rD vx )

L
.
rD ||

(3.79)

The deected tread particle applies a force to the tire. In a rst approximation we get

Fxt = ctx u ,
where

ctx

(3.80)

represents the stiness of one tread particle in longitudinal direction.

On normal wheel loads more than one tread particle is in contact with the track, Figure
3.19a. The number

of the tread particles can be estimated by

p =
where

is the length of one particle and

a)

Figure 3.19:

(3.81)

denotes the distance between the particles.

b)

L
,
s+a

cxt * u

cut * u max

a) Particles, b) Force distribution,

Particles entering the contact patch are undeformed, whereas the ones leaving have the
maximum deection. According to Eq. (3.80), this results in a linear force distribution

38

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

versus the contact length, Figure 3.19b. The resulting force in longitudinal direction for

particles is given by

Fx =

1 t
p c umax .
2 x

(3.82)

Using the Eqs. (3.81) and (3.79) this results in

Fx =

1 L t
L
cx (rD vx )
.
2 s+a
rD ||

(3.83)

A rst approximation of the contact length L was calculated in Eq. (3.34). Approximating
1
F /c results in
the belt deformation by 4zB
2 z R

Fz
,
cR

L2 4 r 0
where

cR

(3.84)

denotes the radial tire stiness, and nonlinearities and dynamic parts in the tire

deformation were neglected. Now, Eq. (3.82) can be written as

Fx = 2

rD v x
r0 ctx
Fz
.
s + a cR
rD ||

(3.85)

The nondimensional relation between the sliding velocity of the tread particles in lonS
gitudinal direction vx = vx rD and the average transport velocity rD || form the
longitudinal slip

sx =
The longitudinal force

sx

Fx

(vx rD )
.
rD ||

(3.86)

is proportional to the wheel load

Fz

and the longitudinal slip

in this rst approximation

Fx = k Fz sx ,
where the constant

summarizes the tire properties

(3.87)

r0 , s, a, ctx

and

cR .

Eq. (3.87) holds only as long as all particles stick to the track. At moderate slip values
the particles at the end of the contact patch start sliding, and at high slip values only the
parts at the beginning of the contact patch still stick to the road, Figure 3.20.

small slip values


Fx = k * Fz* s x
L

moderate slip values


Fx = Fz * f ( s x )

high slip values


Fx = FG

L
t

Fx <= FH
adhesion

L
t

Fx = FG

Fx = FH

sliding

adhesion sliding

Figure 3.20: Longitudinal force distribution for dierent slip values


The resulting nonlinear function of the longitudinal force

Fx

versus the longitudinal slip


sx can be dened by the parameters initial inclination (driving stiness) dFx0 , location
M
S
S
sM
x and magnitude of the maximum Fx , start of full sliding sx and the sliding force Fx ,
Figure 3.21.

39

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Fx
M

adhesion

sliding

Fx
S
Fx dFx0

sSx

sM
x

sx

Figure 3.21: Typical longitudinal force characteristics

3.4.2 Lateral Slip, Lateral Force and Self Aligning Torque


Similar to the longitudinal slip

sx ,

given by Eq. (3.86), the lateral slip can be dened by

vyS
sy =
,
rD ||

(3.88)

where the sliding velocity in lateral direction is given by

vyS = vy

(3.89)

and the lateral component of the contact point velocity

vy

follows from Eq. (3.50).

As long as the tread particles stick to the road (small amounts of slip), an almost linear
distribution of the forces along the length

of the contact patch appears. At moderate

slip values the particles at the end of the contact patch start sliding, and at high slip
values only the parts at the beginning of the contact patch stick to the road, Figure 3.22.

Fy

large slip values


Fy = FG

sliding

Fy

adhesion

moderate slip values


Fy = Fz * f ( s y )

sliding

L
Fy n

adhesion

small slip values


Fy = k * Fz * s y

Figure 3.22: Lateral force distribution over contact patch

The nonlinear characteristics of the lateral force versus the lateral slip can be described
0
M
M
by the initial inclination (cornering stiness) dFy , the location sy and the magnitude Fy

40

FH Regensburg, University of Applied Sciences

of the maximum, the beginning of full sliding

sSy ,

Prof. Dr.-Ing. G. Rill

and the magnitude

FyS

of the sliding

force.
The distribution of the lateral forces over the contact patch length also denes the point
of application of the resulting lateral force. At small slip values this point lies behind
the center of the contact patch (contact point P). With increasing slip values it moves
forward, sometimes even before the center of the contact patch. At extreme slip values,
when practically all particles are sliding, the resulting force is applied at the center of the
contact patch.
The resulting lateral force

Fy

with the dynamic tire oset or pneumatic trail

as a lever

generates the self aligning torque

TS = n Fy .
The lateral force

sy .

Fy

(3.90)

as well as the dynamic tire oset are

functions of the lateral slip

Typical plots of these quantities are shown in Figure 3.23. Characteristic parameters

n/L
(n/L)0

Fy
M

adhesion

Fy

adhesion
adhesion/sliding
full sliding

adhesion/
sliding

s0y

full sliding

Fy dF0y

MS

sSy

sy

adhesion
adhesion/sliding
full sliding

sM
y

sSy

sy

s0y

sSy

sy

Figure 3.23: Typical plot of lateral force, tire oset and self aligning torque

M
0
of the lateral force graph are initial inclination (cornering stiness) dFy , location sy and
S
S
M
magnitude of the maximum Fy , begin of full sliding sy , and the sliding force Fy .
The dynamic tire oset has been normalized by the length of the contact patch L. The
0
S
initial value (n/L)0 as well as the slip values sy and sy suciently characterize the graph.

3.4.3 Wheel Load Influence


The resistance of a real tire against deformations has the eect that with increasing
wheel load the distribution of pressure over the contact area becomes more and more
uneven. The tread particles are deected just as they are transported through the contact
area. The pressure peak in the front of the contact area cannot be used, for these tread
particles are far away from the adhesion limit because of their small deection. In the rear
of the contact area the pressure drop leads to a reduction of the maximally transmittable

41

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Longitudinal force

Fz = 3.0 kN

Fx

Lateral force

Fz = 6.0 kN

Fz = 3.0 kN

Fy

Fz = 6.0 kN

dFx0 = 70 kN

dFx0 = 220 kN

dFy0 = 72 kN

dFy0 = 130 kN

sM
x = 0.160

sM
x = 0.120

sM
y = 0.180

sM
y = 0.200

FxM = 2.90 kN FxM = 5.60 kN FyM = 2.85 kN FyM = 5.40 kN


sSx = 0.500

sSx = 0.500

FxS = 2.65 kN FxS = 5.10 kN

sSy = 0.500

sSy = 0.700

FyS = 2.80 kN FyS = 5.30 kN

Table 3.3: Characteristic tire data with degressive wheel load inuence

friction force. With rising imperfection of the pressure distribution over the contact area,
the ability to transmit forces of friction between tire and road lessens.
In practice, this leads to a degressive inuence of the wheel load on the characteristic
curves of longitudinal and lateral forces. In order to respect this fact in a tire model, the
N
N
characteristic data for two nominal wheel loads Fz and 2 Fz are given in Table 3.3.
M
M
0
0
From this data the initial inclinations dFx , dFy , the maximal forces Fx , Fx and the
M
S
sliding forces Fx , Fy for arbitrary wheel loads Fz are calculated by quadratic functions.
For the maximum longitudinal force it reads as





Fz
1 M
1 M
N
M
N Fz
N
N
M
= N 2 Fx (Fz ) 2 Fx (2Fz ) Fx (Fz ) 2 Fx (2Fz ) N .
Fz
Fz

Fy [kN]

Fx [kN]

FxM (Fz )

-2

-2

-4

-4

-6
-0.4

-0.2

0.2

0.4

-6

-20

-10

Figure 3.24: Longitudinal and lateral force characteristics:

M
sM
x , sy ,

10

20

Fz = 1.8, 3.2, 4.6, 5.4, 6.0 kN

sSx , sSy , at which full sliding


load Fz . For the location of the

and the slip values,

appears, are dened as linear functions of the wheel

42

[deg]

sx [-]

The location of the maxima

(3.91)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

maximum longitudinal force this results in

sM
x (Fz )

N
sM
x (Fz )

N
sM
x (2Fz )


 F
z
1 .
FzN

N
sM
x (Fz )

(3.92)

With the numeric values from Tab. 3.3 a slight shift of the maxima with an increasing
wheel load is also modeled, Figure 3.24.

3.4.4 Two-Dimensional Tire Characteristics


The longitudinal force as a function of the longitudinal slip

Fx = Fx (sx )

and the lateral

force depending on the lateral slip Fy = Fy (sy ) can be dened by their characteristic
0
0
M
M
parameters initial inclination dFx , dFy , location sx , sy and magnitude of the maximum
M
M
S
S
Fx , Fy as well as sliding limit sx , sy and sliding force FxS , FyS , Figure 3.25. During
general driving situations, e.g. acceleration or deceleration in curves, longitudinal sx and

sy

lateral slip

Fx

appear simultaneously.

Fx

Fx

Fy

0
dF x

sx
sSx

sM
x

Fy Fy

dF 0
FS

FM

Fy

F(s)
Fx

sy

dF y
sSy
sS
sy

sM

sM
y

sx

Figure 3.25: Generalized tire characteristics


The longitudinal slip

sx

and the lateral slip

sy

can vectorally be added to a generalized

slip

s
s =
where the slips

sx

and

sy

sx
sx

2


+

sy
sy

2
,

were normalized by appropriate weighting factors

(3.93)

sx

and

sy .

43

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Similar to the graphs of the longitudinal and lateral forces the graph of the generalized tire
0
M
M
S
S
force is dened by the characteristic parameters dF , s , F , s and F . The parameters
are calculated from the corresponding values of the longitudinal and lateral force

dF

sM =
FM =
sS =
FS =

q
2
(dFx0 sx cos )2 + dFy0 sy sin ,
s
2  M
2
sy
sM
x
cos +
sin ,
sx
sy
q
2
(FxM cos )2 + FyM sin ,
s
2  S
2
sy
sSx
cos +
sin ,
sx
sy
q
2
(FxS cos )2 + FyS sin ,

(3.94)

where the slip normalization have also to be considered at the initial inclination. The
angular functions

cos =

sx /
sx
s

and

sin =

sy /
sy
s

(3.95)

grant a smooth transition from the characteristic curve of longitudinal to the curve of

lateral forces in the range of = 0 to = 90 .

F = F (s) is now described in intervals by a broken rational function, a cubic


S
polynomial, and by the sliding force F


,
sM dF 0
= M ,
0 s sM ;

s
s

1 + + dF 0 M 2

F
F (s) =
(3.96)
s sM
M
M
S
2
M
S

(F

F
)

(3

2
)
,

=
,
s
<
s

s
;

sS sM

FS ,
s > sS .
The function

0
When dening the curve parameters, one just has to make sure that the condition dF
M
2 FsM is fullled, because otherwise the function has a turning point in the interval 0
s sM .

<

Now, the longitudinal and the lateral force follow from the according projections in longitudinal and lateral direction

Fx = F cos

and

Fy = F sin .

(3.97)

Hence, within TMeasy the one-dimensional characteristics are automatically converted to


a two-dimensional combination characteristics, Figure 3.26.

44

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

30
3
20
10
F [kN]

1
0

F [kN]

-1

-10

-2

-20

-3
-4

-2

0
F [kN]

-30

-20

0
F [kN]

20

|| = 1, 2, 4, 6, 10, 14

|sx | = 1, 2, 4, 6, 10, 15 %;

Figure 3.26: Two-dimensional tire characteristics at

Fz = 3.2 kN

Fz = 35 kN

3.4.5 Different Friction Coefficients


The tire characteristics are valid for one specic tire road combination only. Hence, different tire road combinations will demand for dierent model parameter.

Fy [N]

4000

L /0

3000

0.2

2000
1000

0.4

0.6

-1000

0.8

-2000

1.0

-3000
-4000
-0.5

sy [-]

0.5

Fz = 3.2 kN

Figure 3.27: Lateral force characteristics for dierent friction coecients


If only the coecient of friction is changed a simple but eective adaption of given model
data is possible. A reduced or changed friction coecient mainly inuences the maximum
force and the sliding force, whereas the initial inclination will remain unchanged. So, by
setting

sM

L M
L M
s , FM
F ,
0
0

sS

L S
L S
s , FS
F ,
0
0

(3.98)

45

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

the essential tire model parameter which are valid for the friction coecient
justed to the new friction coecient

L .

are ad-

The result of this simple approach is shown in

Figure 3.27.
If the road model provides not only the roughness information
local friction coecient

[z, L ] = fR (x, y)

then, braking on

z = fR (x, y)

-split

but also the

maneuvers can easily

be simulated.

3.4.6 Self Aligning Torque


According to Eq. (3.90) the self aligning torque can be calculated via the dynamic tire
oset. The dynamic tire oset

can be normalized by the length

of the contact area,

nN = n/L. It mainly depends on the lateral slip sy . The normalized tire oset starts
at sy = 0 with an initial value (n/L)0 . It tends to zero, n/L 0 at large slip values,
sy sSy . Sometimes the dynamic tire oset overshoots to negative values before it reaches
S
0
zero again. This behavior can be modeled by introducing the parameter sy < sy , Figure
3.28.

n/L

n/L

(n/L)0

(n/L)0

s0y

sy

sSy

s0y

sy

Figure 3.28: Normalized tire oset with and without overshoot

In order to achieve a simple and smooth approximation of the normalized tire oset versus
0
the lateral slip, a linear and a cubic function are overlayed in the rst section sy sy



|sy |

(1w) (1s) + w 1 (32s) s2


|sy | s0y and s = 0

sy

2
0  S
n n
|sy | sy sy |sy |
=
(1w)
s0y < |sy | sSy
L
L 0
0
S s0

s
s

y
y
y

0
|sy | > sSy
where the factor

s0y
w= S
sy

(3.99)

(3.100)

0
weights the linear and the cubic function according to the values of the parameter sy
S
0
S
and sy . No overshoot occurs for sy = sy . Here, w = 1 and (1 w) = 0 will produce a
cubic transition from
sy = s0y .

46

n/L = (n/L)0

to

n/L = 0 with vanishing inclinations at sy = 0 and

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

The characteristic curve parameters, which are used for the description of the dynamic
tire oset, are at rst approximation not wheel load dependent. Similar to the description
of the characteristic curves of longitudinal and lateral force, here also the parameters for
single and double pay load are given. The calculation of the parameters of arbitrary wheel
loads is done similar to Eq. (3.92) by linear inter- or extrapolation.

Tz [Nm]

200

Tire oset parameter

150
100

Fz

50

Fz = 3.0 kN

Fz = 6.0 kN

(n/L)0 = 0.170 (n/L)0 = 0.190

0
-50

s0y = 0.200

s0y = 0.220

-100

sE
y = 0.420

sE
y = 0.400

-150
-200
-30

Figure 3.29: Self aligning torque:


The value of

(n/L)0

-20

-10

10

20

[deg]

30

Fz = 1.5, 3.0, 4.5, 6.0, 7.5 kN

can be estimated very well. At small values of lateral slip

sy 0

one

gets at rst approximation a triangular distribution of lateral forces over the contact area
length cf. Figure 3.22. The working point of the resulting force (dynamic tire oset) is
then given by

n(Fz 0, sy = 0) =
The value

1
L.
6

(3.101)

n = 16 L can only serve as reference point, for the uneven distribution of pressure

in longitudinal direction of the contact area results in a change of the deexion prole
and the dynamic tire oset.
The self aligning torque in Figure 3.29 has been calculated with the tire parameter from
Table 3.3. The degressive inuence of the wheel load on the self aligning torque can be
seen here as well.
With the parameters for the description of the tire oset it has been assumed that at
N
the payload Fz = Fz the related tire oset reaches the value of (n/L)0 = 0.17 1/6 at
sy = 0. The slip value s0y , at which the tire oset passes the x-axis, has been estimated.
Usually the value is somewhat higher than the position of the lateral force maximum.
S
With rising wheel load it moves to higher values. The values for sy are estimated too.

47

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

3.4.7 Camber Influence


At a cambered tire, Figure 3.30, the angular velocity of the wheel

has a component

normal to the road

n = sin .

en

rim
centre
plane

eyR

(3.102)

Fy = Fy (s y ): Parameter

4000

3000
2000

1000
0
-1000

ex
ey

rD ||

y()

v()

-2000
-3000

-4000
-0.5

Figure 3.30: Cambered tire

Fy ()

at

0.5

Fz = 3.2 kN

and

= 0 , 2 , 4 , 6 , 8

Now, the tread particles in the contact patch possess a lateral velocity which depends on
their position

and is provided by

v () = n

L
, = sin ,
2 L/2

L/2 L/2 .

(3.103)

At the contact point it vanishes whereas at the end of the contact patch it takes on the
same value as at the beginning, however, pointing into the opposite direction. Assuming
that the tread particles stick to the track, the deection prole is dened by

y () = v () .

(3.104)

The time derivative can be transformed to a space derivative

y () =
where

rD ||

(3.105)

denotes the average transport velocity. Now, Eq. (3.104) can be written as

d y ()
rD || = sin
d

48

d y () d
d y ()
=
rD ||
d dt
d

or

d y ()
sin L
=
,
d
rD || 2 L/2

(3.106)

FH Regensburg, University of Applied Sciences

where

L/2

Prof. Dr.-Ing. G. Rill

was used to achieve dimensionless terms. Similar to the lateral slip

sy

which

is dened by Eq. (3.88) we can introduce a camber slip now

sin L
.
rD || 2

(3.107)

d y ()

= s
.
d
L/2

(3.108)

s =
Then, Eq. (3.106) simplies to

The shape of the lateral displacement prole is obtained by integration


2

1 L
+ C.
(3.109)
y = s
2 2 L/2

1
The boundary condition y = L = 0 can be used to determine the integration constant
2
C . One gets
1 L
C = s
.
(3.110)
2 2
Then, Eq. (3.109) reads as

1 L
y () = s
2 2

"


1

L/2

2 #
.

(3.111)

The lateral displacements of the tread particles caused by a camber slip are compared
now with the ones caused by pure lateral slip, Figure 3.31. At a tire with pure lateral

a) camber slip
y

b) lateral slip
yy()

y()

_
yy

_
y
-L/2

L/2

-L/2

L/2

Figure 3.31: Displacement proles of tread particles


slip each tread particle in the contact patch possesses the same lateral velocity which

dyy /d rD || = vy ,

y y was
dyy /d . Hence, the deection prole is linear, and
reads as yy = vy /(rD ||) = sy , where the denition in Eq. (3.88) was used to
introduce the lateral slip sy . Then, the average deection of the tread particles under

results in

where according to Eq. (3.105) the time derivative

transformed to the space derivative

pure lateral slip is given by

yy = sy

L
.
2

(3.112)

49

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The average deection of the tread particles under pure camber slip is obtained from

ZL/2 "

1 L 1
y = s
2 2 L


1

x
L/2

2 #

1
L
d = s .
3
2

(3.113)

L/2
A comparison of Eq. (3.112) with Eq. (3.113) shows, that by using

1
s
3

sy =
the lateral camber slip

(3.114)

can be converted to an equivalent lateral slip

sy .

In normal driving conditions, the camber angle and thus, the lateral camber slip are
limited to small values. So, the lateral camber force can be approximated by

Fy dFy0 sy .
If the  global inclination

dFy = Fy /sy

(3.115)

is used instead of the initial inclination

dFy0 ,

one

gets the camber inuence on the lateral force as shown in Figure 3.30.
The camber angle inuences the distribution of pressure in the lateral direction of the
contact patch, and changes the shape of the contact patch from rectangular to trapezoidal. Thus, it is extremely dicult, if not impossible, to quantify the camber inuence
with the aid of such simple models. But, it turns out that this approach is quit a good
approximation.

3.4.8 Bore Torque


In particular during steering motions the angular velocity of the wheel

0W = 0R
+ eyR
has a component in direction of the track normal

(3.116)

en

n = eTn 0W 6= 0 .

(3.117)

Then, a very complicated deection prole of the tread particles in the contact patch
occurs. However, by a simple approach the resulting bore torque can be approximated
quite good by the parameter of the generalized tire force characteristics.
At rst, the complex shape of a tire's contact patch is approximated by a circle, Figure
3.32. By setting

1
R =
2

L B
+
2
2

1
(L + B)
4
length L and

the radius of the circle can be adjusted to the

(3.118)
the width

of the actual

contact patch. The integration over the whole circle area results in the bore torque

TB

50

1
=
A

Z
F r dA ,
A

(3.119)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

ex
B

dr
F
r
L

ey

circular
approximation

normal shape of contact patch

Figure 3.32: Bore torque approximation

denotes the force transmitted by the patch element dA, and


2
circle. With dA = r d dr and A = R Eq. (3.119) reads as
where

TB

1
= 2
R

is the area of the

ZR Z2
F r rd dr
0

(3.120)

which immediately results in

TB

R
2
Z
ZR
1
2
F r r dr = 2
= 2
F r2 dr .
R
R
0

(3.121)

For small slip values the force transmitted in the patch element can be approximated by

F = F (s) dF 0 s
where

denotes the slip of the patch element, and

dF 0

(3.122)
is the initial inclination of the

generalized tire force characteristics. Similar to Eqs. (3.86 and (3.88) we dene

s =
where

r n

r n
rD ||

describes the sliding velocity in the patch element, and

(3.123)

rD

and

denote the

dynamic tire radius and the angular velocity of the wheel.


