Mathematical Foundation of Railroad Vehicle Systems Geometry and Mechanics, Ahmed A. Shabana, 2021

Download as pdf or txt
Download as pdf or txt
You are on page 1of 380

Mathematical Foundation of Railroad Vehicle Systems

Geometry and Mechanics


Mathematical Foundation of Railroad Vehicle
Systems

Geometry and Mechanics

Ahmed A. Shabana
Richard and Loan Hill Professor of Engineering
University of Illinois at Chicago
Chicago, Illinois, USA
This edition first published 2021
© 2021 John Wiley & Sons Ltd

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form
or by any means, electronic, mechanical, photocopying, recording or otherwise, except as permitted by law. Advice on how
to obtain permission to reuse material from this title is available at http://www.wiley.com/go/permissions.

The right of Ahmed A. Shabana to be identified as the author of this work has been asserted in accordance with law.

Registered Offices
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

Editorial Office
The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

For details of our global editorial offices, customer services, and more information about Wiley products visit us at www
.wiley.com.

Wiley also publishes its books in a variety of electronic formats and by print-on-demand. Some content that appears in
standard print versions of this book may not be available in other formats.

Limit of Liability/Disclaimer of Warranty


In view of ongoing research, equipment modifications, changes in governmental regulations, and the constant flow of
information relating to the use of experimental reagents, equipment, and devices, the reader is urged to review and
evaluate the information provided in the package insert or instructions for each chemical, piece of equipment, reagent, or
device for, among other things, any changes in the instructions or indication of usage and for added warnings and
precautions. While the publisher and authors have used their best efforts in preparing this work, they make no
representations or warranties with respect to the accuracy or completeness of the contents of this work and specifically
disclaim all warranties, including without limitation any implied warranties of merchantability or fitness for a particular
purpose. No warranty may be created or extended by sales representatives, written sales materials or promotional
statements for this work. The fact that an organization, website, or product is referred to in this work as a citation and/or
potential source of further information does not mean that the publisher and authors endorse the information or services
the organization, website, or product may provide or recommendations it may make. This work is sold with the
understanding that the publisher is not engaged in rendering professional services. The advice and strategies contained
herein may not be suitable for your situation. You should consult with a specialist where appropriate. Further, readers
should be aware that websites listed in this work may have changed or disappeared between when this work was written
and when it is read. Neither the publisher nor authors shall be liable for any loss of profit or any other commercial
damages, including but not limited to special, incidental, consequential, or other damages.

Library of Congress Cataloging-in-Publication Data

Names: Shabana, Ahmed A., 1951- author.


Title: Mathematical foundation of railroad vehicle systems: geometry and
mechanics / Ahmed A. Shabana, University of Illinois at Chicago Circle,
IL, US.
Description: Hoboken, NJ : John Wiley & Sons, Inc., [2021] | Includes
bibliographical references and index.
Identifiers: LCCN 2020020731 (print) | LCCN 2020020732 (ebook) | ISBN
9781119689041 (cloth) | ISBN 9781119689065 (adobe pdf) | ISBN
9781119689089 (epub)
Subjects: LCSH: Railroad engineering–Mathematics. | Railroad
rails–Mathematical models. | Railroad trains–Dynamics. | Geometric
analysis.
Classification: LCC TF205 .S47 2020 (print) | LCC TF205 (ebook) | DDC
625.1001/5118–dc23
LC record available at https://lccn.loc.gov/2020020731
LC ebook record available at https://lccn.loc.gov/2020020732

Cover Design: Wiley


Cover Image: © Xuanyu Han/Getty Images

Set in 9.5/12.5pt STIXTwoText by SPi Global, Chennai, India

10 9 8 7 6 5 4 3 2 1
v

Contents

Preface ix

1 Introduction 1
1.1 Differential Geometry 4
1.2 Integration of Geometry and Mechanics 9
1.3 Hunting Oscillations 14
1.4 Wheel and Track Geometries 17
1.5 Centrifugal Forces and Balance Speed 22
1.6 Contact Formulations 26
1.7 Computational MBS Approaches 28
1.8 Derailment Criteria 33
1.9 High-Speed Rail Systems 36
1.10 Linear Algebra and Book Notations 41