Now, Eq. (3.121) reads as

TB

2
= 2
R

ZR

dF 0

r n 2
r dr
rD ||

(3.124)

51

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

which nally results in

TB

2
n
= 2 dF 0
R
rD ||

ZR

2
1
R n
n R4
0
r dr = 2 dF
=
R dF 0
R
rD || 4
2
rD ||
3

(3.125)

0
where

sB =

R n
rD ||

(3.126)

can be considered as bore slip. Via the initial inclination

dF 0

the bore torque

TB

takes

the actual tire properties into account.


The bore torque calculated by Eq. (3.125) is only a rst approximation. At large bore slips
S
the generalized tire force F is limited to the sliding force F . Then, Eq. (3.121) changes
to

TBmax

2
= 2
R

ZR

F S r2 dr =

2 S R3
2 S
=
F R.
F
R2
3
3

(3.127)

0
Due to the generalized sliding force

FS

the maximum bore torque

TBmax

depends on the

tire properties and the actual friction value. Now, the bore torque is given by

TB =

1
R n
R dF 0
2
rD ||

where according to Eq. (3.118) the circle radius


the width

|TB |

with

2 S
F R
3

(3.128)

can be replaced by the length

and

of the contact patch.

3.4.9 Typical Tire Characteristics


passenger car tire

truck tire

40
20
F [kN]

2
0

F [kN]

1.8 kN
3.2 kN
4.6 kN
5.4 kN

-2
-4
-6
-40

10 kN
20 kN
30 kN
40 kN
50 kN

-20
-40

-20

20

40

-40

-20

sx [%]
Figure 3.33: Longitudinal force:

Meas.,

sx [%]

20

40

TMeasy

The tire model TMeasy which is based on this approach, can be used for passenger car
tires as well as for truck tires. It approximates the characteristic curves

52

Fx = Fx (sx ),

FH Regensburg, University of Applied Sciences

passenger car

Prof. Dr.-Ing. G. Rill

truck

40

Fy [kN]

F [kN]

20

1.8 kN
3.2 kN
4.6 kN
6.0 kN

-2
-4

-40

100

1000

50

500

[Nm]

1500

[Nm]

150

-50

1.8 kN
3.2 kN
4.6 kN
6.0 kN

-100
-150

-20

-10

[o]

10

20

10 kN
20 kN
30 kN
40 kN

-20

-6

Mz

-500

18.4 kN
36.8 kN
55.2 kN

-1000
-1500

-20

Figure 3.34: Lateral force and self aligning torque:

Fy = Fy ()

and

Mz = Mz ()

-10

[o]

Meas.,

quite well  even for dierent wheel loads

10

20

TMeasy

Fz ,

Figures 3.33

and 3.34.
When experimental tire values are missing, the model parameters can be pragmatically
estimated by adjustment of the data of similar tire types. Furthermore, due to their physical signicance, the parameters can subsequently be improved by means of comparisons
between the simulation and vehicle testing results as far as they are available.

53

4 Suspension System
4.1 Purpose and Components
The automotive industry uses dierent kinds of wheel/axle suspension systems. Important
criteria are costs, space requirements, kinematic properties, and compliance attributes.
The main purposes of a vehicle suspension system are

carry the car and its weight,

maintain correct wheel alignment,

control the vehicles direction of travel,

keep the tires in contact with the road,

reduce the eect of shock forces.

Vehicle suspension systems consist of

guiding elements:
control arms, links,
struts,
leaf springs,

force elements:
coil spring, torsion bar, air spring, leaf spring,
anti-roll bar,
damper,
bushings, hydro-mounts,

tires.

Tires are air springs that support the total weight of the vehicle. The air spring action of
the tire is very important to the ride quality and safe handling of the vehicle.

54

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

4.2 Some Examples


4.2.1 Multi Purpose Systems
The double wishbone suspension, the McPherson suspension and the multi-link suspension
are multi purpose wheel suspension systems, Fig. 4.1.

Figure 4.1: Double wishbone, McPherson and multi-link suspension


They are used as steered front or non steered rear axle suspension systems. These suspension systems are also suitable for driven axles.
In a McPherson suspension the spring is mounted with an inclination to the strut axis.
Thus, bending torques at the strut, which cause high friction forces, can be reduced.

leaf springs

links

Figure 4.2: Solid axles guided by leaf springs and links


At pickups, trucks, and busses solid axles are used often. They are guided either by leaf
springs or by rigid links, Fig. 4.2. Solid axles tend to tramp on rough roads.
Leaf-spring-guided solid axle suspension systems are very robust. Dry friction between
the leafs leads to locking eects in the suspension. Although the leaf springs provide
axle guidance on some solid axle suspension systems, additional links in longitudinal and
lateral direction are used. Thus, the typical wind-up eect on braking can be avoided.
Solid axles suspended by air springs need at least four links for guidance. In addition to
a good driving comfort air springs allow level control too.

55

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

4.2.2 Specific Systems


The semi-trailing arm, the short-long-arm axle (SLA), and the twist beam axle suspension
are suitable only for non-steered axles, Fig. 4.3.

Figure 4.3: Specic wheel/axles suspension systems

The semi-trailing arm is a simple and cheap design which requires only few space. It is
mostly used for driven rear axles.
The short-long-arm axle design allows a nearly independent layout of longitudinal and
lateral axle motions. It is similar to the central control arm axle suspension, where the
trailing arm is completely rigid and hence, only two lateral links are needed.
The twist beam axle suspension exhibits either a trailing arm or a semi-trailing arm
characteristic. It is used for non driven rear axles only. The twist beam axle provides
enough space for spare tire and fuel tank.

4.3 Steering Systems


4.3.1 Requirements
The steering system must guarantee easy and safe steering of the vehicle. The entirety
of the mechanical transmission devices must be able to cope with all loads and stresses
occurring in operation.

56

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

In order to achieve a good maneuverability a maximum steering angle of approx.

30

must

be provided at the front wheels of passenger cars. Depending on the wheel base, busses

and trucks need maximum steering angles up to 55 at the front wheels.


Recently some companies have started investigations on `steer by wire' techniques.

4.3.2 Rack and Pinion Steering


Rack-and-pinion is the most common steering system of passenger cars, Fig. 4.4. The rack
may be located either in front of or behind the axle. Firstly, the rotations of the steering

uR S
nk
drag li

wheel
and
wheel
body

rack

pinion
steering
box

Figure 4.4: Rack and pinion steering


wheel

uR = uR (S ) and then via


1 = 1 (uR ), 2 = 2 (uR ). Hence, the

are transformed by the steering box to the rack travel

the drag links transmitted to the wheel rotations

overall steering ratio depends on the ratio of the steering box and on the kinematics of
the steering linkage.

4.3.3 Lever Arm Steering System


steering box
ste
e

ring

ever

l
ring
stee

lev

er 1

drag link 1

drag link 2

2
wheel and
wheel body
Figure 4.5: Lever arm steering system

Using a lever arm steering system Fig. 4.5, large steering angles at the wheels are possible.
This steering system is used on trucks with large wheel bases and independent wheel

57

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

suspension at the front axle. Here, the steering box can be placed outside of the axle
center.
Firstly, the rotations of the steering wheel
rotation of the steer levers

1 = 1 (L ), 2 = 2 (L ).

are transformed by the steering box to the

L = L (S ). The drag links transmit this rotation to the wheel

Hence, the overall steering ratio again depends on the ratio of

the steering box and on the kinematics of the steering linkage.

4.3.4 Drag Link Steering System


At solid axles the drag link steering system is used, Fig. 4.6. The rotations of the steering

L
wheel
and
wheel
body

g
steerin
r
leve

steer box
O

(90o rotated)

steering link

2
drag link
Figure 4.6: Drag link steering system

S are transformed by the steering box to the rotation of the steering lever arm
L = L (S ) and further on to the rotation of the left wheel, 1 = 1 (L ). The drag link
transmits the rotation of the left wheel to the right wheel, 2 = 2 (1 ). The steering ratio

wheel

is dened by the ratio of the steering box and the kinematics of the steering link. Here,
the ratio

2 = 2 (1 )

given by the kinematics of the drag link can be changed separately.

4.3.5 Bus Steer System


In busses the driver sits more than

2 m in front of the front axle. In addition, large steering

angles at the front wheels are needed to achieve a good manoeuvrability. That is why,
more sophisticated steering systems are needed, Fig. 4.7. The rotations of the steering

S are transformed by the steering box to the rotation of the steering lever arm
L = L (S ). The left lever arm is moved via the steering link A = A (L ). This motion
wheel

is transferred by a coupling link to the right lever arm. Finally, the left and right wheels
are rotated via the drag links,

58

1 = 1 (A )

and

2 = 2 (A ).

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

steer
in

g leve
r

L
steering box
steering link

left
lever
arm

drag link

coupl.
link

2
wheel and
wheel body

Figure 4.7: Typical bus steering system

4.4 Standard Force Elements


4.4.1 Springs
Springs support the weight of the vehicle. In vehicle suspensions coil springs, air springs,
torsion bars, and leaf springs are used, Fig. 4.8.

Coil spring

FS

u
u

FS

Air spring

Torsion bar

Leaf spring

FS

FS
Figure 4.8: Vehicle suspension springs

Coil springs, torsion bars, and leaf springs absorb additional load by compressing. Thus,
the ride height depends on the loading condition. Air springs are rubber cylinders lled
with compressed air. They are becoming more popular on passenger cars, light trucks, and
heavy trucks because here the correct vehicle ride height can be maintained regardless of
the loading condition by adjusting the air pressure.

59

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

FS

FS
u

LF

FS0

L
c

L0

FS
Figure 4.9: Linear coil spring and general spring characteristics

A linear coil spring may be characterized by its free length

LF

and the spring stiness

c,

Fig. 4.9. The force acting on the spring is then given by


FS = c LF L ,
where

(4.1)

L denotes the actual length of the spring. Mounted in a vehicle suspension the spring

has to support the corresponding chassis weight. Hence, the spring will be compressed to
the conguration length

L0 < L F .

Now, Eq. (4.1) can be written as

FS = c LF (L0 u)
where

FS0

is the spring preload and


= c LF L0 + c u = FS0 + c u ,

(4.2)

describes the spring displacement measured from

the spring's conguration length.


In general the spring force
placement

FS

can be dened by a nonlinear function of the spring dis-

u
FS = FS (u) .

(4.3)

Now, arbitrary spring characteristics can be approximated by elementary functions, like


polynomials, or by tables which are then inter- and extrapolated by linear functions or
cubic splines.
The complex behavior of leaf springs and air springs can only be approximated by simple
nonlinear spring characteristics,

FS = FS (u).

For detailed investigations sophisticated or

even dynamic spring models have to be used.

4.4.2 Damper
Dampers are basically oil pumps, Fig. 4.10. As the suspension travels up and down, the
hydraulic uid is forced by a piston through tiny holes, called orices. This slows down
the suspension movement.

60

FH Regensburg, University of Applied Sciences

Remote Gas Chamber

FD

Remote Oil Ch.


Remote orifice

Rebound Ch.

Piston

Prof. Dr.-Ing. G. Rill

FD

Compression
Chamber

Piston orifice

v
Figure 4.10: Principle of a mono-tube damper

Today twin-tube and mono-tube dampers are used in vehicle suspension systems. Dynamic
damper models compute the damper force via the uid pressure applied to each side of
the piston. The change in uid pressures in the compression and rebound chambers are
calculated by applying the conservation of mass.
In standard vehicle dynamics applications simple characteristics

FD = FD (v)
are used to describe the damper force

FD

(4.4)

as a function of the damper velocity

v.

To

obtain this characteristics the damper is excited with a sinusoidal displacement signal

u = u0 sin 2f t. By varying the frequency in several steps from f = f0 to f = fE dierent


force displacement curves FD = FD (u) are obtained, Fig. 4.11. By taking the peak values
of the damper force at the displacement u = u0 which corresponds with the velocity
v = 2f u0 the characteristics FD = FD (v) is generated now. Here, the rebound cycle is
associated with negative damper velocities.

FD = FD(u)

FD = FD(v)

FD [N]

1000
0

Compression

f0

-1000
-2000

Rebound

-3000

fE

-4000
-0.06 -0.04 -0.02

0.02 0.04 0.06 -1.6

u [m]

-1.2

-0.8

-0.4

0.4

0.8

1.2

1.6

v [m/s]

Figure 4.11: Damper characteristics generated from measurements


Typical passenger car or truck dampers will have more resistance during its rebound cycle
then its compression cycle.

61

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

4.4.3 Rubber Elements


Force elements made of natural rubber or urethane compounds are used in many locations
on the vehicle suspension system, Fig. 4.12. Those elements require no lubrication, isolate
minor vibration, reduce transmitted road shock, operate noise free, oer high load carrying
capabilities, and are very durable.

Topmount

Stop

Control arm
bushings

Subframe mounts

Figure 4.12: Rubber elements in vehicle suspension


During suspension travel, the control arm bushings provide a pivot point for the control
arm. They also maintain the exact wheel alignment by xing the lateral and vertical
location of the control arm pivot points. During suspension travel the rubber portion of
the bushing must twist to allow control arm movement. Thus, an additional resistance to
suspension movement is generated.
Bump and rebound stops limit the suspension travel. The compliance of the topmount
avoids the transfer of large shock forces to the chassis. The subframe mounts isolate the
suspension system from the chassis and allow elasto-kinematic steering eects of the whole
axle.
It turns out, that those elastic elements can hardly be described by simple spring and
damper characteristics,

FS = FS (u) and FD = FD (v), because their stiness and damping

properties change with the frequency of the motion. Here, more sophisticated dynamic
models are needed.

62

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

4.5 Dynamic Force Elements


4.5.1 Testing and Evaluating Procedures
The eect of dynamic force elements is usually evaluated in the frequency domain. For
this, on test rigs or in simulation the force element is excited by sine waves

xe (t) = A sin(2 f t) ,
with dierent frequencies

f0 f fE

(4.5)

and amplitudes

Amin A Amax .

Starting at

t = 0,

the system will usually be in a steady state condition after several periods

where

T = 1/f

and

n = 2, 3, . . .

t nT ,

have to be chosen appropriately. Due to the nonlinear

F (t + T ) = F (T ), where T = 1/f , yet


F will be approximated

system behavior the system response is periodic,

not harmonic. That is why, the measured or calculated force


within one period

n T t (n + 1)T ,

by harmonic functions as good as possible

sin(2 f t) + cos(2 f t) .
F (t)
|
{z
}
|{z}
rst harmonic approximation
measured/
calculated

The coecients

1
2

(n+1)T
Z

and

(4.6)

can be calculated from the demand for a minimal overall error

2
sin(2 f t)+ cos(2 f t) F (t) dt

M inimum .

(4.7)

nT
The dierentiation of Eq. (4.7) with respect to

and

yields two linear equations as

necessary conditions

(n+1)T
Z

nT
(n+1)T
Z


sin(2 f t)+ cos(2 f t) F (t) sin(2 f t) dt = 0
(4.8)


sin(2 f t)+ cos(2 f t) F (t) cos(2 f t) dt = 0

nT
with the solutions

R
R
R
F sin dt cos2 dt F cos dt sin cos dt
R
R
R
=
sin2 dt cos2 dt 2 sin cos dt
,
R
R
R
R
F cos dt sin2 dt F sin dt sin cos dt
R
R
R
=
sin2 dt cos2 dt 2 sin cos dt

(4.9)

where the integral limits and arguments of sine and cosine no longer have been written.
Because it is integrated exactly over one period

nT t (n + 1)T ,

for the integrals in

Eq. (4.9)

sin cos dt = 0 ;

sin2 dt =

R
T
T
;
cos2 dt =
2
2

(4.10)

63

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

hold, and as solution

2
=
T

2
=
T

F sin dt ,

Z
F cos dt

(4.11)

remains. However, these are exactly the rst two coecients of a Fourier-"-Approximation.

The rst order harmonic approximation in Eq. (4.6) can now be written as

F (t) = F sin (2 f t + )
where amplitude

and phase angle

F =

(4.12)

are given by

2 + 2

tan =

and

A simple force element consisting of a linear spring with the stiness


with the constant

(4.13)

c and a linear damper

in parallel would respond with

F (t) = c xe + d x e = c A sin 2f t + d 2f A cos 2f t .

(4.14)

Here, amplitude and phase angle are given by

F = A

c2 + (2f d)2

Hence, the response of a pure spring,

tan =

and

c 6= 0

and

d=0

d
d 2f A
= 2f .
cA
c

(4.15)

F = A c and
d 6= 0 results in

is characterized by

tan = 0 or = 0, whereas a pure damper response with c = 0 and


F = 2f dA and tan or = 90 . Hence, the phase angle which

is also called

the dissipation angle can be used to evaluate the damping properties of the force element.
The dynamic stiness, dened by

cdyn =

F
A

(4.16)

is used to evaluate the stiness of the element.


In practice the frequency response of a system is not determined punctually, but continuously. For this, the system is excited by a sweep-sine. In analogy to the simple sine-function

xe (t) = A sin(2 f t) ,
where the period

T = 1/f

(4.17)

appears as pre-factor at dierentiation

x e (t) = A 2 f cos(2 f t) =

2
A cos(2 f t) .
T

(4.18)

A generalized sine-function can be constructed, now. Starting with

xe (t) = A sin(2 h(t)) ,

64

(4.19)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

the time derivative results in

x e (t) = A 2 h(t)
cos(2 h(t)) .
In the following we demand that the function
time, i.e:

h(t)
=
where

p>0

and

q>0

generates periods fading linearly in

1
1
=
,
T (t)
pqt

(4.21)

are constants yet to determine. Eq. (4.21) yields

h(t) =
The initial condition

h(t)

(4.20)

h(t = 0) = 0

1
ln(p q t) + C .
q

(4.22)

xes the integration constant

C =

1
ln p .
q

(4.23)

With Eqs. (4.23) and (4.22) Eq. (4.19) results in a sine-like function

xe (t) = A sin

 2
q

ln

p 
,
pqt

(4.24)

which is characterized by linear fading periods.


The important zero values for determining the period duration lie at

1
p
ln
= 0, 1, 2,
q
p q tn
and

tn =

or

p
= en q ,
p q tn

mit

n = 0, 1, 2,

p
(1 en q ) , n = 0, 1, 2, .
q

(4.25)

(4.26)

The time dierence between two zero values yields the period

p
Tn = tn+1 tn = (1e(n+1) q 1+en q )
q
, n = 0, 1, 2, .
p n q
q
Tn =
e
(1 e )
q
For the rst

(n = 0)

and last

(n = N )

(4.27)

period one nds

p
(1 eq )
q
.
p
=
(1 eq ) eN q = T0 eN q
q

T0 =
TN

With the frequency range to investigate, given by the initial


the parameters

and the ratio

q/p

f0

(4.28)

and nal frequency

fE ,

can be calculated from Eq. (4.28)

1
fE
q =
ln
,
N
f0

n
hf iN o
q
E
= f0 1
,
p
f0

(4.29)

65

Vehicle Dynamics

with

FH Regensburg, University of Applied Sciences

xing the number of frequency intervals. The passing of the whole frequency range

then takes the time

1 e(N +1) q
.
q/p

tN +1 =

(4.30)

Hence, to test or simulate a force element in the frequency range from


with

N = 500

intervals will only take

728 s

or

0.1Hz to f = 100Hz

12min.

4.5.2 Simple Spring Damper Combination


Fig. 4.13 shows a simple dynamic force element where a linear spring with the stiness
and a linear damper with the damping constant

are arranged in series.

Figure 4.13: Spring and damper in series


The displacements of the force element and the spring itself are described by

and

s.

Then, the the forces acting in the spring and damper are given by

FS = c s
FD = FS
displacement s

The force balance


spring

and

FD = d (u s)
.

results in a linear rst order dierential equation for the

d (u s)
= cs

or

where the ratio between the damping coecient


constant,

(4.31)

d
d
s = s + u ,
c
c
d

and the spring stiness

(4.32)

acts as time

T = d/c. Hence, this force element will responds dynamically to any excitation.

The steady state response to a harmonic excitation

u(t) = u0 sin t

respectively

u = u0 cos t

(4.33)

can be calculated easily. The steady state response will be of the same type as the excitation. Inserting

s (t) = u0 (a sin t + b cos t)

(4.34)

into Eq. (4.32) results in

d
d
u0 (a cos t b sin t) = u0 (a sin t + b cos t) +
u0 cos t .
|
{z
}
|
{z
}
c
c | {z }
s
u
s

66

(4.35)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

Collecting all sine and cosine terms we obtain two equations

d
u0 b = u0 a
c

d
d
u0 a = u0 b +
u0
c
c

and

(4.36)

which can be solved for the two unknown parameter

2
2 + (c/d)2

a =

and

b =

.
2
d + (c/d)2

(4.37)

Hence, the steady state force response reads as

FS = c s



c

sin t +
= c u0 2
cos t
d
+ (c/d)2

(4.38)

which can be transformed to

FS = FS sin (t + )
where the force magnitude

FS =

FS

and the phase angle

p
c u0
c u0
2 + (c/d)2 = p
2
2
2
+ (c/d)
+ (c/d)2

The dynamic stiness

cdyn = FS /u0

(4.39)

are given by
and

and the phase angle

= arctan

c/d
.

(4.40)

are plotted in Fig. 4.14 for

dierent damping values.

cdyn

400

4
300

[N/mm]

200

c = 400 N/mm

2
1

100
0
100

d
d1 = 1000 N/(m/s)
d2 = 2000 N/(m/s)

[o]
50

d1 = 3000 N/(m/s)
d2 = 4000 N/(m/s)

3
4
0

20

40

60

80

100

f [Hz]
Figure 4.14: Frequency response of a spring damper combination
With increasing frequency the spring damper combination changes from a pure damper

performance, cdyn 0 and 90 to a pure spring behavior, cdyn c and 0. The


frequency range, where the element provides stiness and damping is controlled by the
value for the damping constant

d.