2 Differential Geometry 45
2.1 Curve Geometry 46
2.2 Surface Geometry 54
2.3 Application to Railroad Geometry 57
2.4 Surface Tangent Plane and Normal Vector 60
2.5 Surface Fundamental Forms 62
2.6 Normal Curvature 69
2.7 Principal Curvatures and Directions 72
2.8 Numerical Representation of the Profile Geometry 76
2.9 Numerical Representation of Surface Geometry 78

3 Motion and Geometry Descriptions 83


3.1 Rigid-Body Kinematics 84
3.2 Direction Cosines and Simple Rotations 86
3.3 Euler Angles 88
3.4 Euler Parameters 91
3.5 Velocity and Acceleration Equations 95
3.6 Generalized Coordinates 97
3.7 Kinematic Singularities 100
vi Contents

3.8 Euler Angles and Track Geometry 102


3.9 Angle Representation of the Curve Geometry 107
3.10 Euler Angles as Field Variables 108
3.11 Euler-Angle Description of the Track Geometry 111
3.12 Geometric Motion Constraints 114
3.13 Trajectory Coordinates 119

4 Railroad Geometry 125


4.1 Wheel Surface Geometry 126
4.2 Wheel Curvatures and Global Vectors 132
4.3 Semi-analytical Approach for Rail Geometry 135
4.4 ANCF Rail Geometry 142
4.5 ANCF Interpolation of Rail Geometry 145
4.6 ANCF Computation of Tangents and Normal 146
4.7 Track Geometry Equations 148
4.8 Numerical Representation of Track Geometry 152
4.9 Track Data 155
4.10 Irregularities and Measured Track Data 162
4.11 Comparison of the Semi-Analytical and ANCF Approaches 169

5 Contact Problem 175


5.1 Wheel/Rail Contact Mechanism 177
5.2 Constraint Contact Formulation (CCF) 183
5.3 Elastic Contact Formulation (ECF) 184
5.4 Normal Contact Forces 187
5.5 Contact Surface Geometry 188
5.6 Contact Ellipse and Normal Contact Force 194
5.7 Creepage Definitions 199
5.8 Creep Force Formulations 203
5.9 Creep Force and Wheel/Rail Contact Formulations 213
5.10 Maglev Forces 219

6 Equations of Motion 225


6.1 Newtonian and Lagrangian Approaches 226
6.2 Virtual Work Principle and Constrained Dynamics 227
6.3 Summary of Rigid-Body Kinematics 232
6.4 Inertia Forces 235
6.5 Applied Forces 239
6.6 Newton–Euler Equations 241
6.7 Augmented Formulation and Embedding Technique 244
6.8 Wheel/Rail Constraint Contact Forces 254
6.9 Wheel/Rail Elastic Contact Forces 259
6.10 Other Force Elements 261
6.11 Trajectory Coordinates 268
6.12 Longitudinal Train Dynamics (LTD) 274
Contents vii

6.13 Hunting Stability 280


6.14 MBS Modeling of Electromechanical Systems 288

7 Pantograph/Catenary Systems 291


7.1 Pantograph/Catenary Design 292
7.2 ANCF Catenary Kinematic Equations 298
7.3 Catenary Inertia and Elastic Forces 304
7.4 Catenary Equations of Motion 306
7.5 Pantograph/Catenary Contact Frame 308
7.6 Constraint Contact Formulation (CCF) 310
7.7 Elastic Contact Formulation (ECF) 314
7.8 Pantograph/Catenary Equations and MBS Algorithms 317
7.9 Pantograph/Catenary Contact Force Control 321
7.10 Aerodynamic Forces 322
7.11 Pantograph/Catenary Wear 324

Appendix Contact Equations and Elliptical Integrals 329


A.1 Derivation of the Contact Equations 329
A.2 Elliptical Integrals 332

Bibliography 335
Index 355
ix

Preface

The design of passenger and freight railroad vehicle systems, which are complex
transportation systems, involves many areas of science and engineering, including
mechanical, civil, electrical, and electronic engineering. Because a comprehensive treat-
ment of such complex systems cannot be covered in a single book, the focus of this book is
on developing the mathematical foundation of railroad vehicle systems with an emphasis
on the integration of geometry and mechanics. Such a geometry/mechanics integration
is necessary for developing a sound mathematical foundation, accurate formulation of
nonlinear dynamic equations, and general computational algorithms that can be used
effectively in the virtual prototyping, analysis, design, and performance evaluation of
railroad vehicle systems. Geometry is particularly important in the formulation of the
railroad system dynamic equations of the motion generated by contact between the wheel
and rail surfaces. The surface geometry, therefore, plays a fundamental role in formulating
the kinematics, forces, and equations of motion of such systems. The theory of curves is
equally important, and it is central in the description of the track geometry. For this reason,
the subject of differential geometry plays a fundamental role in formulating the equations
that govern the motion of railroad vehicle systems.
In addition to basic geometry concepts, principles of mechanics are required in order
to formulate railroad kinematic relationships and dynamic equations of motion. These
mechanics principles allow for systematically modeling motion constraints resulting from
mechanical joints and specified motion trajectories. Railroad vehicles have components
that experience large displacements, including finite rotations; therefore, it is necessary
to avoid linearization of the kinematic and dynamic equations when developing general
computational algorithms. In the mechanics approaches used in this book, a fully non-
linear motion description is used in the formulation of dynamic equations of motion. As
demonstrated in this book, applying mechanics principles can lead to different forms of
dynamic equations of motion. Regardless of the form of the equations of motion obtained,
the integration of geometry and mechanics is necessary.
This book is written to complement a previous book, Railroad Vehicle Dynamics: A Com-
putational Approach. Both books are designed to be self contained, and therefore, overlap
of some topics cannot be avoided. This book is designed for an introductory course on
railroad vehicle system dynamics suitable for senior undergraduate and first-year gradu-
ate students. It can also be used as a reference book by researchers and practicing engi-
neers. The book consists of seven chapters that are organized to introduce the reader to the
x Preface

basic concepts, formulations, and computational algorithms used in railroad vehicle system
dynamics. Unlike other texts in the area of railroad vehicle dynamics, this book empha-
sizes geometry/mechanics integration throughout. It is shown how new mechanics-based
approaches, such as the absolute nodal coordinate formulation (ANCF), can be used to
achieve the geometry/mechanics integration necessary for developing accurate virtual pro-
totyping algorithms.
Chapter 1 introduces the reader to several topics that are covered in subsequent chapters.
Basic geometry concepts are discussed, and the need to integrate geometry and mechan-
ics in the formulation of the railroad vehicle system equations is explained. The hunting
oscillations that characterize the dynamic behavior of railroad vehicles and can lead to
instabilities and derailments are discussed, and the methods used to describe wheel and
track geometries are introduced. During curve negotiations, the effect of centrifugal forces
must be taken into account to define the allowable balance speed that ensures the safe oper-
ation of the rail vehicle. This important topic, as well as the wheel/rail contact formulations
and their integration with multibody system (MBS) computational algorithms, are intro-
duced in Chapter 1. The chapter also discusses other important topics, including derailment
criteria, high-speed rail, pantograph/catenary systems, and linear algebra, which can be
conveniently used to formulate the kinematic and dynamic equations of railroad vehicle
systems.
Chapter 2 covers topics in differential geometry that are fundamental in developing the
nonlinear equations that govern the motion of the vehicle. Curve and surface geometries
and their application to railroad vehicle systems are discussed. It is shown how the principal
curvatures and principal directions of surfaces can be determined using the coefficients
of the first and second fundamental forms of surfaces. The numerical representation of
wheel and rail profile geometries using spline functions is discussed. The use of ANCF
finite elements to describe the rail surface geometry is explained. It is shown that arbitrary
surface geometry for the rail can be systematically described using the displacement field
of fully parameterized ANCF finite elements.
Railroad vehicle motion and geometry descriptions are the subject of Chapter 3, which
demonstrates how body motion equations and geometry are integrated. The coordinate sys-
tems used to describe rigid-body kinematics are first introduced, and different parameters
for describing body orientation are presented. Among the orientation parameters discussed
in this chapter are direction cosines, Euler angles, and Euler parameters. These orientation
parameters are used to define the position, velocity, and acceleration equations in terms of
the body generalized coordinates. The chapter explains the use of Euler angles in railroad
vehicle system mechanics for two fundamentally different purposes: motion description
and track geometry description. Using Euler angles as field variables to describe track geom-
etry is explained. In Chapter 3, geometric motion constraints that result from mechanical
joints used to connect vehicle components are introduced, and the trajectory coordinates
used to develop specialized railroad vehicle system computer programs are defined.
While Chapters 2 and 3 deal, for the most part, with basic geometry and motion kinematic
approaches and concepts, the focus of Chapter 4 is on railroad geometry. Wheel geome-
try equations are first defined, and it is shown how these geometry equations are used to
define curvatures at an arbitrary point on the wheel surface. Two methods are introduced
for describing rail geometry: semi-analytical and ANCF. Limitations of the semi-analytical
Preface xi