67

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

4.5.3 General Dynamic Force Model


To approximate the complex dynamic behavior of bushings and elastic mounts dierent
spring damper models can be combined. A general dynamic force model is constructed
by

parallel force elements, Fig. 4.15. The static load is carried by a single spring with

the stiness

c0

F0 = F0 (u).

or an arbitrary nonlinear force characteristics

u
FM
F1

d1

d2

FM
F2

dN

FM
FN

c0
s2

s1
c1

sN
cN

c2

Figure 4.15: Dynamic force model


Within each force element the spring acts in serial to parallel combination of a damper
and a dry friction element. Now, even hysteresis eects and the stress history of the force
element can be taken into account.
The forces acting in the spring and damper of force element

FSi = ci si
were

and

si

and

are given by

FDi = di (s i u)
,

(4.41)

describe the overall element and the spring displacement.

As long as the absolute value of the spring force FSi is lower than the maximum friction
M
force FF the damper friction combination will not move at all

u s i = 0

for

|FSi | FFM .

(4.42)

In all other cases the force balance

FSi = FDi FFM

(4.43)

holds. Using Eq. 4.41 the force balance results in

di (s i u)
= FSi FFM

(4.44)

which can be combined with Eq. 4.42 to

F + FFM

Si
di s i =
di u

F F M
Si

68

FSi < FFM


for

FFM FSi +FFM


+FFM < FSi

(4.45)

FH Regensburg, University of Applied Sciences

where according to Eq. 4.41 the spring force is given by

Prof. Dr.-Ing. G. Rill

FSi = ci si .

In extension to this linear approach nonlinear springs and dampers may be used. To derive
all the parameters an extensive set of static and dynamic measurements is needed.

4.5.3.1 Hydro-Mount
For the elastic suspension of engines in vehicles very often specially developed hydromounts are used. The dynamic nonlinear behavior of these components guarantees a
good acoustic decoupling but simultaneously provides sucient damping.

xe
main spring
chamber 1
membrane

cF

cT
__
2

ring channel
chamber 2

uF

c__
T
2

MF
dF
__
2

dF
__
2

Figure 4.16: Hydro-mount


Fig. 4.16 shows the principle and mathematical model of a hydro-mount. At small deformations the change of volume in chamber 1 is compensated by displacements of the
membrane. When the membrane reaches the stop, the liquid in chamber 1 is pressed
through a ring channel into chamber 2. The ratio of the chamber cross section to the ring
channel cross section is very large. Thus the uid is moved through the ring channel at
very high speed. This results in remarkable inertia and resistance forces (damping forces).
The force eect of a hydro-mount is combined from the elasticity of the main spring and
the volume change in chamber 1.
With

uF

labeling the displacement of the generalized uid mass

FH = cT xe + FF (xe uF )

MF ,
(4.46)

holds, where the force eect of the main spring has been approximated by a linear spring
with the constant

cT .

69

Vehicle Dynamics

With

MF R

FH Regensburg, University of Applied Sciences

as the actual mass in the ring channel and the cross sections

A K , AR

of

chamber and ring channel the generalized uid mass is given by

MF =

 A 2
K

AR

MF R .

(4.47)

The uid in chamber 1 is not being compressed, unless the membrane can evade no longer.
With the uid stiness

FF (xe uF ) =

cF and the membrane clearance sF , one gets





c
(x

u
)
+
s
(xe uF ) < sF
F
e
F
F

0 


c (x u ) s
F
e
F
F

The hard transition from clearance


tion with

FF 6= 0

for

|xe uf |

sF

(4.48)

(xe uf ) > +sF

FF = 0 and uid compression resp. chamber deforma-

is not realistic and leads to problems, even with the numeric solution.

Therefore, the function (4.48) is smoothed by a parabola in the range

|xe uf | 2 sF .

The motions of the uid mass cause friction losses in the ring channel, which are as a rst
approximation proportional to the speed,

FD = dF u F .

(4.49)

Then, the equation of motion for the uid mass reads as

MF uF = FF FD .

(4.50)

The membrane clearance makes Eq. (4.50) nonlinear and only solvable by numerical integration. The nonlinearity also aects the overall force in the hydro-mount, Eq. (4.46).
The dynamic stiness and the dissipation angle of a hydro-mount are displayed in Fig. 4.17
versus the frequency.
The simulation is based on the following system parameters

mF

25 kg

generalized uid mass

cT

125 000 N/m

stiness of main spring

dF

750 N/(m/s)

damping constant

cF

100 000 N/m

uid stiness

sF

0.0002 mm

clearance in membrane bearing

By the nonlinear and dynamic behavior a very good compromise can be achieved between
noise isolation and vibration damping.

70

FH Regensburg, University of Applied Sciences

400

Prof. Dr.-Ing. G. Rill

Dynamic Stiffness [N/m] at Excitation Amplitudes A = 2.5/0.5/0.1 mm

300
200
100
0
60

Dissipation Angle [deg] at Excitation Amplitudes A = 2.5/0.5/0.1 mm

50
40
30
20
10
0

10

Excitation Frequency [Hz]

10

Figure 4.17: Dynamic stiness [N/mm] and dissipation angle [deg] for a hydro-mount

71

5 Vertical Dynamics
5.1 Goals
The aim of vertical dynamics is the tuning of body suspension and damping to guarantee
good ride comfort, resp. a minimal stress of the load at sucient safety.
The stress of the load can be judged fairly well by maximal or integral values of the body
accelerations.
The wheel load

Fz

is linked to the longitudinal

of friction. The digressive inuence of


at the increase of

Fx

and

Fy

Fz

on

Fx

Fx

and

and lateral force

Fy

Fy

by the coecient

as well as non-stationary processes

in the average lead to lower longitudinal and lateral forces

at wheel load variations.


Maximal driving safety can therefore be achieved with minimal variations of the wheel
load. Small variations of the wheel load also reduce the stress on the track.
The comfort of a vehicle is subjectively judged by the driver. In literature dierent approaches of describing the human sense of vibrations by dierent metrics can be found.
Transferred to vehicle vertical dynamics, the driver primarily registers the amplitudes and
accelerations of the body vibrations. These values are thus used as objective criteria in
practice.

5.2 Modelling Aspects


5.2.1 Full Vehicle Model
For detailed investigations of ride comfort and ride safety sophisticated road and vehicle
models are needed. The three-dimensional vehicle model, shown in Fig. 5.1, includes
an elastically suspended engine, and dynamic seat models. The elasto-kinematics of the
wheel suspension was described fully nonlinear. In addition, dynamic force elements for
the damper elements and the hydro-mounts are used. Such sophisticated models not only
provide simulation results which are in good conformity to measurements but also make
it possible to investigate the vehicle dynamic attitude in an early design stage.

72

Ford

FH Regensburg, University of Applied Sciences


Time =

0.000000

Prof. Dr.-Ing. G. Rill

ZZ

Y
Y

X
X

Figure 5.1: Full vehicle model

Thilo Seibert
Ext. 37598
Vehicle Dynamics, Ford Research Center Aachen

/export/ford/dffa089/u/tseiber1/vedyna/work/results/mview.mvw

07/02/98
AA/FFA

5.2.2 Twodimensional Models

Much simpler models can be used, however, for basic studies on ride comfort and ride
safety. A two-dimensional vehicle model, for instance, suits perfectly with a single track
road model, Fig. 5.2. Neglecting longitudinal accelerations, the vehicle chassis only per-

a2
a1

M2

zC1
M1

M*

zB

zA2
m2

M,

zR(s-a2)

zA1
zR(s+a1)

zC2

pitch

m1

hub
C
yB

xB
zR(s)

Figure 5.2: Vehicle model for basic comfort and safety analysis
forms hub and pitch motions. Here, the chassis is considered as one rigid body. Then,
mass and inertia properties can be represented by three point masses which are located in
the chassis center of gravity and on top of the front and the rear axle. The lumped mass
model has 4 degrees of freedom. The hub and pitch motion of the chassis are represented
by the vertical motions of the chassis in the front

zA1

and

zA2

zC1

zC2 . The coordinates


axle. The function zR (s)

and in the rear

describe the vertical motions of the front and rear

provides road irregularities in the space domain, where

denotes the distance covered by

the vehicle and measured at the chassis center of gravity. Then, the irregularities at the
front and rear axle are given by

zR (s a2 )
gravity C .

zR (s + a1 )

locate the position of the chassis center of

and

respectively, where

a1

and

a2

73

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The point masses must add up to the chassis mass

M1 + M + M2 = M

(5.1)

and they have to provide the same inertia around an axis located in the chassis center

and pointing into the lateral direction

a21 M1 + a22 M2 = .

(5.2)

The correct location of the center of gravity is assured by

a1 M 1 = a2 M 2 .

(5.3)

Now, Eqs. (5.2) and (5.3) yield the main masses

M1 =

a1 (a1 +a2 )

and

M2 =

,
a2 (a1 +a2 )

(5.4)

and the coupling mass

M = M

1
M a1 a2


(5.5)

follows from Eq. (5.1).

= M a1 a2 then,

the coupling mass vanishes M = 0, and the vehicle can be represented by two uncoupled

If the mass and the inertia properties of a real vehicle happen to result in

two mass systems describing the vertical motion of the axle and the hub motion of the
chassis mass on top of each axle.
mid

full

sports

size

size

utility

car

car

vehicle

m1 [kg]

80

100

125

120

600

m2 [kg]

80

100

125

180

1100

a1 [m]
a2 [m]

1.10
1.40

1.40
1.40

1.45
1.38

1.90
1.40

2.90
1.90

M [kg]

1100

1400

1950

3200

14300

[kg m2 ]

1500

2350

3750

5800

50000

M1
M [kg]
M2

545
126
429

600
200
600

914
76
960

925
1020
1255

3592
5225
5483

vehicles
properties
front axle
mass
rear axle
mass
center
of
gravity
chassis
mass
chassis
inertia
lumped
mass
model

commercial

heavy

vehicle

truck

Table 5.1: Mass and inertia properties of dierent vehicles

74

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

Depending on the actual mass and inertia properties the vertical dynamics of a vehicle
can be investigated by two simple decoupled mass models describing the vibrations of the
front and rear axle and the corresponding chassis masses. By using half of the chassis and
half of the axle mass we nally end up in quarter car models.
The data in Table 5.1 show that for a wide range of passenger cars the coupling mass is
smaller than the corresponding chassis masses,

M < M1

and

M < M2 .

Here, the two

mass model or the quarter car model represent a quite good approximation to the lumped
mass model. For commercial vehicles and trucks, where the coupling mass has the same
magnitude as the corresponding chassis masses, the quarter car model serves for basic
studies only.

5.2.3 Simple Models


At most vehicles, c.f. Table 5.1, the axle mass is much smaller than the corresponding
chassis mass,

mi  Mi , i = 1, 2.

Hence, for a rst basic study axle and chassis motions

can be investigated independently. The quarter car model is now further simplied to two
single mass models, Fig. 5.3.

``

6
zC

M
`

cS ``
`
`

`
cS `
`

dS

dS

6
zW

`
c

6
zR

``
`` c
`` T
c

6
zR

Figure 5.3: Simple vertical vehicle models


The chassis model neglects the tire deection and the inertia forces of the wheel. For the
high frequent wheel motions the chassis can be considered as xed to the inertia frame.
The equations of motion for the models read as

M zC + dS zC + cS zC = dS zR + cS zR

(5.6)

m zW + dS zW + (cS + cT ) zW = cT zR ,

(5.7)

and

where

zC

and

zW

label the vertical motions of the corresponding chassis mass and the

wheel mass with respect to the steady state position. The constants

cS , dS

describe the

suspension stiness and damping. The dynamic wheel load is calculated by

FTD = cT (zR zW )

(5.8)

75

Vehicle Dynamics

where

cT

FH Regensburg, University of Applied Sciences

is the vertical or radial stiness of the tire and

zR

denotes the road irregularities.

In this simple approach the damping eects in the tire are not taken into account.

5.3 Basic Tuning


5.3.1 Natural Frequency and Damping Rate
At an ideally even track the right side of the equations of motion (5.6), (5.7) vanishes
because of

zR = 0

zR = 0.

and

The remaining homogeneous second order dierential

equations can be written in a more general form as

z + 2 0 z + 02 z = 0 ,

(5.9)

where 0 represents the undamped natural frequency, and is a dimensionless parameter


called viscous damping ratio. For the chassis and the wheel model the new parameter are
given by

Chassis:

Wheel:

z zC ,

C =

z zW ,

W =

d
S
,
2 cS M
dS
2

(cS +cT )m

2
02 0C
=

2
02 0W
=

cS
;
M
cS +cT
.
m
(5.10)

The solution of Eq. (5.9) is of the type

z(t) = z0 et ,
where

z0

and

(5.11)

are constants. Inserting Eq. (5.11) into Eq. (5.9) results in

(2 + 2 0 + 02 ) z0 et = 0 .
Non-trivial solutions

z0 6= 0

(5.12)

are possible, if

2 + 2 0 + 02 = 0
will hold. The roots of the characteristic equation (5.13) depend on the value of

<1 :
1 :

p
1,2 = 0 pm i 0 1 2 ,


p
2
1,2 = 0 1 .

(5.13)

(5.14)

Figure 5.4 shows the root locus of the eigenvalues for dierent values of the viscous
damping rate

76

FH Regensburg, University of Applied Sciences

=0

=0.5

=0.2

=0.7

-3

=1.25

-2.5

-2

-1

-1.5

-0.5

=0.9
=0.2

=0.5

1 the eigenvalues 1,2

Re()/0

-0.5

=0.7

For

1.0

=1.25 =1.5

=1

Figure 5.4: Eigenvalues

Im()/0

0.5

=0.9
=1.5

Prof. Dr.-Ing. G. Rill

and

=0

-1.0

for dierent values of

are both real and negative. Hence, Eq. (5.11) will produce a

< 1 holds, the eigenvalues 1,2 will become


complex conjugate of 1 . Now, the solution can be written as

 p
0 t
2
z(t) = A e
sin 0 1 t ,

exponentially decaying solution. If


where

is the

complex,

(5.15)

A and are constants which have to be adjusted to given initial conditions z(0) = z0
and z(0)

= z0 . The real part Re (1,2 ) =


p 0 is negative and determines the decay of
1 2 part denes the actual frequency of
the solution. The imaginary Im (1,2 ) = 0

where

the vibration. The actual frequency

= 0
tends to zero,

0,

1 2

if the viscous damping ratio will approach the value one,

(5.16)

1.

In a more general way the relative damping may be judged by the ratio

Re(1,2 )
.
| 1,2 |

D =

(5.17)

For complex eigenvalues which characterize vibrations

D =

(5.18)

holds, because the absolute value of the complex eigenvalues is given by

| 1,2 | =

r
Re(1,2 )2 + Im(1,2 )2 =

 p
2
( 0 )2 + 0 1 2
= 0 ,

(5.19)

and hence, Eq. (5.17) results in

D =
For

+ 0
= .
0

(5.20)

1 the eigenvalues become real and negative. Then, Eq. (5.17) will always produce
D = 1. In this case the viscous damping rate is more

the relative damping value


sensitive.

77

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

5.3.2 Spring Rates


5.3.2.1 Minimum Spring Rates
The suspension spring is loaded with the corresponding vehicle weight. At linear spring
characteristics the steady state spring deection is calculated from

u0 =

Mg
.
cS

(5.21)

At a conventional suspension without niveau regulation a load variation


leads to changed spring deections

In analogy to (5.21) the additional

4M g
.
cS

(5.22)

deection follows from

4u =
If for the maximum load variation

M M + 4M

u0 u0 + 4u.

4M

the additional spring deection is limited to

4u

the suspension spring rate can be estimated by a lower bound

cS

4M g
.
4u

(5.23)

In the standard design of a passenger car the engine is located in the front and the trunk
in the rear part of the vehicle. Hence, most of the load is supported by the rear axle
suspension.
For an example we assume that

150 kg

wheel by

500 kg are going to the


4MF = 150 kg/2 = 75 kg and each rear

of the permissible load of

front axle. Then, each front wheel is loaded by

4MR = (500 150) kg/2 = 175 kg .


u umax . At standard passenger cars it
umax 0.10 m. By setting 4u = umax /2 we demand

The maximum wheel travel is limited,


range of

umax 0.8 m

to

is in the
that the

spring deection caused by the load should not exceed half of the maximum value. Then,
according to Eq. (5.23) a lower bound of the spring rate at the front axle can be estimated
by

cmin
= ( 75 kg 9.81 m/s2 )/(0.08/2) m = 18400 N/m .
S

(5.24)

4MR = 175 kg . Assuming


umax 0.10 m the minimum

The maximum load over one rear wheel was estimated here by
that the suspension travel at the rear axle is slightly larger,
spring rate at the rear axle can be estimated by

cmin
= ( 175 kg 9.81 m/s2 )/(0.10/2) m = 34300 N/m ,
S

(5.25)

which is nearly two times the minimum value of the spring rate at the front axle. In order
to reduce this dierence we will choose a spring rate of

cS = 20 000 N/m at the front axle.

In Tab. 5.1 the lumped mass chassis model of a full size passenger car is described by

M1 = M2 = 600 kg

and

M = 200.

To approximate the lumped mass model by two

decoupled two mass models we have to neglect the coupling mass or, in order to achieve

78

FH Regensburg, University of Applied Sciences

the same chassis mass, to distribute

Prof. Dr.-Ing. G. Rill

equally to the front and the rear. Then, the

corresponding cassis mass of a quarter car model is given here by


M = M1 + M /2 /2 = (600 kg + 200/2 kg)/2 = 350 kg .

(5.26)

According to Eq. 5.10 the undamped natural eigen frequency of the simple chassis model
2
is then given by 0C = cS /M . Hence, for a spring rate of cS = 20000 N/m the undamped
natural frequency of the unloaded car amounts to

f0C =

20000 N/m 350 kg/(2 ) = 1.2 Hz ,

(5.27)

which is a typical value for most of all passenger cars. Due to the small amount of load
the undamped natural frequency for the loaded car does not change very much,

f0C =

p
20000 N/m (350 + 75) kg/(2 ) = 1.1 Hz .

(5.28)

The corresponding cassis mass over the rear axle is given here by


M = M2 + M /2 /2 = (600 kg + 200/2 kg)/2 = 350 kg .

(5.29)

Now the undamped natural frequencies for the unloaded

0
f0C
=

p
34300 N/m/350 kg/(2 ) = 1.6 Hz

(5.30)

and the loaded car

L
f0C
=

34300 N/m/(350 + 175) kg/(2 ) = 1.3 Hz

(5.31)

are larger and dier more.

5.3.2.2 Nonlinear Springs


In order to reduce the spring rate at the rear axle and to avoid too large spring deections
when loaded nonlinear spring characteristics are used, Fig. 5.5. Adding soft bump stops
the overall spring force in the compression mode

u0

can be modeled by the nonlinear

function

FS = FS0 + c0 u
where

FS0

is the spring preload,

cS


1+k

u
4u

2 !
,

describes the spring rate at

(5.32)

u = 0,

k > 0 characat u = 4u the

and

terizes the intensity of the nonlinearity. The linear characteristic provides


FSlin (4u) = FS0 + cS 4u. To achieve the same value with the nonlinear spring

value

FS0 + c0 4u (1 + k) = FS0 + cS 4u
must hold, where

cS

or

c0 (1 + k) = cS

(5.33)

describes the spring rate of the corresponding linear characteristics.

The local spring rate is determined by the derivative

dFS
= c0
du


1 + 3k

u
4u

2 !
.

(5.34)

79

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

FS

dFS
du

8000
u=u

FS [N]
6000
M g

63 kN/m
44 kN/m

cS
FS0

dFS
du

4000
29 kN/m

u=0

20 kN/m
2000

0.05

u [m]

0.1

Figure 5.5: Principle and realizations of nonlinear spring characteristics

Hence, the spring rate for the loaded car at

u = 4u

is given by

cL = c0 (1 + 3 k) .
The intensity of the nonlinearity

(5.35)

can be xed, for instance, by choosing an appropriate

spring rate for the unloaded vehicle. With

c0 = 20000 N/m

the spring rates on the front

and rear axle will be the same for the unloaded vehicle. With
yields

k =

cS = 34300 N/m

Eq. (5.33)

34300
cS
1 = 0.715 .
1 =
c0
20000

(5.36)

The solid line in Fig. 5.5 shows the resulting nonlinear spring characteristics which is
characterized by the spring rates

3 0.715) = 62 900 N/m

c0 = 20 000 N/m

and

cL = c0 (1 + 3k) = 20 000 (1 +

for the unloaded and the loaded vehicle. Again, the undamped

natural frequencies

s
0
f0C
=

20000 N/m 1
= 1.20 Hz
350 kg
2

s
or

L
f0C
=

92000 N/m 1
= 1.74 Hz
(350+175) kg 2

(5.37)

for the unloaded and the loaded vehicle dier quite a lot.
The unloaded and the loaded vehicle have the same undamped natural frequencies if

c0
cL
=
M
M + 4M

or

cL
M + 4M
=
c0
M

(5.38)

will hold. Combing this relationship with Eq. (5.35) one obtains

M
1 + 3k =
M + 4M

80

or

1
k =
3


M + 4M
1 4M
1 =
.
M
3 M

(5.39)

FH Regensburg, University of Applied Sciences

Hence, for the quarter car model with


sity of the nonlinear spring amounts to

Prof. Dr.-Ing. G. Rill

M = 350 kg and 4M = 175


k = 1/3 175/350 = 0.1667.

the intenThis value

cS = 34300 N/m will produce the dotted line in Fig. 5.5. The spring rates
c0 = cS /(1 + k) = 34 300 N/m / (1 + 0.1667) = 29 400 N/m for the unloaded and
cL = c0 (1 + 3k) = 29 400 N/m (1 + 3 0.1667) = 44 100 N/m for the loaded vehicle fol-

and

low from Eqs. (5.34) and (5.35). Now, the undamped natural frequency for the unloaded
p
p
0
0
= cL /(M + 4M ) = 1.46 Hz are
f0C
= c0 /M = 1.46 Hz and the loaded vehicle f0C
in deed the same.