approach are discussed to provide a justification for using the more general ANCF approach
to describe rail geometry. Using the ANCF interpolation to define the tangents and nor-
mal at an arbitrary point on the rail surface is explained. As discussed in this chapter,
such an ANCF approach, which employs position gradients, allows using higher-order
interpolation for the position coordinates and avoids the interpolation of rotations. Fully
parametrized ANCF finite elements can be used to systematically define surfaces by intro-
ducing a functional relationship between the element spatial coordinates. The structure of
the track data used in computer simulations of railroad vehicle systems is discussed, and
methods for describing track deviations that represent track irregularities are presented.
The chapter concludes by providing a comparison between the semi-analytical and ANCF
approaches used to describe rail geometry.
The contact problem is covered in Chapter 5, starting with a discussion of the wheel/rail
contact mechanism in order to provide a better understanding of the need to account for
the creep phenomenon in railroad vehicle system applications. The constraint contact for-
mulation (CCF) and the elastic contact formulation (ECF), which are used to determine the
locations of wheel/rail contact points online, are introduced. As explained in this chapter,
the CCF approach does not allow wheel/rail separation, but the ECF approach does allow
this separation. Therefore, when the ECF approach is used, no degrees of freedom are
eliminated. Determining the normal contact force in both the CCF and ECF approaches
is discussed in this chapter, and it is explained how the surface geometry and Hertz con-
tact theory can be used to determine the dimensions of the contact area. To determine the
tangential creep forces, velocity creepages used in wheel/rail creep force formulations are
defined. Several creep force formulations are discussed in Chapter 5, and the assumptions
used in these formulations are highlighted. The chapter concludes with a discussion of mag-
netically levitated (maglev) trains.
Methods for developing the nonlinear dynamic equations of motion of railroad vehi-
cle systems are introduced in Chapter 6. The Newtonian and Lagrangian approaches are
compared to highlight the basic concepts used in the two approaches. As discussed in this
chapter, the Lagrangian approach does not require the use of free-body diagrams or mak-
ing cuts at the joints because this approach is based on using connectivity conditions. The
constraint forces can be defined systematically using algebraic constraint equations, and
therefore, the Lagrangian approach lends itself to developing general computational algo-
rithms for railroad vehicle systems. Using generalized coordinates, the inertia and applied
and contact forces can be formulated and used to develop equations of motion. The form of
the equations of motion can be defined using the augmented formulation or the embedding
technique. The augmented formulation leads to a large, sparse system expressed in terms
of redundant coordinates and constraint forces, while the embedding technique leads to
a minimum number of equations of motion expressed in terms of the system degrees of
freedom; therefore, these equations do not include any constraint forces. The chapter also
discusses the formulation of other force elements used in railroad vehicle dynamics and
explains using trajectory coordinates to develop longitudinal train dynamics (LTD) algo-
rithms. Simple models that can be used to study hunting oscillations are also developed in
Chapter 6. The chapter concludes with a discussion of MBS modeling of electromechanical
systems.
xii Preface

Pantograph/catenary systems that provide the power supply for high-speed trains are
the subject of Chapter 7. The chapter discusses in detail their design and the formula-
tion of catenary equations of motion using ANCF finite elements. The formulation of
pantograph/catenary contact forces is presented using the constraint and elastic contact
formulations. It is shown how to account for lateral relative sliding between the pan-head
and catenary when using the constraint and elastic contact formulations. Chapter 6 also
discusses pantograph/catenary force control, the effect of aerodynamic forces, and the
wear problem when such current-collection systems are used for high-speed trains.
The development of the materials presented in this book is based on research collabo-
ration between the author and many colleagues and students. The author would like to
acknowledge the contributions of his colleagues and students including UIC current and
former students and visiting scholars Zhengfeng Bai, Issac Banes, Dario Bettamin, Mah-
moud Elbakly, Ahmed Eldeek, Sibi Kandasamy, Hao Ling, Ramon Martinez, Huailong Shi,
and Dayu Zhang for their help in the preparation and proofreading of the book manuscript.
The author would like to thank Drs. Khaled Zaazaa and Hiroyuki Sugiyama for past col-
laboration on writing the book Railroad Vehicle Dynamics: A Computational Approach. The
editorial and production staff of Wiley & Sons deserve special thanks for their cooperation
and professional work in producing this book. Finally, the author would like to thank his
family for their patience and understanding during the time spent preparing this book.

Ahmed A. Shabana
Chicago, Illinois, USA
1

Chapter 1

INTRODUCTION

Passenger and freight railroad vehicles are complex transportation systems whose design
involves many areas of science and engineering, including mechanical, civil, electrical, and
electronic engineering. Because, a comprehensive treatment of such railroad vehicle sys-
tems cannot be covered in a single volume book, the focus of this book is on developing
the mathematical foundation with the emphasis placed on the integration of geometry and
mechanics.

Integration of Geometry and Analysis As will be explained in this book, a sound


mathematical foundation, the formulation of nonlinear dynamic equations, and the design
of general computational algorithms of such complex systems require the integration of
basic concepts of geometry and mechanics. Geometry is particularly important in formu-
lating the railroad system dynamic equations that govern the vehicle motion produced by
contact between the wheel and rail surfaces. The surface geometry, therefore, plays a fun-
damental role in formulating the kinematics, forces, and equations of motion of such trans-
portation systems. Without an accurate description of the surface geometry, it is not possible
to develop accurate formulations and computational algorithms that account for the non-
linear dynamic behavior of railroad vehicle systems. The theory of curves is equally impor-
tant, and it is central to the description of track geometry. High-speed trains are electrically
powered using pantograph/catenary systems. As discussed in this book, the catenary sys-
tem consists of wires whose geometry and deformations are critical in ensuring high-quality
electric current collection. These catenary wires also define space curves whose geometry
must be accurately described in order to better understand their dynamic behavior and their
response to contact and aerodynamic forces.
While geometry enters into formulating railroad kinematic relationships and dynamic
equations of motion, these equations are formulated using the principles of mechanics,
which allow for systematically modeling motion constraints resulting from mechanical
joints and specified motion trajectories. Because railroad vehicles have components
that experience large displacements, including finite rotations, it is important to avoid
linearization of the kinematic and dynamic equations when developing general computa-
tional algorithms. In the mechanics approaches used in this book, fully nonlinear motion
descriptions are adopted in formulating the dynamic equations of motion. Applying the
principles of mechanics can lead to different forms of the dynamic equations of motion,
as discussed in this book. Regardless of the form of the equations of motion obtained, the
Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,
First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
2 Mathematical Foundation of Railroad Vehicle Systems

integration of geometry and mechanics is necessary in order to establish the foundation of


those equations.

Passenger and Freight Trains The methods developed in this book are applicable to
both passenger and freight railroad vehicle systems. These two types of systems can oper-
ate at significantly different speeds, have significantly different axle loads, and require the
use of different track quality. Figure 1 shows a high-speed passenger train and a freight
train that operate under different rules and guidelines and require different track-quality

(a)

(b)