5.3.3 Influence of Damping


To investigate the inuence of the suspension damping to the chassis and wheel motion the
simple vehicle models are exposed to initial disturbances. Fig. 5.6 shows the time response
and wheel displacement zW (t) as well as the chassis acceleration z
C
0
D
and the wheel load FT = FT + FT for dierent damping rates C and W . The dynamic
0
wheel load follows from Eq. (5.8), and the static wheel load is given by FT = (M + m) g ,

of the chassis

where

zC (t)

labels the constant of gravity.

To achieve the same damping rates for the chassis and the wheel model dierent values
for the damping parameter

dS

were needed.

With increased damping the overshoot eect in the time history of the chassis displacement and the wheel load becomes smaller and smaller till it vanishes completely at
and

W = 1.

The viscous damping rate

C = 1

=1

5.3.4 Optimal Damping


5.3.4.1 Avoiding Overshoots
If avoiding overshoot eects is the design goal then,
ratio. For

=1

=1

will be the optimal damping

the eigenvalues of the single mass oscillator change from complex to real.

Thus, producing a non oscillating solution without any sine and cosine terms.

C = 1 and W = 1 results in the optimal damping parameter


p
p
C =1
W =1


dopt
=
2
cS M , and dopt
(cS +cT )m .
(5.40)
S Comfort
S Safety = 2

According to Eq. (5.10)

So, the damping values

C =1
N

dopt
S Comfort = 5292
m/s

and

W =1
N

dopt
S Safety = 6928
m/s

will avoid an overshoot eect in the time history of the chassis displacement
in the time history of the wheel load

FT (t).

(5.41)

zC (t) or in the

Usually, as it is here, the damping values for

optimal comfort and optimal ride safety will be dierent. Hence, a simple linear damper
can either avoid overshoots in the chassis motions or in the wheel loads.

81

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

chassis model
350 kg
dS

20000 N/m

wheel model

C [ - ]

dS [Ns/m]

W [ - ]

dS [Ns/m]

0.25
0.50
0.75
1.00
1.25

1323
2646
3969
5292
6614

0.25
0.50
0.75
1.00
1.25

1732
3464
5196
6928
8660

displacement [mm]
200

15

100

10

50

-50

-5

-100

50 kg
220000 N/m

displacement [mm]

20

150

20000 N/m

dS

-10
0

0.5

t [s]

0.05

0.1

t [s] 0.15

0.1

t [s]

wheel load [N]

acceleration [g]

1.5
6000
5000

0.5
4000
0

3000

2000

-0.5

1000
-1

0
0

0.5

t [s]

1.5

0.05

0.15

Figure 5.6: Time response of simple vehicle models to initial disturbances

The overshot in the time history of the chassis accelerations

zC (t)

will only vanish for

which surely is not a desirable conguration, because then, it takes a very long

time till the initial chassis displacement has fully disappeared.

5.3.4.2 Fast Approach to Steady State


Instead of avoiding overshoot eects we better demand that the time history of the system
response will approach the steady state value as fast as possible. Fig. 5.7 shows the typical
time response of a damped single-mass oscillator to the initial disturbance
and

82

z(t
= 0) = 0.

z(t = 0) = z0

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

z0
zS
tE

z(t)

Figure 5.7: Evaluating a damped vibration

Counting the dierences of the system response

z(t) from the steady state value zS = 0 as

errors allows to judge the attenuation. If the overall quadratic error becomes a minimum

t=t
Z E

 =

z(t)2 dt M in ,

(5.42)

t=0
the system approaches the steady state position as fast as possible. In theory

tE

holds, for practical applications a nite

tE

have to be chosen appropriately.

To judge ride comfort and ride safety the hub motion of the chassis zC , its acceleration
zC and the variations of the dynamic wheel load FTD can be used. In the absence of road
D
irregularities zR = 0 the dynamic wheel load from Eq. (5.8) simplies to FT = cT zW .
Hence, the demands

2C

t=t
Z E

2

g1 zC

g2 zC

2 i

dt M in

(5.43)

t=0
and

t=t
Z E

2S =

cT zW

2

dt M in

(5.44)

t=0
will guarantee optimal ride comfort and optimal ride safety. By the factors

g1

and

g2

the

acceleration and the hub motion can be weighted dierently.


The equation of motion for the chassis model can be resolved for the acceleration


2
zC + 2C zC ,
zC = 0C
where, the system parameter

C 0C = dS /(2M )

M , dS

and

cS

(5.45)

were substituted by the damping rate

and by the undamped natural frequency

0C = cS /M .

C =

Then, the

problem in Eq. (5.43) can be written as

2C =

t=t
Z E

t=0
t=t
Z E

=
t=0

2
g12 0C
zC + 2C zC

zC zC

g12

2

2 2
(0C
)

g12

2
0C

i
+ g22 zC2 dt

g22

2C

g12
g12

2
0C

2C
2

(2C )

zC
zC

(5.46)

M in ,

83

Vehicle Dynamics

where

FH Regensburg, University of Applied Sciences



xTC = zC zC

is the state vector of the chassis model. In a similar way Eq. (5.44)

can be transformed to

2S

t=t
Z E

c2T

2
zW

t=t
Z E

dt =

zW zW

c2T 0
0 0

#"

t=0

t=0
where

"


xTW = zW zW

zW
zW

#
M in ,

(5.47)

denotes the state vector of the wheel model.

The problems given in Eqs. (5.46) and (5.47) are called disturbance-reaction problems.
They can be written in a more general form

t=t
Z E

xT (t) Q x(t) dt M in

(5.48)

t=0
where

x(t)

denotes the state vector and

Q = QT

is a symmetric weighting matrix. For

single mass oscillators described by Eq. (5.9) the state equation reads as


 

0
1
z
z
=
.
2
0 2
z
z
|
{z
} | {z }
| {z }
x
x
A


(5.49)

the time response of the system exposed to the initial disturbance x(t = 0) =
x0 vanishes x(t ) = 0, and the integral in Eq.(5.48) can be solved by

For tE

t=t
Z E

xT (t) Q x(t) dt = xT0 R x0 ,

(5.50)

t=0
where the symmetric matrix

R = RT

is given by the Ljapunov equation

AT R + R A + Q = 0 .

(5.51)

For the single mass oscillator the Ljapunov equation

0 02
1 2



R11 R12
R12 R22


+

R11 R12
R12 R22



0
1
02 2


+

Q11 Q12
Q12 Q22


.

(5.52)

results in 3 linear equations

02 R12 02 R12 + Q11 = 0


02 R22 + R11 2 R12 + Q12 = 0
R12 2 R22 + R12 2 R22 + Q22 = 0
which easily can be solved for the elements of


R11 =

84

1
+
2
0 4


Q11 Q12 +

02
Q22 ,
4

(5.53)

R12 =

Q11
,
202

R22 =

Q11
Q22
+
.
2
4 0
4

(5.54)

FH Regensburg, University of Applied Sciences

For the initial disturbance

t=t
Z E

x0 = [ z0 0 ]T

x (t) Q x(t) dt =

z02

R11 =

Prof. Dr.-Ing. G. Rill

Eq. (5.50) nally results in

z02



t=0

+
2
0 4


02
Q11 Q12 +
Q22 .
4

(5.55)

Now, the integral in Eq. (5.46) evaluating the ride comfort is solved by

2C




2
2
C
1
0C
2
2
2
2
2 2
2
=
+
g1 0C + g2 g1 0C 2 C +
g (2C )
2
0C
4C
4C 1
#
"

2 ! 
2
g
g

2
2
0C
2
2
g12 +
+
C 0C .
= z0C
0C
2
2
4C
0C
0C
2
z0C



where the abbreviation


By setting

g1 = 1

and

was nally replaced by

g2 = 0

(5.56)

C 0C .

the time history of the chassis acceleration

zC

is weighted

only. Eq. (5.56) then simplies to


2
2 0C
2C z = z0C
0C
C
4C
which will become a minimum for

0C 0

(5.57)

C .

or

achieved by a soft suspension spring

cS 0

C
0C 0 is

As mentioned before,

surely is not a desirable conguration. A low undamped natural frequency


or a large chassis mass

M .

However, a

large chassis mass is uneconomic and the suspension stiness is limited by the the loading
conditions. Hence, weighting the chassis accelerations only does not lead to a specic
result for the system parameter.
The combination of

g1 = 0 and g2 = 1 weights the time history of the chassis displacement

only. Then, Eq. (5.56) results in

2C z
C

z2
= 0C
0C

1
+ C
4C

0C or


2
d 2C |zC
z0C
1
=
+1 = 0.
d C
0C 4C2

(5.58)

which will become a minimum for

A high undamped natural frequency

(5.59)

0C

contradicts the demand for rapidly van1


ishing accelerations. The viscous damping ratio C =
solves Eq. (5.59) and minimizes
2
the merit function in Eq. (5.58). But again, this value does not correspond with C
which minimizes the merit function in Eq. (5.57).
Hence, practical results can be achieved only if the chassis displacements and the chassis
accelerations will be evaluated simultaneously. To do so, appropriate weighting factors
have to be chosen. In the equation of motion for the chassis (5.6) the terms

cS zC

are added. Hence,

g1 = M

and

g1 = 1

g2 = cS
and

M zC

and

or

g2 =

cS
2
= 0C
M

(5.60)

85

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

provide system-tted weighting factors. Now, Eq. (5.56) reads as

2C

2
z0C

2
0C

0C
+ C 0C
2C

Again, a good ride comfort will be achieved by

0C 0.

(5.61)

For nite undamped natural

frequencies Eq. (5.61) becomes a minimum, if the viscous damping rate

d 2C |zC
2
2
0C
= z0C
d C

Hence, a viscous damping rate of

C =

0C
+ 0C
2C2

will satisfy


= 0.

(5.62)

1
2
2

(5.63)

or a damping parameter of

1
opt C = 2 2
dS Comfort

2 cS M ,

(5.64)

will provide optimal comfort by minimizing the merit function in Eq. (5.61).
For the passenger car with
parameter will amount to

which is

70%

M = 350 kg

and

cS = 20 000 N/m

the optimal damping

C = 12 2
N

dopt
S Comfort = 3742
m/s

(5.65)

of the value needed to avoid overshot eects in the chassis displacements.

The integral in Eq. (5.47) evaluating the ride safety is solved by

2S

z2
= 0W
0W


W

1
+
4W

c2T

(5.66)

where the model parameter m, cS , dS and cT where replaced by the undamped natural
2
frequency 0W = (cS + cT )/m and by the damping ratio W = W 0W = dS /(2m).
2
A soft tire cT 0 make the safety criteria Eq. (5.66) small S 0 and thus, reduces the
dynamic wheel load variations. However, the tire spring stiness can not be reduced to
arbitrary low values, because this would cause too large tire deformations. Small wheel
masses

m 0

and/or a hard body suspension

cS

will increase

0W

and thus,

reduce the safety criteria Eq. (5.66). The use of light metal rims improves, because of
wheel weight reduction, the ride safety of a car. Hard body suspensions contradict a good
driving comfort.
With xed values for
if

cT

and

0W

the merit function in Eq. (5.66) will become a minimum

2
2S
z0W
=
W
0W

1
1+ 2
4W

c2T = 0

(5.67)

will hold. Hence, a viscous damping rate of

W =

86

1
2

(5.68)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

or the damping parameter

p


dopt
(cS + cT ) m
S Safety =

(5.69)

will guarantee optimal ride safety by minimizing the merit function in Eq. (5.66).
For the passenger car with

M = 350 kg

and

cS = 20 000 N/m

the optimal damping

parameter will now amount to

W = 12
N

dopt
S Safety = 3464
m/s
which is

50%

(5.70)

of the value needed to avoid overshot eects in the wheel loads.

5.4 Sky Hook Damper


5.4.1 Modelling Aspects
In standard vehicle suspension systems the damper is mounted between the wheel and
the body. Hence, the damper aects body and wheel/axle motions simultaneously.

sky
zC

zC

dB

dW
cS
zW
zR

cS

dS

FD

zW

m
cT

zR

a) Standard Damper

m
cT

b) Sky Hook Damper

Figure 5.8: Quarter car model with standard and sky hook damper
To take this situation into account the simple quarter car models of section 5.2.3 must be
combined to a more enhanced model, Fig. 5.8a.
Assuming a linear characteristics the suspension damper force is given by

FD = dS (zW zC ) ,
where

dS denotes the damping constant, and zC , zW

(5.71)

are the time derivatives of the absolute

vertical body and wheel displacements.

87

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The sky hook damping concept starts with two independent dampers for the body and
the wheel/axle mass, Fig. 5.8b. A practical realization in form of a controllable damper
will then provide the damping force

FD = dW zW dC zC ,
dS

where instead of the single damping constant

(5.72)

now two design parameter

dW

and

dC

are available.
The equations of motion for the quarter car model are given by

M zC = FS + FD M g ,

(5.73)

m zW = FT FS FD m g ,
M, m
ments, and g
where

are the sprung and unsprung mass,

zC , zW

denote their vertical displace-

is the constant of gravity.

The suspension spring force is modeled by

FS = FS0 + cS (zW zC ) ,
where

FS0 = mC g

is the spring preload, and

cS

(5.74)

the spring stiness.

Finally, the vertical tire force is given by

FT = FT0 + cT (zR zW ) ,
where

FT0 = (M + m) g

cS

is the tire preload,

the road roughness. The condition

FT 0

(5.75)

the vertical tire stiness, and

zR

describes

takes the tire lift o into account.

5.4.2 Eigenfrequencies and Damping Ratios


Using the force denitions in Eqs. (5.72), (5.74) and (5.75) the equations of motion in
Eq. (5.73) can be transformed to the state equation

zC

zW

z
C
zW
| {z
x

= 0
cS

cS
M
cS
cS +cT
m
m

{z
A

dMC
dC
m

zC

h i


1
zW + 0 zR ,
0
dW
z
| {z }
M C
u
dW
cT
m
zW
m
} | {z }
| {z
}
x
B

(5.76)

M g , mg were compensated by the preloads FS0 , FT0 , the term B u


excitation, x denotes the state vector, and A is the state matrix. In this

where the weight forces


describes the

linear approach the tire lift o is no longer taken into consideration.


The eigenvalues

of the state matrix A will characterize the eigen dynamics of the quarter

car model. In case of complex eigenvalues the damped natural eigenfrequencies are given

88

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

Damping ratio = D

Frequencies [Hz]
12

10

Wheel

350 kg

0.8
0.7
0.6

dS 0.5

20000 N/m

50 kg

220000 N/m

Chassis

0.4
0.2

3880
3220

1000

2000

3000

4000

5000

1000

2000

3000

4000

5000

dS [N/(m/s)]

dS [N/(m/s)]
Figure 5.9: Quarter car model with standard damper

= Im(), and according to Eq. (vdyn-eq:


= D = Re()/ ||. evaluates the damping ratio.

by the imaginary parts,


ratio lambda)
By setting

dC = dS

and

dW = dS

relative damping

Eq. (5.76) represents a quarter car model with the

standard damper described by Eq. (5.71). Fig. 5.9 shows the eigenfrequencies
and the damping ratios

= D

f = /(2)
dS .

for dierent values of the damping parameter

C = 21 2 0.7 for the chassis motion


parameter dS = 3880 N/(m/s), and the damping

Optimal ride comfort with a damping ratio of


could be achieved with the damping
parameter

W = 0.5

dS = 3220 N/(m/s)

would provide for the wheel motion a damping ratio of

which correspond to minimal wheel load variations. This damping parameter

are very close to the values

3742 N/(m/s)

and

3464 N/(m/s)

which very calculated in

Eqs. (5.65) and (5.70) with the single mass models. Hence, the very simple single mass
models can be used for a rst damper layout. Usually, as it is here, optimal ride comfort
and optimal ride safety cannot achieved both by a standard linear damper.
The sky-hook damper, modeled by Eq. (5.72), provides with
rameter. Their inuence to the eigenfrequencies

dW

and

dS

two design pa-

and the damping ratios

is shown in

Fig. 5.10.
The the sky-hook damping parameter

dC , dW

the damping ratios. The chassis damping ratio


damping ratio

mainly depends on

dW .

have a nearly independent inuence on

mainly depends on

dC ,

and the wheel

Hence, the damping of the chassis and the

wheel motion can be adjusted to nearly each design goal. Here, a sky-hook damper with

dC = 3900 N/(m/s) and dW = 3200 N/(m/s) would generate the damping ratios dC = 0.7
and dW = 0.5 hence, combining ride comfort and ride safety within one damper layout.

89

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Damping ratios C, W

Frequencies [Hz]
12

10

0.8
0.7

350 kg

dC

50 kg
220000 N/m

2
0

1000

2000

3000

4000

4500
4000
3500
3000
2500
2000
1500

dC 0.6
dW 0.5

20000 N/m

dC [N/(m/s)]

5000

0.4
0.2
0

W
0

1000

2000

3000

4000

5000

dW [N/(m/s)]

dW [N/(m/s)]

Figure 5.10: Quarter car model with sky-hook damper

5.4.3 Technical Realization


By modifying the damper law in Eq. (5.72) to

FD = dW zW dC zC + =

dW zW dC zC
(zW zC ) = dS (zW zC )
zW zC
|
{z
}
dS

(5.77)

the sky-hook damper can be realized by a standard damper in the form of Eq. (5.71). The

new damping parameter dS now nonlinearly depends on the absolute vertical velocities of

the chassis and the wheel dS = dS (zC , zW ). As, a standard damper operates in a dissipative

mode only the damping parameter will be restricted to positive values, dS > 0. Hence, the
passive realization of a sky-hook damper will only match with some properties of the ideal
damper law in Eq. (5.72). But, compared with the standard damper it still can provide a
better ride comfort combined with an increased ride safety.

5.5 Nonlinear Force Elements


5.5.1 Quarter Car Model
The principal inuence of nonlinear characteristics on driving comfort and safety can
already be studied on a quarter car model Fig. 5.11.
The equations of motion read as

M zC = FS + FD M g
m zW = FT FS FD m g ,

g = 9.81 m/s2 labels the constant of gravity, M , m are the masses of the chassis
the wheel, FS , FD , FT describe the spring, the damper, and the vertical tire force,

where
and

90

(5.78)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

zC
nonlinear spring

nonlinear damper

FS

FD
FS

FD

zW

zR

cT

Figure 5.11: Quarter car model with nonlinear spring and damper characteristics

and the vertical displacements of the chassis

zC

and the wheel

zW

are measured from the

equilibrium position.
In extension to Eq. (5.32) the spring characteristics is modeled by

FS =

FS0

c0 u

c0 u

where

FS0 = M g


1 + kr

1 + kc

u
4ur

2 !

u
4uc

2 !

u<0
(5.79)

u0

is the spring preload, and

u = zW zC

(5.80)

u < 0 marks tension (rebound), and u 0 compression. Two sets of kr , ur and kc , uc dene the spring nonlinearity during rebound and
compression. For kr = 0 and kc = 0 a linear spring characteristics is obtained.

describes the spring travel. Here,

A degressive damper characteristics can be modeled by

FD (v) =

where

d0

d0 v
1 pr v

v <0,

d0 v
1 + pc v

v 0,

denotes the damping constant at

v = 0,

(5.81)

and the damper velocity is dened by

v = zW zC .

(5.82)

The sign convention of the damper velocity was chosen consistent to the spring travel.
Hence, rebound is characterized by

v<0

and compression by

v 0.

The parameter

pr

91

Vehicle Dynamics

and

pc

FH Regensburg, University of Applied Sciences

make it possible to model the damper nonlinearity dierently in the rebound and

compression mode. A linear damper characteristics is obtained with

pr = 0

and

pc = 0.

The nonlinear spring design in Section 5.3.2 holds for the compression mode. Hence,

c0 = 29 400 N/m, uc = 4u = umax /2 = 0.10/2 = 0.05 and


ur = uc and kr = 0 a simple linear is used in the rebound

using the same data we obtain:

kc = k = 0.1667.

By setting

mode, Fig. 5.12a.

a) Spring

b) Damper
5000

cS = 34300 N/m
6000 c0 = 29400 N/m
ur = 0.05 m
5000 kr = 0
uc = 0.05 m
4000 kc = 0.1667

FD [N/m]

FS [N/m]

7000

2500

d0 = 4200 N/(m/s)
pr = 0.4 1/(m/s)
pc = 1.2 1/(m/s)

3000
2000

-2500

1000
0
-0.1

rebound
u<0

-0.05

compression
u>0

0.05
u [m]

0.1

-5000
-1

rebound
v<0

-0.5

compression
v>0

0.5
v [m/s]

Figure 5.12: Spring and damper characteristics: - - - linear,  nonlinear


According to Section 5.3.4 damping coecients optimizing the ride comfort and the ride
safety can be calculated from Eqs. (5.65) and (5.69). For
equivalent linear spring rate,

(dS )C
opt =
(dS )Sopt

2 cS M =

M = 350 kg , m = 50 kg

and

cS = 34 300 N/m which is the


cT = 220 000 N/m we obtain

2 34 300 350 = 4900 N/(m/s) ,


(5.83)

p
(cS + cT ) m = (18 000 + 220 000) 50 = 3570 N/(m/s) .

1
The mean value d0 = 4200N/(m/s) may serve as compromise. With pr = 0.4(m/s)
and
pc = 1.2 (m/s)1 the nonlinearity becomes more intensive in compression than rebound,
Fig. 5.12b.

5.5.2 Results
The quarter car model is driven with constant velocity over a single obstacle. Here, a
cosine shaped bump with a height of

H = 0.08 m

and a length of

L = 2.0 m

was used.

The results are plotted in Fig. 5.13.


Compared to the linear model the nonlinear spring and damper characteristics result in
2
2
signicantly reduced peak values for the chassis acceleration (6.0m/s instead of 7.1m/s )

92

FH Regensburg, University of Applied Sciences

Chassis acceleration [m/s2]


10

Wheel load [N]


7000

7.1
6.0

Suspension travel [m]


0.04

6660
6160

6000

Prof. Dr.-Ing. G. Rill

0.02

5000
0

4000

-5

3000

-0.02

2000
-10
-15

0.5

1.5

-0.04

linear
nonlinear

1000
0

0.5

time [s]

1.5

-0.06

0.5

time [s]

Figure 5.13: Quarter car model driving with

and for the wheel load (6160

instead of

1.5

time [s]

v = 20 km h

6660 N ).

over a single obstacle

Even the tire lift o at

t 0.25 s

can be avoided. While crossing the bump large damper velocities occur. Here, the degressive damper characteristics provides less damping compared to the linear damper which
increases the suspension travel.

Chassis acceleration [m/s2]


10
5

Wheel load [N]

Suspension travel [m]

7000

0.04

6000

0.02

5000
0

4000

-5

3000

0
-0.02
-0.04

2000
-10
-15

linear,
low damping

1000
0

0.5

time [s]

1.5

-0.06

nonlinear

0.5

time [s]

1.5

-0.08

0.5

1.5

time [s]

Figure 5.14: Results for low damping compared to nonlinear model


A linear damper with a lower damping coecient,

d0 = 3000N/m for instance, also reduces

the peaks in the chassis acceleration and in the wheel load, but then the attenuation of
the disturbances will take more time. Fig. 5.14. Which surely is not optimal.

93

6 Longitudinal Dynamics
6.1 Dynamic Wheel Loads
6.1.1 Simple Vehicle Model
The vehicle is considered as one rigid body which moves along an ideally even and horizontal road. At each axle the forces in the wheel contact points are combined in one
normal and one longitudinal force.

S
h

Fz1

Fx1

mg

Fx2

a2

a1

Fz2

Figure 6.1: Simple vehicle model

If aerodynamic forces (drag, positive and negative lift) are neglected at rst, the equations
of motions in the

where

x-, z -plane

will read as

m v = Fx1 + Fx2 ,

(6.1)

0 = Fz1 + Fz2 m g ,

(6.2)

0 = Fz1 a1 Fz2 a2 + (Fx1 + Fx2 ) h ,

(6.3)

indicates the vehicle's acceleration,

wheel base, and

is the mass of the vehicle,

a1 + a2

is the

is the height of the center of gravity.

These are only three equations for the four unknown forces

Fx1 , Fx2 , Fz1 , Fz2 .

But, if we

insert Eq. (6.1) in Eq. (6.3), we can eliminate two unknowns at a stroke

0 = Fz1 a1 Fz2 a2 + m v h .

94

(6.4)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

The equations Eqs. (6.2) and (6.4) can be resolved for the axle loads now

Fz1 = m g

h
a2

m v ,
a1 + a2
a1 + a2

(6.5)

Fz2 = m g

h
a1
+
m v .
a1 + a2
a1 + a2

(6.6)

The static parts

st
Fz1
= mg

a2
,
a1 + a2

st
Fz2
= mg

a1
a1 + a2

(6.7)

describe the weight distribution according to the horizontal position of the center of
gravity. The height of the center of gravity only inuences the dynamic part of the axle
loads,

dyn
Fz1
= m g
When accelerating

v > 0,

h
v
,
a1 + a2 g

dyn
Fz2
= +m g

h
v
.
a1 + a2 g

(6.8)

the front axle is relieved as the rear axle is when decelerating

v < 0.

6.1.2 Influence of Grade

v
x
Fx1

mg
Fz1

a1
a2

h
Fx2

Fz2

Figure 6.2: Vehicle on grade

For a vehicle on a grade, Fig.6.2, the equations of motion Eq. (6.1) to Eq. (6.3) can easily
be extended to

m v = Fx1 + Fx2 m g sin ,


0 = Fz1 + Fz2 m g cos ,

(6.9)

0 = Fz1 a1 Fz2 a2 + (Fx1 + Fx2 ) h ,

95

Vehicle Dynamics

where

FH Regensburg, University of Applied Sciences

denotes the grade angle. Now, the axle loads are given by

Fz1 = m g cos

h
a2 h tan

m v ,
a1 + a2
a1 + a2

(6.10)

Fz2 = m g cos

a1 + h tan
h
+
m v ,
a1 + a2
a1 + a2

(6.11)

where the dynamic parts remain unchanged, whereas now the static parts also depend on
the grade angle and the height of the center of gravity.

6.1.3 Aerodynamic Forces


The shape of most vehicles or specic wings mounted at the vehicle produce aerodynamic
forces and torques. The eect of these aerodynamic forces and torques can be represented
by a resistant force applied at the center of gravity and down forces acting at the front
and rear axle, Fig. 6.3.

FD1

FD2
FAR
h
mg

Fx1

a2

a1

Fz1

Fx2
Fz2

Figure 6.3: Vehicle with aerodynamic forces


If we assume a positive driving speed,

v > 0,

the equations of motion will read as

m v = Fx1 + Fx2 FAR ,


0 = Fz1 FD1 + Fz2 FD2 m g ,

(6.12)

0 = (Fz1 FD1 ) a1 (Fz2 FD2 ) a2 + (Fx1 + Fx2 ) h ,


where

FAR

and

FD1 , FD2

describe the air resistance and the down forces. For the dynamic

axle loads we get

The down forces

Fz1 = FD1 + m g

a2
h

(m v + FAR ) ,
a1 + a2
a1 + a2

(6.13)

Fz2 = FD2 + m g

a1
h
+
(m v + FAR ) .
a1 + a2
a1 + a2

(6.14)

FD1 , FD2

increase the static axle loads, and the air resistance

generates an additional dynamic term.

96

FAR

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

6.2 Maximum Acceleration


6.2.1 Tilting Limits
Ordinary automotive vehicles can only apply pressure forces to the road. If we take the
demands

Fz1 0

and

Fz2 0

into account, Eqs. (6.10) and (6.11) will result in

v
a2

cos sin
g
h

and

v
a1
cos sin .
g
h

(6.15)

These two conditions can be combined in one

a1
cos
h

v
+ sin
g

Hence, the maximum achievable accelerations (v

limited by the grade angle

and the

> 0)
position a1 , a2 , h of

a2
cos .
h

(6.16)

and decelerations (v

the center of gravity.

< 0) are
For v
0

the tilting condition Eq. (6.16) results in

a2
a1
tan
h
h

(6.17)

which describes the climbing and downhill capacity of a vehicle.


The presence of aerodynamic forces complicates the tilting condition. Aerodynamic forces
become important only at high speeds. Here, the vehicle acceleration is normally limited
by the engine power.

6.2.2 Friction Limits


The maximum acceleration is also restricted by the friction conditions

|Fx1 | Fz1
where the same friction coecient

and

|Fx2 | Fz2

(6.18)

has been assumed at front and rear axle. In the limit

case

Fx1 = Fz1

and

Fx2 = Fz2

(6.19)

the linear momentum in Eq. (6.9) can be written as

m v max = (Fz1 + Fz2 ) m g sin .

(6.20)

Using Eqs. (6.10) and (6.11) one obtains

 
v
=
g max
That means climbing (v

a friction coecient

> 0, > 0)
tan ||.

cos sin .

or downhill stopping (v

< 0, < 0)

(6.21)

requires at least

97

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

According to the vehicle dimensions and the friction values the maximal acceleration or
deceleration is restricted either by Eq. (6.16) or by Eq. (6.21).
If we take aerodynamic forces into account, the maximum acceleration and deceleration
on a horizontal road will be limited by

FD1 FD2
+
1 +
mg
mg

FAR

mg

v
g

FD1 FD2
1 +
+
mg
mg

FAR
.
mg

(6.22)

In particular the aerodynamic forces enhance the braking performance of the vehicle.

6.3 Driving and Braking


6.3.1 Single Axle Drive
With the rear axle driven in limit situations,

Fx1 = 0

and

Fx2 = Fz2

hold. Then, using

Eq. (6.6) the linear momentum Eq. (6.1) results in


m v R WD = m g

a1
h
v R WD
+
a1 + a2 a1 + a2 g


,

(6.23)

where the subscript R WD indicates the rear wheel drive. Hence, the maximum acceleration
for a rear wheel driven vehicle is given by

v R WD
=
g

By setting

Fx1 = Fz1

and

Fx2 = 0,

h
1
a1 + a2

a1
.
a1 + a2

(6.24)

the maximum acceleration for a front wheel driven

vehicle can be calculated in a similar way. One gets

v F WD
=
g

a2
,
h
a1 + a2
1+
a1 + a2

where the subscript F WD denotes front wheel drive. Depending on the parameter
a2 and h the accelerations may be limited by the tilting condition vg ah2 .

(6.25)

, a1 ,

The maximum accelerations of a single axle driven vehicle are plotted in Fig. 6.4. For rear

a2 /(a1+a2 ) which describes the static axle load


distribution is in the range of 0.4 a2 /(a1+a2 ) 0.5. For = 1 and h = 0.55 this results
in maximum accelerations in between 0.77 v/g
0.64. Front wheel driven passenger
cars usually cover the range 0.55 a2 /(a1 +a2 ) 0.60 which produces accelerations in
the range of 0.45 v/g
0.49. Hence, rear wheel driven vehicles can accelerate much
wheel driven passenger cars, the parameter

faster than front wheel driven vehicles.

98

FH Regensburg, University of Applied Sciences

range of load distribution

v/g

0.8

FWD

RWD

Prof. Dr.-Ing. G. Rill

FWD

0.6
0.4

RWD

0.2
0

0.2

0.4

0.6

Figure 6.4: Single axle driven passenger car:

0.8

a2 / (a1+a2)

= 1, h = 0.55 m, a1 +a2 = 2.5 m

6.3.2 Braking at Single Axle


If only the front axle is braked, in the limit case

Fx1 = Fz1

and

Fx2 = 0

will hold. With

Eq. (6.5) one gets from Eq. (6.1)


m v F WB = m g

h
v F WB
a2

a1 + a2 a1 + a2 g


,

(6.26)

where the subscript F WB indicates front wheel braking. Then, the maximum deceleration
is given by

v F WB
=
g
If only the rear axle is braked (Fx1
deceleration

h
1
a1 + a2

a2
.
a1 + a2

= 0, Fx2 = Fz2 ),

v R WB
=
g

h
1+
a1 + a2

(6.27)

one will obtain the maximum

a1
,
a1 + a2

(6.28)

where the subscript R WB denotes a braked rear axle. Depending on the parameters
v
and h, the decelerations may be limited by the tilting condition
ah1 .
g

, a1 ,

a2 ,

The maximum decelerations of a single axle braked vehicle are plotted in Fig. 6.5. For
passenger cars the load distribution parameter
to

0.6.

a2 /(a1 +a2 ) usually covers the range of 0.4


v/g
= 0.51 to v/g
= 0.77

If only the front axle is braked, decelerations from

will be achieved. This is a quite large value compared to the deceleration range of a braked
rear axle which is in the range of

v/g
= 0.49

to

v/g
= 0.33.

Therefore, the braking

system at the front axle has a redundant design.

99

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

range of
load
distribution

v/g

-0.2

FWB

-0.4
-0.6
-0.8
-1

RWB
0

0.2

0.4

0.6

Figure 6.5: Single axle braked passenger car:

0.8

a2 / (a1+a2)

= 1, h = 0.55 m, a1 +a2 = 2.5 m

6.3.3 Optimal Distribution of Drive and Brake Forces


The sum of the longitudinal forces accelerates or decelerates the vehicle. In dimensionless
style Eq. (6.1) reads as

Fx1
Fx2
v
=
+
.
g
mg
mg

(6.29)

A certain acceleration or deceleration can only be achieved by dierent combinations of


the longitudinal forces

Fx1

and

Fx2 .

According to Eq. (6.19) the longitudinal forces are

limited by wheel load and friction.


The optimal combination of

Fx1

and

Fx2

will be achieved, when front and rear axle have

the same skid resistance.

Fx1 = Fz1

and

Fx2 = Fz2 .

(6.30)

With Eq. (6.5) and Eq. (6.6) one obtains

and

Fx1
=
mg

Fx2
=
mg

a2
v

h
g

a1
v
+
h
g

h
a1 + a2

(6.31)

h
.
a1 + a2

(6.32)

With Eq. (6.31) and Eq. (6.32) one gets from Eq. (6.29)

v
= ,
g
where it has been assumed that

Fx1

and

Fx2

(6.33)

have the same sign. Finally, if Eq. (6.33 is

inserted in Eqs. (6.31) and (6.32) one will obtain

Fx1
v
=
mg
g

100

a2
v

h
g

h
a1 + a2

(6.34)

FH Regensburg, University of Applied Sciences

and

Fx2
v
=
mg
g

Depending on the desired acceleration

a1
v
+
h
g

Prof. Dr.-Ing. G. Rill

h
.
a1 + a2

(6.35)

v > 0 or deceleration v < 0, the longitudinal forces

B2/mg

that grant the same skid resistance at both axles can be calculated now.

braking

Fx1/mg

-a1/h

dFx2
dFx1 0

-1

-2

B1/mg

a =1.15
1

driving

a =1.35
2

h=0.55
=1.20

tilting limits

Fx2/mg

a2/h

Figure 6.6: Optimal distribution of driving and braking forces


Fig. 6.6 shows the curve of optimal drive and brake forces for typical passenger car values.
At the tilting limits

v/g
= a1 /h and v/g
= +a2 /h, no longitudinal forces can be applied

at the lifting axle. The initial gradient only depends on the steady state distribution of
the wheel loads. From Eqs. (6.34) and (6.35) it follows

Fx1


a2
v
h
mg
=
2
v
h
g a1 + a2
d
g

(6.36)

Fx2


a1
v
h
mg
=
+2
.
v
h
g a1 + a2
d
g

(6.37)

and

101

Vehicle Dynamics

For

v/g
=0

FH Regensburg, University of Applied Sciences

the initial gradient remains as


d Fx2
a1
.
=

d Fx1 0
a2

(6.38)

6.3.4 Different Distributions of Brake Forces


Practical applications aim at approximating the optimal distribution of brake forces by
constant distribution, limitation, or reduction of brake forces as good as possible. Fig. 6.7.

limitation

Fx2/mg

Fx2/mg

constant
distribution

Fx1/mg
Fx2/mg

Fx1/mg

Fx1/mg

reduction

Figure 6.7: Dierent distributions of brake forces


When braking, the stability of a vehicle depends on the potential of generating a lateral
force at the rear axle. Thus, a greater skid (locking) resistance is realized at the rear axle
than at the front axle. Therefore, the brake force distribution are all below the optimal
curve in the physically relevant area. This restricts the achievable deceleration, specially
at low friction values.
Because the optimal curve depends on the center of gravity of the vehicle an additional
safety margin have to be installed when designing real brake force distributions. The
distribution of brake forces is often tted to the axle loads. There, the inuence of the
height of the center of gravity, which may also vary much on trucks, is not taken into
account and has to be compensated by a safety margin from the optimal curve. Only the
control of brake pressure in anti-lock-systems provides an optimal distribution of brake
forces independently from loading conditions.

6.3.5 Anti-Lock-Systems
On hard braking maneuvers large longitudinal slip values occur. Then, the stability and/or
steerability is no longer given because nearly no lateral forces can be generated. By controlling the brake torque or brake pressure respectively, the longitudinal slip can be restricted
to values that allow considerable lateral forces.
Here, the angular wheel acceleration

is used as a control variable. Angular accelerations

of the wheel are derived from the measured angular speeds of the wheel by dierentiation.
The rolling condition is fullled with a longitudinal slip of

rD = x

102

sL = 0.

Then
(6.39)

FH Regensburg, University of Applied Sciences

holds, where

rD

labels the dynamic tire radius and

Prof. Dr.-Ing. G. Rill

names the longitudinal acceleration

of the vehicle. According to Eq. (6.21), the maximum acceleration/deceleration of a vehicle


depends on the friction coecient,

|
x| = g .

For a given friction coecient

a simple

control law can be realized for each wheel


||

1
|
x| .
rD

(6.40)

Because no reliable possibility to determine the local friction coecient between tire and
road has been found until today, useful information can only be gained from Eq. (6.40)
at optimal conditions on dry road. Therefore, the longitudinal slip is used as a second
control variable.
In order to calculate longitudinal slips, a reference speed is estimated from all measured
wheel speeds which is used for the calculation of slip at all wheels, then. This method
is too imprecise at low speeds. Therefore, no control is applied below a limit velocity.
Problems also arise when all wheels lock simultaneously for example which may happen
on icy roads.
The control of the brake torque is done via the brake pressure which can be

held,

or

decreased

increased,

by a three-way valve. To prevent vibrations, the decrement is usually

made slower than the increment.


To prevent a strong yaw reaction, the

select low principle is often used with -split braking

at the rear axle. Here, the break pressure at both wheels is controlled by the wheel running
on lower friction. Thus, at least the brake forces at the rear axle cause no yaw torque.
However, the maximum achievable deceleration is reduced by this.

6.4 Drive and Brake Pitch


6.4.1 Vehicle Model
The vehicle model in Fig. 6.8 consists of ve rigid bodies. The body has three degrees
of freedom: Longitudinal motion

z1

and

z2

xA ,

vertical motion

zA

and pitch

A .

The coordinates

describe the vertical motions of wheel and axle bodies relative to the body.

The longitudinal and rotational motions of the wheel bodies relative to the body can be
described via suspension kinematics as functions of the vertical wheel motion:

x1 = x1 (z1 ) , 1 = 1 (z1 ) ;
x2 = x2 (z2 ) , 2 = 2 (z2 ) .
The rotation angles

R1

and

R2

describe the wheel rotations relative to the wheel bodies.

The forces between wheel body and vehicle body are labeled
drive torques

(6.41)

FF 1

and

FF 2 .

At the wheels

MA1 , MA2 and brake torques MB1 , MB2 , longitudinal forces Fx1 , Fx2 and
Fz1 , Fz2 apply. The brake torques are directly supported by the wheel

the wheel loads

bodies, whereas the drive torques are transmitted by the drive shafts to the vehicle body.

103

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

zA
xA

FF1

MA1

z1
MB1

MA1

hR

R1

MB1
Fz1

z2
MB2

MA2

Fx1
R

a1

FF2

MA2

R2
a2

MB2
Fz2

Fx2

Figure 6.8: Simple vehicle model

The forces and torques that apply to the single bodies are listed in the last column of the
tables 6.1 and 6.2.
The velocity of the vehicle body and its angular velocity are given by

v0A,0

x A
0
= 0 + 0 ;
0
zA

0
= A .
0

0A,0

(6.42)

At small rotational motions of the body one gets for the velocities of the wheel bodies
and wheels

x1
z1

v0RK1 ,0 = v0R1 ,0

x A
0
hR A
+
0
= 0 + 0 +

0
zA
a1 A

x2
z2

v0RK2 ,0 = v0R2 ,0

x A
0
hR A
+
0
= 0 + 0 +
0
zA
+a2 A

z1

0
z1
z2

0
z2

(6.43)

(6.44)

The angular velocities of the wheel bodies and wheels are obtained from

0RK1 ,0

0
0
= A + 1
0
0

and

0R1 ,0

0
0
0
= A + 1 + R1
0
0
0
(6.45)

104

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

as well as

0
0
= A + 2
0
0

0RK2 ,0

0R2 ,0

and

0
0
0
= A + 2 + R2
0
0
0
(6.46)

Introducing a vector of generalized velocities

z =

x A zA A 1 R1 2 R2

T

(6.47)

the velocities and angular velocities given by Eqs. (6.42), (6.43), (6.44), (6.45), and (6.46)
can be written as

v0i =

7
X
v0i
j=1

zj

zj

and

0i =

7
X
0i
j=1

zj

zj

(6.48)

6.4.2 Equations of Motion


The partial velocities

0i
v0i
and partial angular velocities
for the ve bodies
zj
zj

and for the seven generalized speeds

j = 1(1)7

partial velocities
bodies
chassis

mA

wheel body
front

mRK1

wheel
front

mR1

wheel body
rear

mRK2

wheel
rear

mR2

x A
1
0
0
1
0
0
1
0
0
1
0
0
1
0
0

zA
0
0
1
0
0
1
0
0
1
0
0
1
0
0
1

A
0
0
0
hR
0
a1
hR
0
a1
hR
0
a2
hR
0
a2

z1
0
0
0
x1
z1

0
1
x1
z1

0
1
0
0
0
0
0
0

i = 1(1)5

are arranged in the tables 6.1 and 6.2.

v0i /zj
R1
z2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
x2
0
z2
0
0
0
1
x2
0
z2
0
0
0
1

applied forces

R2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

Fie
0
0
FF 1 +FF 2 mA g
0
0
FF 1 mRK1 g
Fx1
0
Fz1 mR1 g
0
0
FF 2 mRK2 g
Fx2
0
Fz2 mR2 g

Table 6.1: Partial velocities and applied forces


With the aid of the partial velocities and partial angular velocities the elements of the
mass matrix

and the components of the vector of generalized forces and torques

Q can

be calculated.