Figure 1.1 Passenger and freight trains. Sources: (a) WANG SHIH-WEi/123 RF (2016); (b) Mike
Danneman/Moment/Getty Images.
Introduction 3

standards developed based on safety considerations. For both systems, however, regardless
of the power supply needed for their operation, motion is generated by the force of interac-
tion of wheels and rails like the ones shown in Figure 2. For the most part, freight trains use
diesel engines and often run on lower-quality tracks, so their speeds cannot exceed a cer-
tain limit. High-speed trains, on the other hand, are electrically powered and can operate
at a much higher speed as compared to freight trains. While the mathematical foundations
developed in this book can be applied to both cases, it is important to recognize that prac-
tical issues must be taken into consideration when modeling these two different systems.
While the basic geometric and mechanical formulations are the same, the analysis models,
including dimensions, system configurations, number of rail cars, suspension characteris-
tics, ride-comfort criteria, noise level, power supply, operating speeds, etc., can be signif-
icantly different. Developing accurate virtual prototyping models for both systems, based
on proper integration of geometry and mechanics, is necessary in order to avoid deadly,
costly, and environmentally damaging accidents. Passenger trains transport a large number
of people around the world, while freight trains transport goods and hazardous materials.
As the demand increases for higher-speed, heavier-axle loads; stricter operational and safety
guidelines; less noise; a greater degree of comfort; and more robust transportation systems;
more accurate virtual models with significant details are needed.

Tread contact Flange contact

Figure 1.2 Wheel/rail contact.

Organization and Scope of This Chapter This chapter introduces the basic top-
ics discussed in this book, starting with the differential geometry that covers the theory of
curves and surfaces. The integration of geometry and motion description plays an important
role in formulating the railroad vehicle equations of motion, as discussed in Section 1.1. In
Section 1.2, the important topic of the integration of geometry and mechanics is discussed in
more detail. Railroad wheelsets exhibit a type of vibration known as hunting oscillations dur-
ing which the wheelset lateral displacement and yaw angle are related because of wheelset
conicity. Hunting oscillations are discussed in Section 1.3 based on pure geometric consid-
erations. Using a basic description of differential geometry and motion, track and wheel
geometries can be defined, as explained in Section 1.4. Section 1.5 explains the effect of
centrifugal forces during curve negotiations and presents a simple analysis for defining the
balance speed that must not be exceeded by the vehicle when it travels on a curved section
of track, to avoid derailment. Section 1.6 introduces the wheel/rail contact problem and the
4 Mathematical Foundation of Railroad Vehicle Systems

forces that influence vehicle stability. While the main focus in this book is on trains driven
by the wheel/rail contact forces, maglev trains are also briefly discussed in Section 1.6.
In Section 1.7, the multibody system (MBS) approach for formulating governing dynamic
equations using the principles of mechanics is discussed. Section 1.8 discusses existing
derailment criteria and the role of three-dimensional computational dynamics in devel-
oping more accurate and general derailment criteria. Section 1.9 reviews topics related to
the operation of high-speed trains, including the pantograph/catenary systems. Section 1.10
discusses some mathematical preliminaries that are used throughout the chapters and the
notations adopted in this book.

1.1 DIFFERENTIAL GEOMETRY


Figure 3 shows a train negotiating a curved track, the wheel and rail that come into con-
tact, and the pantograph/catenary system used to supply the electric power necessary for
the operation of high-speed trains. The description of the motion of the vehicle on the track,
the layout of the track, the mathematical definition of the surfaces of the wheel and rail that
come into contact to produce the train motion, and the geometry and deformations of the
catenary cables that carry the electric current necessary for the operation of the high-speed
rail system are examples that demonstrate the importance of geometry in railroad vehi-
cle system dynamics. For this reason, understanding the differential-geometry theories of
curves and surfaces is the first step in correctly formulating the dynamic equations that
govern the motion of these complex systems (Do Carmo 1976; Goetz 1970; Kreyszig 1991).
For example, the track and wheel geometries must be accurately described in order to
correctly predict the wheel/rail contact forces. Using both the differential-geometry theory
of curves and the theory of surfaces is necessary in order to be able to formulate the wheel
and rail kinematic and force equations. These theories are used to define nonlinear geo-
metric equations that can be solved for the locations of the wheel/rail contact points, as
discussed in this book. Furthermore, the spatial wheel and rail geometric representation
is crucial in the study of railroad vehicle nonlinear dynamics in different motion scenar-
ios, including curve negotiations and travel on tracks with irregularities and worn profiles
that influence ride comfort, vehicle stability, and safe operation. Therefore, geometric con-
cepts are an integral part of formulating nonlinear dynamic equations and computational
algorithms that can be used effectively to develop credible virtual prototyping models that
contribute to developing operational and safety guidelines and to reducing the possibility
of train derailments and accidents.