M (i, j) =

T
5 
X
v0k
k=1

zi

T
5 
X
v0k
0k
0k
mk
+
k
;
zj
zj
zi
k=1

i, j = 1(1)7 ;

(6.49)

105

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

partial angular velocities


bodies
chassis

wheel body
front

RK1

wheel
front

R1

wheel body
rear

RK2

wheel
rear

R2

x A
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

zA
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

A
0
1
0
0
1
0
0
1
0
0
1
0
0
1
0

z1
0
0
0
0

R1
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0

1
z1

0
0
1
z1

0
0
0
0
0
0
0

0i /zj
applied torques
z2 R2
Mie
0
0
0
0
0 MA1MA2 a1 FF 1 +a2 FF 2
0
0
0
0
0
0
0
0
MB1
0
0
0
0
0
0
0
0
MA1 MB1 R Fx1
0
0
0
0
0
0
2
0
M
B2
z2
0
0
0
0
0
0
2
1
M
M
A2
B2 R Fx2
z2
0
0
0

Table 6.2: Partial angular velocities and applied torques

Q(i) =

T
5 
X
v0k
k=1

zi

Fke

T
5 
X
0k
k=1

zi

Mke ;

i = 1(1)7 .

(6.50)

Then, the equations of motion for the plane vehicle model are given by

M z = Q .

(6.51)

6.4.3 Equilibrium
With the abbreviations

m1 = mRK1 + mR1 ;

m2 = mRK2 + mR2 ;

mG = mA + m1 + m2

(6.52)

and

h = hR + R

(6.53)

The components of the vector of generalized forces and torques read as

Q(1) = Fx1 + Fx2 ;


Q(2) = Fz1 + Fz2 mG g ;

(6.54)

Q(3) = a1 Fz1 + a2 Fz2 h(Fx1 + Fx2 ) + a1 m1 g a2 m2 g ;


Q(4) = Fz1 FF 1 +

x1
z1

Fx1 m1 g +

Q(5) = MA1 MB1 R Fx1 ;

106

1
(MA1
z1

R Fx1 ) ;

(6.55)

FH Regensburg, University of Applied Sciences

Q(6) = Fz2 FF 2 +

x2
z2

Fx2 m2 g +

2
(MA2
z2

Prof. Dr.-Ing. G. Rill

R Fx2 ) ;

(6.56)

Q(7) = MA2 MB2 R Fx2 .


Without drive and brake forces

MA1 = 0 ;

MA2 = 0 ;

MB1 = 0 ;

MB2 = 0

(6.57)

from Eqs. (6.54), (6.55) and (6.56) one gets the steady state longitudinal forces, the spring
preloads, and the wheel loads

0
Fx1
= 0;

FF0 1 =
0
Fz1

0
Fx2
= 0;

b
a+b

= m1 g +

a
a+b

FF0 2 =

mA g ;
a2
a1 +a2

0
Fz2

mA g ;

mA g ;

= m2 g +

(6.58)

a1
a1 +a2

mA g .

6.4.4 Driving and Braking


Assuming that on accelerating or decelerating the vehicle the wheels neither slip nor lock,

R R1 = x A hR A +
R R2 = x A hR A +

x1
z1
x2
z2

z1 ;

(6.59)

z2

hold. In steady state the pitch motion of the body and the vertical motion of the wheels
reach constant values

A = Ast = const. ,

z1 = z1st = const. ,

z2 = z2st = const.

(6.60)

and Eq. (6.59) simplies to

R R1 = x A ;

R R2 = x A .

(6.61)

With Eqs. (6.60), (6.61) and (6.53) the equation of motion (6.51) results in

a
a
mG xA = Fx1
+ Fx2
;
a
a
0 = Fz1
+ Fz2
;

hR (m1 +m2 ) xA + R1

x
A
R

+ R2

x
A
R

a
a
a
a
= a Fz1
+ b Fz2
(hR + R)(Fx1
+ Fx2
);
(6.62)

x1
z1

x2
z2

m1 xA +

m2 xA +

1
z1

2
z2

R1 xRA

a
Fz1

FFa 1

x1
z1

a
Fx1

R1

x
A
R

a
= MA1 MB1 R Fx1
;

R2

x
A
R

a
= Fz2
FFa 2 +

R2

x
A
R

a
= MA2 MB2 R Fx2
;

x2
z2

a
Fx2
+

1
(MA1
z1

2
(MA2
z2

a
R Fx2
);

a
R Fx1
)

(6.63)

(6.64)

107

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

where the steady state spring forces, longitudinal forces, and wheel loads have been separated into initial and acceleration-dependent terms

a
0
st
;
+ Fxi
= Fxi
Fxi

Fzist = Fzi0 + Fzia ;

FFsti = FF0 i + FFa i ;

i = 1, 2 .

(6.65)

a
With given torques of drive and brake the vehicle acceleration x
A , the wheel forces Fx1
,
a
a
a
a
a
Fx2 , Fz1 , Fz2 and the spring forces FF 1 , FF 2 can be calculated from Eq. (6.62), Eq. (6.63)
and Eq. (6.64)
Via the spring characteristics which have been assumed as linear the accelerationdependent forces also cause a vertical displacement and pitch motion of the body besides
the vertical motions of the wheels,

FFa 1
FFa 2
a
Fz1
a
Fz2
Especially the pitch of the vehicle

= cA1 z1a ,
= cA2 z2a ,
= cR1 (zAa a Aa + z1a ) ,
= cR2 (zAa + b Aa + z2a ) .

(6.66)

Aa 6= 0, caused by drive or brake will be felt as annoying,

if too distinct.
By an axle kinematics with 'anti dive' and/or 'anti squat' properties,

the drive and/or

brake pitch angle can be reduced by rotating the wheel body and moving the wheel center
in longitudinal direction during the suspension travel.

6.4.5 Brake Pitch Pole


For real suspension systems the brake pitch pole can be calculated from the motions of
the wheel contact points in the

x-, z -plane,

Fig. 6.9.

pitch pole

x-, z- motion of the contact points


during compression and rebound
Figure 6.9: Brake pitch pole
Increasing the pitch pole height above the track level means a decrease in the brake pitch
angle. However, the pitch pole is not set above the height of the center of gravity in
practice, because the front of the vehicle would rise at braking then.

108

7 Lateral Dynamics
7.1 Kinematic Approach
7.1.1 Kinematic Tire Model
When a vehicle drives through a curve at low lateral acceleration, small lateral forces
will be needed for course holding. Then, hardly lateral slip occurs at the wheels. In the
ideal case at vanishing lateral slip the wheels only move in circumferential direction. The
velocity component of the contact point in the lateral direction of the tire vanishes then

vy = eTy v0P = 0 .

(7.1)

This constraint equation can be used as 'kinematic tire model' for course calculation of
vehicles moving in the low lateral acceleration range.

7.1.2 Ackermann Geometry


Within the validity limits of the kinematic tire model the necessary steering angle of the
front wheels can be constructed via the given momentary pivot pole

M,

Fig. 7.1.

At slowly moving vehicles the lay out of the steering linkage is usually done according to
the Ackermann geometry. Then, the following relations apply

a
a
and
tan 2 =
,
(7.2)
R
R+s
where s labels the track width and a denotes the wheel base. Eliminating the curve radius
R, we get
a
a tan 1
tan 2 =
or
tan 2 =
.
(7.3)
a
a + s tan 1
+s
tan 1
a
A
a
The deviations 42 = 2 2 of the actual steering angle 2 from the Ackermann steering
A
angle 2 , which follows from Eq. (7.3), are used, especially on commercial vehicles, to
tan 1 =

judge the quality of a steering system.


At a rotation around the momentary pole
every point of the vehicle. The angle

M,

the direction of the velocity is xed for

between the velocity vector

and the longitudinal

axis of the vehicle is called side slip angle. The side slip angle at point

tan P =
where

denes the distance of

x
R
P to

or

tan P =

x
tan 1 ,
a

is given by
(7.4)

the inner rear wheel.

109

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

v
P
P

a
x

Figure 7.1: Ackermann steering geometry at a two-axled vehicle

7.1.3 Space Requirement


The Ackermann approach can also be used to calculate the space requirement of a vehicle
during cornering, Fig. 7.2. If the front wheels of a two-axled vehicle are steered according
to the Ackermann geometry, the outer point of the vehicle front will run on the maximum
radius

Rmax ,

whereas a point on the inner side of the vehicle at the location of the rear

axle will run on the minimum radius

Rmin .

Hence, it holds

2
Rmax
= (Rmin + b)2 + (a + f )2 ,
where

a, b are the wheel base and the width of the vehicle, and f

(7.5)
species the distance from

the front of the vehicle to the front axle. Then, the space requirement
can be specied as a function of the cornering radius

4R = Rmax Rmin =
The space requirement

4R

Rmin

110

2.5

for a given vehicle dimension

(Rmin + b)2 + (a + f )2 Rmin .

(7.6)

of a typical passenger car and a bus is plotted in Fig. 7.3

Rmin = 5.0 m, a bus requires


needs only 1.5 times the width.

versus the minimum cornering radius. In narrow curves


space of

4R = Rmax Rmin

times the width, whereas a passenger car

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

ma

Rmin

Figure 7.2: Space requirement

bus: a=6.25 m, b=2.50 m, f=2.25 m


car: a=2.50 m, b=1.60 m, f=1.00 m

R [m]

5
4
3
2
1
0

10

20
30
R min [m]

40

50

Figure 7.3: Space requirement of a typical passenger car and bus

111

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

7.1.4 Vehicle Model with Trailer


7.1.4.1 Kinematics
Fig. 7.4 shows a simple lateral dynamics model for a two-axled vehicle with a single-axled
trailer. Vehicle and trailer move on a horizontal track. The position and the orientation

x0 , y0 , z0

= y
R

of the vehicle relative to the track xed frame


to the rear axle center

r02,0
and the rotation matrix

(7.7)

A02
Here, the tire radius

is dened by the position vector

cos sin 0
cos 0 .
= sin
0
0
1

is considered to be constant, and

(7.8)

x, y

as well as the yaw angle

are generalized coordinates.

a
b

A1

y2

x2

A2

y0

y3

A3

x0

Figure 7.4: Kinematic model with trailer


The position vector

r01,0 = r02,0 + A02 r21,2

112

with

r21,2

a
= 0
0

(7.9)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

and the rotation matrix

A01 = A02 A21

with

cos sin 0
cos 0
= sin
0
0
1

A21

describe the position and the orientation of the front axle, where
wheel base and

a = const

(7.10)

labels the

the steering angle.

The position vector

r03,0 = r02,0 + A02 r2K,2 + A23 rK3,3


with

r2K,2

b
= 0
0

(7.11)

c
= 0
0

(7.12)

cos sin 0
cos 0
= sin
0
0
1

(7.13)

rK3,2

and

and the rotation matrix

A03 = A02 A23

with

A23

dene the position and the orientation of the trailer axis, with
between vehicle and trailer, and
coupling point

b, c

labeling the bend angle

marking the distances from the rear axle

and from the coupling point

to the trailer axis

to the

3.

7.1.4.2 Vehicle Motion


According to the kinematic tire model, cf. section 7.1.1, the velocity at the rear axle can
only have a component in the longitudinal direction of the tire which here corresponds
with the longitudinal direction of the vehicle

v02,2

vx2
= 0 .
0

(7.14)

The time derivative of Eq. (7.7) results in

v02,0 = r02,0

x
= y .
0

The transformation of Eq. (7.14) into the system

v02,0 = A02 v02,2 = A02

(7.15)

vx2
cos vx2
0 = sin vx2
0
0

(7.16)

113

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

compared to Eq. (7.15) results in two rst order dierential equations for the position
coordinates

and

y
x = vx2 cos ,

y = vx2 sin .

(7.17)

The velocity at the front axle follows from Eq. (7.9)

v01,0 = r01,0 = r02,0 + 02,0 A02 r21,2 .


x2 , y2 , z2 we obtain

vx2
0
a
vx2
= 0 + 0 0 = a .
0

0
0
| {z }
| {z } | {z }
v02,2
02,2
r21,2

(7.18)

Transformed into the vehicle xed system

v01,2

(7.19)

The unit vectors

ex1,2

cos
= sin
0

and

ey1,2

sin
= cos
0

(7.20)

dene the longitudinal and lateral direction at the front axle. According to Eq. (7.1) the
velocity component lateral to the wheel must vanish,

eTy1,2 v01,2 = sin vx2 + cos a = 0 .

(7.21)

Whereas in longitudinal direction the velocity

eTx1,2 v01,2 = cos vx2 + sin a = vx1

(7.22)

remains. From Eq. (7.21) a rst order dierential equation follows for the yaw angle

vx2
tan .
a

(7.23)

7.1.4.3 Entering a Curve


In analogy to Eq. (7.2) the steering angle
or with

k = 1/R

can be related to the current track radius

to the current track curvature

tan =

a
= ak .
R

(7.24)

Then, the dierential equation for the yaw angle reads as

= vx2 k .
With the curvature gradient

k = k(t) = kC

114

(7.25)

t
,
T

(7.26)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

the entering of a curve is described as a continuous transition from a straight line with

k=0

the curvature

into a circle with the curvature

k = kC .

The yaw angle of the vehicle can be calculated by simple integration now

vx2 kC t2
,
T
2
angle, (t = 0) = 0,

(t) =
where at time

t=0

a vanishing yaw

(7.27)
has been assumed. Then, the

position of the vehicle follows with Eq. (7.27) from the dierential equations Eq. (7.17)

Zt=T
x = vx2


cos

vx2 kC t2
T
2

Zt=T


dt ,

y = vx2

t=0


sin

vx2 kC t2
T
2


dt .

(7.28)

t=0

At constant vehicle speed,

vx2 = const., Eq. (7.28) is the parameterized form of a clothoide.

From Eq. (7.24) the necessary steering angle can be calculated, too. If only small steering
angles are necessary for driving through the curve, the

tan-function can be approximated

by its argument, and

= (t) a k = a kC

t
T

(7.29)

holds, i.e. the driving through a clothoide is manageable by a continuous steer motion.

7.1.4.4 Trailer Motions


The velocity of the trailer axis can be obtained by dierentiation of the position vector
Eq. (7.11)

v03,0 = r03,0 = r02,0 + 02,0 A02 r23,2 + A02 r23,2 .


The velocity

r02,0 = v02,0

and the angular velocity

02,0

(7.30)

of the vehicle are dened in

Eqs. (7.16) and (7.19). The position vector from the rear axle to the axle of the trailer is
given by

r23,2 = r2K,2 + A23 rK3,3


where

r2K,2

and

rK3,3

b c cos
c sin ,
=
0

(7.31)

are dened in Eq. (7.12). The time derivative of Eq. (7.31) results

in


c cos
c sin
0
= 0 c sin = c cos .
0
0

{z
}
| {z } |
23,2
A23 rK3,3

r23,2

(7.32)

x2 , y2 , z2 now

vx2
0
b c cos
c sin
vx2 + c sin (+
)

c sin + c cos = b c cos (+


)
.
= 0 + 0
0

0
0
0
{z
} |
{z
}
| {z } | {z } |
v02,2
02,2
r23,2
r23,2

Eq. (7.30) is transformed into the vehicle xed frame

v03,2

(7.33)

115

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The longitudinal and lateral direction at the trailer axle are dened by the unit vectors

ex3,2

cos
= sin
0

ey3,2

and

sin
= cos .
0

(7.34)

At the trailer axis the lateral velocity must also vanish



)
+ cos b c cos (+
)

eTy3,2 v03,2 = sin vx2 + c sin (+


= 0,

(7.35)

whereas in longitudinal direction the velocity



)

eTx3,2 v03,2 = cos vx2 + c sin (+


+ sin b c cos (+
= vx3

(7.36)

remains. If Eq. (7.23) is inserted into Eq. (7.35) now, one will get a rst order dierential
equation for the bend angle

vx2
=
a

a
sin +
c

b
cos + 1 tan
c


.

(7.37)

The dierential equations Eq. (7.17) and Eq. (7.23) describe position and orientation

x0 , y0

within the

plane. The position of the trailer relative to the vehicle follows from

Eq. (7.37).

7.1.4.5 Course Calculations

front axle
rear axle
trailer axle

20
[m]
10
0

-30

-20

-10

10

20

30 [m] 40

50

60

30

[o]

front axle steering angle

20
10
0

10

15

20

[s]

25

30

Figure 7.5: Entering a curve


For a given set of vehicle parameters
velocity,

vx2 = vx2 (t)

a, b , c ,

and predened time functions of the vehicle

and the steering angle,

= (t),

the course of vehicle and trailer

can be calculated by numerical integration of the dierential equations Eqs. (7.17), (7.23)
and (7.37). If the steering angle is slowly increased at constant driving speed, the vehicle
drives a gure which will be similar to a clothoide, Fig. 7.5.

116

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

7.2 Steady State Cornering


7.2.1 Cornering Resistance
In a body xed reference frame

B,

Fig. 7.6, the velocity state of the vehicle can be

described by

v0C,B
where

v cos
= v sin
0

and

0F,B

0
= 0 ,

denotes the side slip angle of the vehicle measured at the center of gravity. The

angular velocity of a vehicle cornering with constant velocity


is given by

=
where

(7.38)

on an at horizontal road

v
,
R

(7.39)

denotes the radius of curvature.

Fx2

a2

Fy2

a1

xB

Fx1

R
yB

Fy1

Figure 7.6: Cornering resistance

In the body xed reference frame, linear and angular momentum result in


v2
m sin
= Fx1 cos Fy1 sin + Fx2 ,
R
 2

v
m
cos
= Fx1 sin + Fy1 cos + Fy2 ,
R
0 = a1 (Fx1 sin + Fy1 cos ) a2 Fy2 ,


(7.40)

(7.41)
(7.42)

117

Vehicle Dynamics

where

FH Regensburg, University of Applied Sciences

denotes the mass of the vehicle,

Fx1 , Fx2 , Fy1 , Fy2

are the resulting forces in

longitudinal and vertical direction applied at the front and rear axle, and

species the

average steer angle at the front axle.


The engine torque is distributed by the center dierential to the front and rear axle. Then,
in steady state condition we obtain

Fx1 = k FD
where

FD

is the driving force and by

k=0

and

Fx2 = (1 k) FD ,

dierent driving conditions

rear wheel drive

0<k<1

all wheel drive

k=1

front wheel drive

(7.43)
can be modeled:

Fx1 = 0, Fx2 = FD
Fx1
k
=
Fx2
1k
Fx1 = FD , Fx2 = 0

If we insert Eq. (7.43) into Eq. (7.40) we will get


k cos + (1k) FD
k sin FD +

sin Fy1
cos Fy1 +

Fy2

mv 2
=
sin ,
R
mv 2
=
cos ,
R

(7.44)

a1 k sin FD + a1 cos Fy1 a2 Fy2 = 0 .


These equations can be resolved for the driving force

FD

a2
cos sin sin cos
mv 2
a1 + a2
=
.
k + (1 k) cos
R

(7.45)

The driving force will vanish, if

a2
cos sin = sin cos
a1 + a2

or

a2
tan = tan
a1 + a2

(7.46)

holds. This fully corresponds with the Ackermann geometry. But, the Ackermann geometry applies only for small lateral accelerations. In real driving situations, the side slip
angle of a vehicle at the center of gravity is always smaller than the Ackermann side slip
a2
angle. Then, due to tan <
tan a driving force FD > 0 is needed to overcome the
a1 +a2
'cornering resistance' of the vehicle.

7.2.2 Overturning Limit


The overturning hazard of a vehicle is primarily determined by the track width and the
height of the center of gravity. With trucks however, also the tire deection and the body
roll have to be respected., Fig. 7.7.

118

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

2 1
m ay
h2
mg

h1

F yL

F yR

FzL

s/2

s/2

FzR

Figure 7.7: Overturning hazard on trucks

The balance of torques at the height of the track plane applied at the already inclined
vehicle results in

s
= m ay (h1 + h2 ) + m g [(h1 + h2 )1 + h2 2 ] ,
2
where ay describes the lateral acceleration, m is the sprung mass, and small
the axle and the body were assumed, 1  1, 2  1.
(FzL FzR )

(7.47)
roll angles of

On a left-hand tilt, the right tire raises

T
FzR
= 0,

(7.48)

whereas the left tire carries the complete vehicle weight

T
FzL
= mg .

(7.49)

Using Eqs. (7.48) and (7.49) one gets from Eq. (7.47)

s
aTy
h2
2
=
1T
2T .
g
h1 + h2
h1 + h2

(7.50)

T
rises above the limit ay . Roll
T
T
of axle and body reduce the overturning limit. The angles 1 and 2 can be calculated
from the tire stiness cR and the roll stiness of the axle suspension.
The vehicle will turn over, when the lateral acceleration

ay

119

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

If the vehicle drives straight ahead, the weight of the vehicle will be equally distributed
to both sides

stat
stat
FzR
= FzL
=

1
mg .
2

(7.51)

With

T
stat
FzL
= FzL
+ 4Fz

(7.52)

and Eqs. (7.49), (7.51), one obtains for the increase of the wheel load at the overturning
limit

4Fz =

1
mg .
2

(7.53)

Then, the resulting tire deection follows from

4Fz = cR 4r ,
where

cR

(7.54)

is the radial tire stiness.

Because the right tire simultaneously rebounds with the same amount, for the roll angle
of the axle

2 4r = s 1T

or

1T =

mg
2 4r
=
s
s cR

(7.55)

holds. In analogy to Eq. (7.47) the balance of torques at the body applied at the roll
center of the body yields

cW 2 = m ay h2 + m g h2 (1 + 2 ) ,

(7.56)

cW names the roll stiness of the body suspension. In particular, at the overturning
ay = aTy
aTy
mgh2
mgh2
T
2 =
+
1T
(7.57)
g cW mgh2
cW mgh2

where
limit

T
applies. Not allowing the vehicle to overturn already at ay = 0 demands a minimum of
min
roll stiness cW > cW
= mgh2 . With Eqs. (7.55) and (7.57) the overturning condition
Eq. (7.50) reads as

(h1 + h2 )

aTy
aTy
s
1
1
1
1
h2
,
=
(h1 + h2 ) h2

g
2
cR
g cW 1
cW 1 cR

(7.58)

where, for abbreviation purposes, the dimensionless stinesses

cR
cR = m g
s

and

cW =

cW
m g h2

(7.59)

have been used. Resolved for the normalized lateral acceleration

aTy
=
g

120

s
2
h2
h1 + h2 +
cW 1

1
cR

(7.60)

FH Regensburg, University of Applied Sciences

roll angle =T1+2


T

overturning limit ay

0.6

Prof. Dr.-Ing. G. Rill

20

0.5
15
0.4
0.3

10

0.2
5
0.1
0

0
10
20
normalized roll stiffness c W
*

0
10
20
normalized roll stiffness c W
*

Figure 7.8: Tilting limit for a typical truck at steady state cornering

remains.

m = 13 000 kg . The radial stiness


width can be set to s = 2 m. The values

At heavy trucks, a twin tire axle may be loaded with

cR = 800 000 N/m, and the track


h1 = 0.8 m and h2 = 1.0 m hold at maximal load. These values produce the results shown

in Fig. 7.8. Even with a rigid body suspension cW , the vehicle turns over at a lateral
acceleration of ay 0.5 g . Then, the roll angle of the vehicle solely results from the tire
of one tire is

deection.
At a normalized roll stiness of

cW = 5,

the overturning limit lies at

ay 0.45 g

and

so reaches already 90% of the maximum. The vehicle will turn over at a roll angle of
= 1 + 2 10 then.

7.2.3 Roll Support and Camber Compensation


When a vehicle drives through a curve with the lateral acceleration

ay ,

centrifugal forces

will be applied to the single masses. At the simple roll model in Fig. 7.9, these are the
forces

m A ay

and

m R ay ,

where

Through the centrifugal force

mA

m A ay

names the body mass and

the wheel mass.

applied to the body at the center of gravity, a torque

is generated, which rolls the body with the angle


of the tires

mR

and leads to an opposite deection

z1 = z2 .

At steady state cornering, the vehicle forces are balanced. With the principle of virtual
work

W = 0 ,

(7.61)

the equilibrium position can be calculated.


At the simple vehicle model in Fig. 7.9 the suspension forces

Fy1 , Fz1 , Fy2 , Fz2 ,

FF 1 , FF 2

and tire forces

are approximated by linear spring elements with the constants

cA

and

121

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

b/2

b/2
zA

mA a y

A
SA

yA
FF1

FF2
h0

z2

z1

mR a y

mR a y

S2

r0

y2

Q2

S1
Q1

F y2

Fy2

F z1

y1
F y1

Figure 7.9: Simple vehicle roll model

cQ , cR .

The work

potential

V.

of these forces can be calculated directly or using

W = V

via the

At small deections with linearized kinematics one gets

W = mA ay yA
mR ay (yA + hR A + y1 )2 mR ay (yA + hR A + y2 )2
21 cA z12

1
2

cA z22

(7.62)

21 cS (z1 z2 )2
21 cQ (yA + h0 A + y1 + r0 1 )2 12 cQ (yA + h0 A + y2 + r0 2 )2
2
2
21 cR zA + 2b A + z1 21 cR zA 2b A + z2 ,
where the abbreviation

hR = h0 r0

has been used, and

cS

describes the spring constant

of the anti roll bar, converted to the vertical displacement of the wheel centers.
The kinematics of the wheel suspension are symmetrical. With the linear approaches

y1 =
the work

y
z1 ,
z

1 =

1
z

and

y2 =

y
z2 ,
z

2 =

2
z

(7.63)

can be described as a function of the position vector

y = [ yA , zA , A , z1 , z2 ]T .

(7.64)

W = W (y)

(7.65)

Due to

the principle of virtual work Eq. (7.61) leads to

W =

122

W
y = 0 .
y

(7.66)

FH Regensburg, University of Applied Sciences

Because of

y 6= 0,

Prof. Dr.-Ing. G. Rill

a system of linear equations in the form of

Ky = b
results from Eq. (7.66). The matrix

2 cQ

K = 2 cQ h0

yQ c
z Q

Q
yz cQ

Q
yz cQ

cR
Q
b
c +h0 yz cQ
2 R

cA + cS + cR

2b cR h0 yz cQ

and

are given by

y Q
c
z Q

Q
b
c +h0 yz cQ
2 R

cR

and the vector

2 cQ h0

2 cR

cR

(7.67)

b =

cS

mA + 2 mR
0
(m1 + m2 ) hR
mR y/z
mR y/z

cR

Q
y
b
2 cR h0 z cQ

cS

cA + cS + cR

(7.68)

ay .

(7.69)

The following abbreviations have been used:

y
y Q

=
+ r0
,
z
z
z

cA


= cA + cQ

y
z

2
,

c =

2 cQ h20

 2
b
+ 2 cR
.
2

(7.70)

The system of linear equations Eq. (7.67) can be solved numerically, e.g. with MATLAB.
Thus, the inuence of axle suspension and axle kinematics on the roll behavior of the
vehicle can be investigated.

a)

b)

roll center

1 0

roll center

2 0

Figure 7.10: Roll behavior at cornering: a) without and b) with camber compensation
If the wheels only move vertically to the body at jounce and rebound, at fast cornering
the wheels will be no longer perpendicular to the track Fig. 7.10 a. The camber angles

1 > 0

and

2 > 0

result in an unfavorable pressure distribution in the contact area,

which leads to a reduction of the maximally transmittable lateral forces. Thus, at more

123

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

sportive vehicles axle kinematics are employed, where the wheels are rotated around the

1 = 1 (z1 ) and 2 = 2 (z2 ). Hereby, a camber


1 0 and 2 0. Fig. 7.10 b. By the rotation of

longitudinal axis at jounce and rebound,


compensation can be achieved with

the wheels around the longitudinal axis on jounce and rebound, the wheel contact points
are moved outwards, i.e against the lateral force. By this, a 'roll support' is achieved that
reduces the body roll.

7.2.4 Roll Center and Roll Axis

roll axis
roll center front

roll center rear

Figure 7.11: Roll axis


The 'roll center' can be constructed from the lateral motion of the wheel contact points

Q1

and

Q2 ,

Fig. 7.10. The line through the roll center at the front and rear axle is called

'roll axis', Fig. 7.11.

7.2.5 Wheel Loads


-TT

+TT

PR0+P
PF0+P

PF0-P

PR0+PR

PR0-P
PF0+PF

PR0-PR

PF0-PF

Figure 7.12: Wheel loads for a exible and a rigid chassis


The roll angle of a vehicle during cornering depends on the roll stiness of the axle and
on the position of the roll center. Dierent axle layouts at the front and rear axle may
result in dierent roll angles of the front and rear part of the chassis, Fig. 7.12.

124

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

On most passenger cars the chassis is rather sti. Hence, front and rear part of the chassis
are forced by an internal torque to an overall chassis roll angle. This torque aects the
wheel loads and generates dierent wheel load dierences at the front and rear axle. Due
to the degressive inuence of the wheel load to longitudinal and lateral tire forces the
steering tendency of a vehicle can be aected.

7.3 Simple Handling Model


7.3.1 Modeling Concept
x0

a2
a1

y0

Fy2
x2

xB

Fy1

y1

y2

yB

x1

Figure 7.13: Simple handling model


The main vehicle motions take place in a horizontal plane dened by the earth-xed
frame

0,

Fig. 7.13. The tire forces at the wheels of one axle are combined to one resulting

force. Tire torques, rolling resistance, and aerodynamic forces and torques, applied at the
vehicle, are not taken into consideration.

7.3.2 Kinematics
The vehicle velocity at the center of gravity can be expressed easily in the body xed
frame

xB , yB , zB

vC,B
where

v cos
= v sin ,
0

denotes the side slip angle, and

(7.71)

is the magnitude of the velocity.

125

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The velocity vectors and the unit vectors in longitudinal and lateral direction of the axles
are needed for the computation of the lateral slips. One gets

ex1 ,B

cos
= sin ,
0

and

ey1 ,B

sin
= cos ,
0

ex2 ,B
where
and

a1

and

a2

1
= 0 ,
0

v01,B

ey2 ,B

0
= 1 ,
0

v cos
= v sin + a1
0

(7.72)

v02,B

v cos
= v sin a2 ,
0

(7.73)

are the distances from the center of gravity to the front and rear axle,

denotes the yaw angular velocity of the vehicle.

7.3.3 Tire Forces


Unlike with the kinematic tire model, now small lateral motions in the contact points
are permitted. At small lateral slips, the lateral force can be approximated by a linear
approach

Fy = cS sy ,
where

cS

is a constant depending on the wheel load

(7.74)

Fz , and the lateral slip sy

is dened by

Eq. (3.88). Because the vehicle is neither accelerated nor decelerated, the rolling condition
is fullled at each wheel

rD = eTx v0P .
Here,

rD

is the dynamic tire radius,

v0P

the contact point velocity, and

(7.75)

ex

the unit vector

in longitudinal direction. With the lateral tire velocity

vy = eTy v0P

(7.76)

and the rolling condition Eq. (7.75), the lateral slip can be calculated from

eTy v0P
,
sy = T
| ex v0P |
with

ey

(7.77)

labeling the unit vector in the lateral direction direction of the tire. So, the lateral

forces are given by

Fy1 = cS1 sy1 ; Fy2 = cS2 sy2 .

(7.78)

7.3.4 Lateral Slips


With Eq. (7.73), the lateral slip at the front axle follows from Eq. (7.77):

sy1 =

126

+ sin (v cos ) cos (v sin + a1 )

.
| cos (v cos ) + sin (v sin + a1 )
|

(7.79)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

The lateral slip at the rear axle is given by

sy2 =
The yaw velocity of the vehicle

v sin a2
.
| v cos |

the side slip angle

(7.80)

and the steering angle

are

considered to be small

| a1 |  |v| ;
| |  1

| a2 |  |v|

(7.81)

||  1 .

(7.82)

and

Because the side slip angle always labels the smaller angle between the velocity vector
and the vehicle longitudinal axis, instead of

v sin v

the approximation

v sin |v|

(7.83)

has to be used. Now, Eqs. (7.79) and (7.80) result in

sy1 =

a1
v
+

|v|
|v|

(7.84)

a2
,
|v|

(7.85)

and

sy2 = +

where the consequences of Eqs. (7.81), (7.82), and (7.83) were already taken into consideration.

7.3.5 Equations of Motion


The velocities, angular velocities, and the accelerations are needed to derive the equations

 1, Eq. (7.71)

v
vC,B = |v| .
0

of motion, For small side slip angles

can be approximated by

(7.86)

The angular velocity is given by

0F,B

0
= 0 .

If the vehicle accelerations are also expressed in the vehicle xed frame

(7.87)

xF , yF , zF ,

one

v = const and with neglecting small higher-order terms

0
= 0F,B vC,B + v C,B = v + |v| .
(7.88)
0

will nd at constant vehicle speed

aC,B

127

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

The angular acceleration is given by

0F,B

0
= 0 ,

(7.89)

where the substitution

(7.90)

was used. The linear momentum in the lateral direction of the vehicle reads as

= Fy1 + Fy2 ,
m (v + |v| )
where, due to the small steering angle, the term
and

Fy1 cos

(7.91)

has been approximated by

describes the vehicle mass. With Eq. (7.90) the angular momentum yields

= a1 Fy1 a2 Fy2 ,
where

Fy1 ,

(7.92)

names the inertia of vehicle around the vertical axis. With the linear description

of the lateral forces Eq. (7.78) and the lateral slips Eqs. (7.84), (7.85), one gets from
Eqs. (7.91) and (7.92) two coupled, but linear rst order dierential equations

cS1
m |v|

a1 cS1



a2
cS2
v

|v|
m |v|
|v|




a1
v
a2
a2 cS2

+

,
|v|
|v|
|v|

a1
v
+

|v|
|v|

(7.93)

(7.94)

which can be written in the form of a state equation

cS1 + cS2

m |v|

a2 cS2 a1 cS1
| {z }
x

|


v
a2 cS2 a1 cS1
v cS1

 
m |v||v|
|v|
|v| m |v|
+

| {z } v a1 cS1
a21 cS1 + a22 cS2

x
|v|
|v|
{z
}
|
{z
A
B

 

.
|{z}
u

(7.95)

If a system can be at least approximatively described by a linear state equation, stability,


steady state solutions, transient response, and optimal controlling can be calculated with
classic methods of system dynamics.

7.3.6 Stability
7.3.6.1 Eigenvalues
The homogeneous state equation

x = A x

128

(7.96)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

describes the eigen-dynamics. If the approach

xh (t) = x0 et

(7.97)

is inserted into Eq. (7.96), the homogeneous equation will remain

( E A) x0 = 0 .
One gets non-trivial solutions

x0 6= 0

(7.98)

for

det | E A| = 0 .
The eigenvalues

(7.99)

provide information concerning the stability of the system.

7.3.6.2 Low Speed Approximation


The state matrix

cS1 + cS2

m |v|
=

Av0

v
a2 cS2 a1 cS1

m |v||v|
|v|

2
2

a1 cS1 + a2 cS2

|v|

(7.100)

v 0.

The matrix in

a21 cS1 + a22 cS2


.
|v|

(7.101)

approximates the eigen-dynamics of vehicles at low speeds,


Eq. (7.100) has the eigenvalues

1v0 =

cS1 + cS2
m |v|

and

2v0 =

The eigenvalues are real and always negative independent from the driving direction.
Thus, vehicles possess an asymptotically stable driving behavior at low speed!

7.3.6.3 High Speed Approximation


At high driving velocities,

v ,

the state matrix can be approximated by

Av

=
a2 cS2 a1 cS1

|v|
.

(7.102)

Using Eq. (7.102) one receives from Eq. (7.99) the relation

2v +

v a2 cS2 a1 cS1
= 0
|v|

(7.103)

129

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

with the solutions

1,2v
When driving forward with

v > 0,

r
v a2 cS2 a1 cS1
.
=
|v|

(7.104)

the root argument will be positive, if

a2 cS2 a1 cS1 < 0

(7.105)

holds. Then however, one eigenvalue is positive, and the system is unstable. Two zeroeigenvalues

1 = 0

and

2 = 0

are obtained for

a1 cS1 = a2 cS2 .

(7.106)

The driving behavior is indierent then. Slight parameter variations, however, can lead
to an unstable behavior. With

a2 cS2 a1 cS1 > 0


and

v > 0

or

a1 cS1 < a2 cS2

(7.107)

the root argument in Eq. (7.104) becomes negative. Then, the eigenvalues

are imaginary, and disturbances lead to undamped vibrations. To avoid instability, highspeed vehicles have to satisfy the condition Eq. (7.107). The root argument in Eq. (7.104)
changes at backward driving its sign. Hence, a vehicle showing stable driving behavior at
forward driving becomes unstable at fast backward driving!

7.3.7 Steady State Solution


7.3.7.1 Side Slip Angle and Yaw Velocity

= 0 , a stable system reaches steady state after a certain time.


x st = 0, the state equation Eq. (7.95) is reduced to a system of linear

At a given steering angle


With

xst = const.

or

equations

A xst = B u .
With the elements from the state matrix

and the vector

two equations to determine the steady state side slip angle


velocity

st

(7.108)

at a constant given steering angle

B,

one gets from Eq. (7.108)

st and the steady state angular

= 0

|v| (cS1 + cS2 ) st + (m v |v| + a1 cS1 a2 cS2 ) st = v cS1 0 ,

(7.109)

|v| (a1 cS1 a2 cS2 ) st + (a21 cS1 + a22 cS2 ) st = v a1 cS1 0 ,

(7.110)

where the rst equation has been multiplied by

m |v| |v|

and the second with

|v|.

The solution can be derived from

st =

130

v cS1 0

m v |v| + a1 cS1 a2 cS2

v a1 cS1 0

a21 cS1 + a22 cS2

|v| (cS1 + cS2 )

m v |v| + a1 cS1 a2 cS2

|v| (a1 cS1 a2 cS2 )

a21 cS1 + a22 cS2

(7.111)

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

and

|v| (cS1 + cS2 )

v cS1 0

|v| (a1 cS1 a2 cS2 ) v a1 cS1 0

st =

(7.112)

|v| (cS1 + cS2 )

m v |v| + a1 cS1 a2 cS2

|v| (a1 cS1 a2 cS2 )

a21 cS1 + a22 cS2

The denominator results in

detD = |v| cS1 cS2 (a1 + a2 )2 + m v |v| (a2 cS2 a1 cS1 )


For a non vanishing denominator

detD 6= 0,

(7.113)

steady state solutions exist

a1
v
cS2 (a1 + a2 )
=
0 ,
|v| a + a + m v |v| a2 cS2 a1 cS1
1
2
cS1 cS2 (a1 + a2 )
a2 m v |v|

st

st =

a1 + a2 + m v |v|
At forward driving vehicles

a2 cS2 a1 cS1 0 .
cS1 cS2 (a1 + a2 )

(7.114)

(7.115)

v > 0, the steady state side slip angle starts with the kinematic

value

v0
st
=

v
a2
0
|v| a1 + a2

v0
st
=

and

v
0
a1 + a2

(7.116)

and decreases with increasing speed. At speeds larger than

s
vst=0 =

a2 cS2 (a1 + a2 )
a1 m

(7.117)

the side slip angle changes the sign. Using the kinematic value of the yaw velocity
Eq. (7.115) can be written as

st =

v
a1 + a2

1
1 +

where

s
vch =

|v|
v

0 ,


(7.118)

v 2
vch


cS1 cS2 (a1 + a2 )2


m (a2 cS2 a1 cS1 )

(7.119)

is called the 'characteristic' speed of the vehicle.

, and the driven curve radius R are plotted versus the


v . The steering angle has been set to 0 = 1.4321 , in order to let the vehicle

In Fig. 7.14 the side slip angle


driving speed

131

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

steady state side slip angle

radius of curvrature

200

0
150
r [m]

[deg]

-2
-4
-6

a1*c S1/a2*c S2 = 0.66667


a1*cS1/a2*c S2 = 1
a1*cS1/a2*c S2 = 1.3333

-8
-10

100

10

20
v [m/s]

m=700 kg;
=1000 kg m2 ;

30

a1 =1.2 m;
a2 =1.3 m;

a1*cS1/a2*cS2 = 0.66667
a1*cS1/a2*cS2 = 1
a1*cS1/a2*cS2 = 1.3333

50
0

40

10

cS1 = 80 000 N m;

20
v [m/s]

cS2

30

40

110 770 N m
= 73 846 N m
55 385 N m

Figure 7.14: Steady state cornering

drive a circle with the radius

R0 = 100 m

at

has been calculated via

st =

v 0.

The actually driven circle radius

v
.
R

(7.120)

Some concepts for an additional steering of the rear axle were trying to keep the side
slip angle of the vehicle, measured at the center of the vehicle to zero by an appropriate
steering or controlling. Due to numerous problems, production stage could not yet be
reached.

7.3.7.2 Steering Tendency


After reaching the steady state solution, the vehicle moves on a circle. When inserting
Eq. (7.120) into Eq. (7.115) and resolving for the steering angle, one gets

0 =

a1 + a2
v 2 v a2 cS2 a1 cS1
+ m
.
R
R |v| cS1 cS2 (a1 + a2 )

(7.121)

The rst term is the Ackermann steering angle, which follows from Eq. (7.2) with the
wheel base

a = a1 + a2

and the approximation for small steering angles

tan 0 0 .

The Ackermann-steering angle provides a good approximation for slowly moving vehicles,

v 0. Depending on
v > 0, backward: v < 0),

because the second expression in Eq. (7.121) becomes very small at


the value of

a2 cS2 a1 cS1

and the driving direction (forward:

the necessary steering angle diers from the Ackermann-steering angle at higher speeds.
The dierence is proportional to the lateral acceleration

ay =

132

v2
.
R

(7.122)

FH Regensburg, University of Applied Sciences

At

v > 0 the
a1 , a2

Prof. Dr.-Ing. G. Rill

steering tendency of a vehicle is dened by the position of the center of

gravity

and the cornering stinesses at the axles

cS1 , cS2 .

The various steering

tendencies are arranged in the table 7.1.

understeering

0 > 0A

or

a1 cS1 < a2 cS2

or

a1 cS1
<1
a2 cS2

neutral

0 = 0A

or

a1 cS1 = a2 cS2

or

a1 cS1
=1
a2 cS2

oversteering

0 < 0A

or

a1 cS1 > a2 cS2

or

a1 cS1
>1
a2 cS2

Table 7.1: Steering tendency of a vehicle at forward driving

7.3.7.3 Slip Angles


With the conditions for a steady state solution

st = 0, st = 0

and the relation

Eq. (7.120), the equations of motion Eq. (7.91) and Eq. (7.92) can be dissolved for the
lateral forces

Fy1st =
Fy2st =

a2
v2
,
m
a1 + a2
R

or

a1
v2
m
a1 + a2
R

a1
Fy2st
.
=
a2
Fy1st

(7.123)

With the linear tire model in Eq. (7.74) one gets

st
Fy1
= cS1 sst
y1
where

sst
yA1

and

sst
yA2

and

st
Fy2
= cS2 sst
y2 ,

(7.124)

label the steady state lateral slips at the axles. Now, from Eqs. (7.123)

and (7.124) it follows

st
Fy2
cS2 sst
a1
y2
= st =
a2
Fy1
cS1 sst
y1

or

sst
a1 cS1
y2
= st .
a2 cS2
sy1

That means, at a vehicle with understeering tendency (a1 cS1

< a2 cS2 )

(7.125)

during steady

state cornering the slip angles at the front axle are larger than the slip angles at the rear
st
st
axle, sy1 > sy2 . So, the steering tendency can also be determined from the slip angle at
the axles.

7.3.8 Influence of Wheel Load on Cornering Stiffness


With identical tires at the front and rear axle, given a linear inuence of wheel load on
the raise of the lateral force over the lateral slip,

clin
S1 = cS Fz1

and

clin
S2 = cS Fz2 .

(7.126)

133

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

holds. The weight of the vehicle

G = mg

is distributed over the axles according to the

position of the center of gravity

Fz1 =

a2
G
a1 + a2

.Fz2 =

and

a1
G
a1 + a2

(7.127)

With Eq. (7.126) and Eq. (7.127) one obtains

a1 clin
S1 = a1 cS

a2
G
a1 + a2

(7.128)

a2 clin
S2 = a2 cS

a1
G.
a1 + a2

(7.129)

and

Thus, a vehicle with identical tires would be steering neutrally at a linear inuence of the
wheel load on the cornering stiness, because of

lin
a1 clin
S1 = a2 cS2

(7.130)

The lateral force is applied behind the center of the contact patch at the caster oset
v
distance. Hence, the lever arms of the lateral forces change to a1 a1
n and
|v| L1
v
a2 a2 + |v| nL1 , which will stabilize the vehicle, independently from the driving direction.
6
5

Fz [N ]

Fy [kN]

4
3
2
1
0

Fz [kN]
Figure 7.15: Lateral force

Fy

over wheel load

Fz

Fy [N ]

1000

758

2000

1438

3000

2043

4000

2576

5000

3039

6000

3434

7000

3762

8000

4025

at dierent slip angles

At a real tire, a degressive inuence of the wheel load on the tire forces is observed,
Fig. 7.15. According to Eq. (7.92) the rotation of the vehicle is stable, if the torque from
the lateral forces

Fy1

and

Fy2

is aligning, i.e.

a1 Fy1 a2 Fy2 < 0

(7.131)

a = 2.45 m the axle loads Fz1 = 4000 N and Fz2 =


3000 N yield the position of the center of gravity a1 = 1.05 m and a2 = 1.40 m. At equal
slip on front and rear axle one gets from the table in 7.15 Fy1 = 2576 N and Fy2 = 2043 N .
holds. At a vehicle with the wheel base

134

FH Regensburg, University of Applied Sciences

With this, the condition in Eq. (7.131) yields

Prof. Dr.-Ing. G. Rill

1.05 2576 1.45 2043 = 257.55 .

The

value is signicantly negative and thus stabilizing.


Vehicles with

a1 < a 2

have a stable, i.e. understeering driving behavior. If the axle load at

the rear axle is larger than at the front axle (a1

> a2 ), generally a stable driving behavior

can only be achieved with dierent tires.


At increasing lateral acceleration the vehicle is more and more supported by the outer
wheels. The wheel load dierences can dier at a suciently rigid vehicle body, because of
dierent kinematics (roll support) or dierent roll stiness. Due to the degressive inuence
of wheel load, the lateral force at an axle decreases with increasing wheel load dierence.
If the wheel load is split more strongly at the front axle than at the rear axle, the lateral
force potential at the front axle will decrease more than at the rear axle and the vehicle
will become more stable with an increasing lateral force, i.e. more understeering.

135

8 Driving Behavior of Single Vehicles


8.1 Standard Driving Maneuvers
8.1.1 Steady State Cornering
The steering tendency of a real vehicle is determined by the driving maneuver called
steady state cornering. The maneuver is performed quasi-static. The driver tries to keep
the vehicle on a circle with the given radius

R. He slowly increases the driving


2
ay = vR until reaching the limit.

and, with this also the lateral acceleration due

speed

Typical

80

60

side slip angle [deg]

steer angle [deg]

results are displayed in Fig. 8.1.

40
20

0
-2

-4

wheel loads [kN]

roll angle [deg]

5
3
2
1
0

4
3
2
1

0.2
0.4
0.6
0.8
lateral acceleration [g]

0.2
0.4
0.6
0.8
lateral acceleration [g]

Figure 8.1: Steady state cornering: rear-wheel-driven car on

R = 100 m

The vehicle is under-steering and thus stable according to Eq. (7.121) with Eq. (7.122).
The inclination in the diagram steering angle versus lateral velocity decides about the
steering tendency and stability behavior.

136

FH Regensburg, University of Applied Sciences

Prof. Dr.-Ing. G. Rill

The nonlinear inuence of the wheel load on the tire performance is here used to design
a vehicle that is weakly stable, but sensitive to steer input in the lower range of lateral
acceleration, and is very stable but less sensitive to steer input in limit conditions.
With the increase of the lateral acceleration the roll angle becomes larger. The overturning
torque is intercepted by according wheel load dierences between the outer and inner
wheels. With a suciently rigid frame the use of an anti roll bar at the front axle allows
to increase the wheel load dierence there and to decrease it at the rear axle accordingly.
Thus, the digressive inuence of the wheel load on the tire properties, cornering stiness
and maximum possible lateral force, is stressed more strongly at the front axle, and the
vehicle becomes more under-steering and stable at increasing lateral acceleration, until it
drifts out of the curve over the front axle in the limit situation.
Problems occur at front driven vehicles, because due to the demand for traction, the front
axle cannot be relieved at will.
Having a suciently large test site, the steady state cornering maneuver can also be
carried out at constant speed. There, the steering wheel is slowly turned until the vehicle
reaches the limit range. That way also weakly motorized vehicles can be tested at high
lateral accelerations.

8.1.2 Step Steer Input


The dynamic response of a vehicle is often tested with a step steer input. Methods for
the calculation and evaluation of an ideal response, as used in system theory or control
technics, can not be used with a real car, for a step input at the steering wheel is not
possible in practice. A real steering angle gradient is displayed in Fig. 8.2.

steering angle [deg]

40
30
20
10
0

0.2

0.4
0.6
time [s]

0.8

Figure 8.2: Step steer input


Not the angle at the steering wheel is the decisive factor for the driving behavior, but the
steering angle at the wheels, which can dier from the steering wheel angle because of
elasticities, friction inuences, and a servo-support. At very fast steering movements, also
the dynamics of the tire forces plays an important role.
In practice, a step steer input is usually only used to judge vehicles subjectively. Exceeds
in yaw velocity, roll angle, and especially sideslip angle are felt as annoying.

137

FH Regensburg, University of Applied Sciences

12

0.5

10

0.4

yaw velocity [deg/s]

0.6

0.3
0.2
0.1

6
4
2

2.5

0.5
side slip angle [deg]

roll angle [deg]

lateral acceleration [g]

Vehicle Dynamics

2
1.5
1
0.5
0

0
-0.5
-1
-1.5
-2

Figure 8.3: Step steer: passenger car at

[t]

v = 100 km/h

The vehicle under consideration behaves dynamically very well, Fig. 8.3. Almost no overshoots occur in the time history of the roll angle and the lateral acceleration. However,
small overshoots can be noticed at yaw the velocity and the sideslip angle.

8.1.3 Driving Straight Ahead


8.1.3.1 Random Road Profile
The irregularities of a track are of stochastic nature. Fig. 8.4 shows a country road prole in
dierent scalings. To limit the eort of the stochastic description of a track, one usually
employs simplifying models. Instead of a fully two-dimensional description either two
parallel tracks are evaluated

z = z(x, y)

z1 = z1 (s1 ) ,

and

z2 = z2 (s2 )

(8.1)

or one uses an isotropic track. The statistic properties are direction-independent at an


isotropic track. Then, a two-dimensional track can be approximated by a single random
process

z = z(x, y)

138

z = z(s) ;

(8.2)

FH Regensburg, University of Applied Sciences

0.05
0.04
0.03
0.02
0.01
0
-0.01
-0.02
-0.03
-0.04
-0.05
0

10

20

30

40

50

60

70

80

90

100

Prof. Dr.-Ing. G. Rill

Figure 8.4: Track irregularities

A normally distributed, stationary and ergodic random process

z = z(s)

is completely

characterized by the rst two expectation values, the mean value

1
mz = lim
s 2s

Zs
z(s) ds

(8.3)

s
and the correlation function

Zs

1
Rzz () = lim
s 2s

z(s) z(s ) ds .

(8.4)

s
A vanishing mean value

mz = 0

can always be achieved by an appropriate coordinate

transformation. The correlation function is symmetric,

Rzz () = Rzz () ,
and

1
Rzz (0) = lim
s 2s

Zs
z(s)

(8.5)

2

ds

(8.6)

s
describes the variance of

zs .

Stochastic track irregularities are mostly described by power spectral densities (abbreviated by psd). Correlating function and the one-sided power spectral density are linked by
the Fourier-transformation

Z
Rzz () =

Szz () cos() d

(8.7)

0
where

denotes the space circular frequency. With Eq. (8.7) follows from Eq. (8.6)

Z
Rzz (0) =

Szz () d .

(8.8)

139

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Thus, the psd gives information, how the variance is compiled from the single frequency
shares.
The power spectral densities of real tracks can be approximated by the relation


Szz () = S0

w
,

(8.9)

0 = 1 m1 . The reference psd S0 = Szz (0 )


acts as a measurement for unevennes and the waviness w indicates, whether the track has

where the reference frequency is xed to

notable irregularities in the short or long wave spectrum. At real tracks, the reference6
psd S0 lies within the range from 1 10
m3 to 100 106 m3 and the waviness can be
approximated by

w = 2.

8.1.3.2 Steering Activity


-6

highway: S 0=1*10

-5

m ; w=2

country road: S0=2*10

1000

1000

500

500

-2

[deg] 2

-2

m ; w=2

[deg] 2

Figure 8.5: Steering activity on dierent roads


A straightforward drive upon an uneven track makes continuous steering corrections necessary. The histograms of the steering angle at a driving speed of

v = 90 km/h

are

displayed in Fig. 8.5. The track quality is reected in the amount of steering actions. The
steering activity is often used to judge a vehicle in practice.

8.2 Coach with different Loading Conditions


8.2.1 Data
The dierence between empty and laden is sometimes very large at trucks and coaches.
In the table 8.1 all relevant data of a travel coach in fully laden and empty condition are
listed.

140

FH Regensburg, University of Applied Sciences

vehicle

mass

[kg]

center of gravity

[m]

empty

12 500

3.800 | 0.000 | 1.500

fully laden

18 000

3.860 | 0.000 | 1.600

Prof. Dr.-Ing. G. Rill

inertias

[kg m2 ]

12 500
0
0
0
155 000
0
0
0
155 000
15 400
0
250
0
200 550
0
250
0
202 160

Table 8.1: Data for a laden and empty coach

The coach has a wheel base of

a = 6.25m. The front axle with the track width sv = 2.046m

has a double wishbone single wheel suspension. The twin-tire rear axle with the track
i
o
widths sh = 2.152 m and sh = 1.492 m is guided by two longitudinal links and an a-arm.
The air-springs are tted to load variations via a niveau-control.

suspension travel [cm]

8.2.2 Roll Steering


10
5
0
-5
-10
-1

0
steer angle [deg]

Figure 8.6: Roll steer: - - front,  rear


While the kinematics at the front axle hardly cause steering movements at roll motions,
the kinematics at the rear axle are tuned in a way to cause a notable roll steering eect,
Fig. 8.6.

8.2.3 Steady State Cornering


Fig. 8.7 shows the results of a steady state cornering on a

100 m-Radius.

The fully occu-

pied vehicle is slightly more understeering than the empty one. The higher wheel loads
cause greater tire aligning torques and increase the degressive wheel load inuence on the
increase of the lateral forces. Additionally roll steering at the rear axle occurs.

141

Vehicle Dynamics

steer angle
250

FH Regensburg, University of Applied Sciences

LW

[deg]

vehicle course
200

200
[m]

150

150

100

100

50

50

0.1 0.2 0.3 0.4


lateral acceleration a y [g]
wheel loads [kN]

100

50

-100

0
[m]

100

wheel loads [kN]

100

50

0.1 0.2 0.3 0.4


lateral acceleration a y [g]

0.1 0.2 0.3 0.4


lateral acceleration a y [g]

Figure 8.7: Steady state cornering: coach - - empty,

 fully occupied

Both vehicles can not be kept on the given radius in the limit range. Due to the high position of the center of gravity the maximal lateral acceleration is limited by the overturning
hazard. At the empty vehicle, the inner front wheel lift o at a lateral acceleration of

ay 0.4 g

. If the vehicle is fully occupied, this eect will occur already at

ay 0.35 g .

8.2.4 Step Steer Input


The results of a step steer input at the driving speed of

v = 80 km/h

can be seen in

Fig. 8.8. To achieve comparable acceleration values in steady state condition, the step

steer input was done at the empty vehicle with = 90 and at the fully occupied one

with = 135 . The steady state roll angle is 50% larger at the fully occupied bus than
at the empty one. By the niveau-control, the air spring stiness increases with the load.
Because the damper eect remains unchanged, the fully laden vehicle is not damped as
well as the empty one. This results in larger overshoots in the time histories of the lateral
acceleration, the yaw angular velocity, and the sideslip angle.

142

FH Regensburg, University of Applied Sciences

yaw velocity

lateral acceleration a y [g]


0.4

10

0.2

0.1

2
0

roll angle

[deg]

side slip angle

[deg]

0
-1

2
0

Z [deg/s]

0.3

Prof. Dr.-Ing. G. Rill

-2
0

[s] 6

Figure 8.8: Step steer input:

- - coach empty,

[s] 6

 coach fully occupied

8.3 Different Rear Axle Concepts for a Passenger Car


A medium-sized passenger car is equipped in standard design with a semi-trailing rear
axle. By accordingly changed data this axle can easily be transformed into a trailing arm
or a single wishbone axis. According to the roll support, the semi-trailing axle realized
in serial production represents a compromise between the trailing arm and the single
wishbone, Fig. 8.9, .
The inuences on the driving behavior at steady state cornering on a

100 m

radius are

shown in Fig. 8.10.


Substituting the semi-trailing arm at the standard car by a single wishbone, one gets,
without adaption of the other system parameters a vehicle oversteering in the limit range.
Compared to the semi-trailing arm the single wishbone causes a notably higher roll support. This increases the wheel load dierence at the rear axle, Fig. 8.10. Because the wheel
load dierence is simultaneously reduced at the front axle, the understeering tendency is
reduced. In the limit range, this even leads to an oversteering behavior.

143

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

vertical motion [cm]

10

Figure 8.9: Rear axle:

steer angle

5
0
-5
-10
-5

 semi-trailing arm, - - single wishbone,

LW

100

0
lateral motion [cm]

[deg]

roll angle

trailing arm

[Grad]

4
3

50

2
1

0.2

0.4

0.6

0.8

wheel loads front [kN]

2
0

0.2
0.4
0.6
0.8
lateral acceleration a y [g]

0.2

0.4

0.6

0.8

wheel loads rear [kN]

0.2
0.4
0.6
0.8
lateral acceleration a y [g]

Figure 8.10: Steady state cornering,  semi-trailing arm, - - single wishbone,

trailing

arm

The vehicle with a trailing arm rear axle is, compared to the serial car, more understeering.
The lack of roll support at the rear axle also causes a larger roll angle.

144

Index

Ackermann geometry, 109

Damping rate, 76

Ackermann steering angle, 109, 132

Deviation, 13

Aerodynamic forces, 96

Disturbance-reaction problem, 84

Air resistance, 96

Disturbing force lever, 9

All wheel drive, 118

Down forces, 96

Anti dive, 108

Downhill capacity, 97

Anti roll bar, 122

Drag link, 57, 58

Anti squat, 108

Drive pitch angle, 103

Anti-lock-system, 102

Driver, 3

Auto-correlation, 13

Driving force distribution, 100

Axle kinematics, 108

Driving safety, 72

Double wishbone, 7

Dynamic axle load, 95

McPherson, 7

Dynamic force elements, 63

Multi-link, 7

Dynamic wheel loads, 94

Axle load, 95
Axle suspension
Solid axle, 55
Twist beam, 56
Bend angle, 113, 116
Brake pitch angle, 103
Brake pitch pole, 108
Braking force distribution, 100
Camber angle, 6, 24
Camber compensation, 121, 124
Camber slip, 49
Caster, 8, 9
Characteristic speed, 131

Eective value, 13
Eigenvalues, 128
Environment, 4
First harmonic oscillation, 63
Fourier-approximation, 64
Frequency domain, 63
Friction, 97
Front wheel drive, 98, 118
Generalized uid mass, 70
Grade, 95
Hydro-mount, 69

Climbing capacity, 97

Kingpin, 7

Comfort, 72

Kingpin Angle, 8

Contact point, 24
Cornering resistance, 117, 118

Lateral acceleration, 121, 132

Cornering stiness, 41, 133

Lateral force, 126

Curvature gradient, 114

Lateral slip, 126

Vehicle Dynamics

FH Regensburg, University of Applied Sciences

Ljapunov equation, 84

State vector, 88

Load, 4

Steady state cornering, 117, 136, 141

Maximum acceleration, 97, 98


Maximum deceleration, 97, 99
Mean value, 13
Natural frequency, 76

Steering activity, 140


Steering angle, 114
Steering box, 57, 58
Steering lever, 58
Steering oset, 9
Steering system

Optimal damping, 81, 87

Drag link steering system, 58

Oversteering, 133

Lever arm, 57

Overturning limit, 118

Rack and pinion, 57

Parallel track model, 10


Parallel tracks, 138
Pinion, 57
Pivot pole, 109
Power spectral density, 139
Quarter car model, 87, 90

Steering tendency, 125, 132


Step steer input, 137, 143
Suspension model, 75
Suspension spring rate, 78
System response, 63
Tilting condition, 97
Tire

Rack, 57

Bore slip, 52

Random road prole, 138

Bore torque, 20, 50

Rear wheel drive, 98, 118

Camber angle, 24

Reference frames

Camber inuence, 48

Ground xed, 5

Characteristics, 52

Inertial, 5

Circumferential direction, 24

Vehicle xed, 5

Composites, 19

Relative damping rate, 77

Contact forces, 20

Ride comfort, 83

Contact patch, 20

Ride safety, 83

Contact point, 23

Road, 10, 23

Contact point velocity, 31

Roll axis, 124

Contact torques, 20

Roll center, 124

Cornering stiness, 41

Roll steer, 141

Deection, 26

Roll stiness, 120

Deformation velocity, 31

Roll support, 121, 124

Development, 19

Rolling condition, 126

Dynamic oset, 41
Dynamic radius, 32, 33

Safety, 72

Friction coecient, 45

Side slip angle, 109

Lateral direction, 24

Sky hook damper, 87

Lateral force, 20

Space requirement, 110

Lateral force characteristics, 41

Spring rate, 78

Lateral force distribution, 40

Stability, 128

Lateral slip, 40

State equation, 128

Lateral velocity, 31

State matrix, 88

ii

FH Regensburg, University of Applied Sciences

Lift o, 88

Prof. Dr.-Ing. G. Rill

Vehicle model, 75, 90, 94, 103, 112, 121,

Linear model, 126

125

Loaded radius, 24, 32

Vertical dynamics, 72

Longitudinal force, 20, 38, 39

Virtual work, 121

Longitudinal force characteristics, 39


Longitudinal force distribution, 39
Longitudinal slip, 39
Longitudinal velocity, 31
Model, 52
Normal force, 20
Pneumatic trail, 41
Radial damping, 35
Radial direction, 24
Radial stiness, 34, 121
Rolling resistance, 20, 36, 37
Self aligning torque, 20, 41
Sliding velocity, 40
Static radius, 24, 32, 33
Tilting torque, 20
Track normal, 24, 26
Transport velocity, 33
Tread deection, 38
Tread particles, 37
Unloaded radius, 32
Vertical force, 34
Wheel load inuence, 41

Waviness, 140
Wheel
Angular velocity, 50
Wheel base, 109
Wheel load, 20
Wheel loads, 94
Wheel rotation axis, 5
Wheel Suspension
Semi-trailing arm, 143
Single wishbone, 143
Trailing arm, 143
Wheel suspension
Central control arm, 56
Double wishbone, 55
McPherson, 55
Multi-Link, 55
Semi-trailing arm, 56
SLA, 56
Yaw angle, 112, 115
Yaw velocity, 126

Tire Model
Kinematic, 109
Linear, 133
TMeasy, 52
Toe angle, 6
Toe-in angle, 6
Track, 23
Track curvature, 114
Track normal, 5
Track radius, 114
Track width, 109, 121
Trailer, 112, 115
Understeering, 133
Variance, 13
Vehicle, 3
Vehicle comfort, 72
Vehicle dynamics, 2

iii

You might also like