Theory of Curves While a distinction is made between curve and surface geometry and
dynamic motion descriptions, simple motion examples can be used to explain some con-
cepts that are used in this book. In this section, the motion of a particle is used to explain
basic concepts related to the geometry of curves and surfaces. In general, three coordinates
(parameters) are required in order to define the position of a particle in space. These three
coordinates can be selected to describe the location of the particle with respect to the origin
of a coordinate system formed by three orthogonal axes, as shown in Figure 4. In this case,
the position of the particle is defined by the vector r = [x y z]T , where x, y, and z are
Introduction 5

(a)

Pantograph/catenary system

Wheel Rail

(b)

Figure 1.3 Geometry of railroad vehicle systems. Sources: (a) serjiob74/Adobe Stock; (b) Susan
Isakson/Alamy Stock Photo.

three independent parameters that define the position coordinates of the particle along the
three orthogonal axes of the coordinate system XYZ. If the particle is constrained to trace a
space curve like the one shown in Figure 4, the particle has the freedom to move only along
the curve; the particle freedom to move along two directions perpendicular to the curve is
eliminated. That is, the curve equation is completely defined in terms of one parameter, and
the three coordinates of the particle are no longer independent.
To demonstrate this simple concept, consider a particle that traces the circle shown in
Figure 5. The circle represents a curve with constant curvature defined by the radius of cur-
vature a. If the motion of the particle is restricted to trace this circular curve, two constraints
6 Mathematical Foundation of Railroad Vehicle Systems

Z r

Figure 1.4 Curve geometry.

are imposed on that motion. These constraints, which relate the particle coordinates, can
be described using the algebraic equations
(x)2 + (y)2 = (a)2 , z=0 (1.1)
The first equation relates the coordinates x and y, while the second equation implies that
the motion on the circle is planar. Using
√ the first equation, the coordinate y can be written
in terms of the coordinate x as y = ± (a)2 − (x)2 . Therefore, the particle position, or the
[ ]T
position of an arbitrary point on the curve r = x y z , can be written as
[ ]T [ √ ]T
r = x y z = x ± (a)2 − (x)2 0 (1.2)
This equation shows that points on the curve can be traced if the parameter (coordinate)
x is given, and the curve equation can be written in terms of one parameter, which is x in
this example. Given the parameter x, the location of an arbitrary point on the curve can be
completely determined.

a r
y
θ
x X

Figure 1.5 Circle geometry.

Curve Parameterization The curve parameter is not unique: that is, other parameters
can be used to develop the curve equation. For example, the angle 𝜃 shown in Figure 5 can
be used as the curve parameter. In this case, one must be able to write the curve Cartesian
coordinates x, y, and z in terms of the parameter 𝜃. It is clear from Figure 5 that the
coordinates x and y can be written, respectively, in terms of the parameter 𝜃 as x = a cos 𝜃
Introduction 7

and y = a sin 𝜃. Therefore, the curve equation can be written in terms of the parameter 𝜃 as
[ ]T [ ]T
r = x y z = a cos 𝜃 a sin 𝜃 0 (1.3)
Equations 2 and 3 clearly show that the same curve can have different parameterizations.
The different parameters, however, are related by algebraic equations, as previously
explained using the simple example considered in this section. Equation 2 or 3 is called the
curve equation in its parametric form. The equation (x)2 + (y)2 = (a)2 , which defines the
circle equation, is called the curve equation in its implicit form.

Arc Length It is clear from the simple analysis presented in this section that different
parameters can be used to define the parametric form of the curve equation. As discussed
earlier, these parameters are related, as evidenced by the fact that one can write the param-
eter x in terms of the parameter 𝜃 as x = a cos 𝜃, and the parameter 𝜃 can be written in terms
of the parameter x as 𝜃 = cos−1 (x/a). Similarly, a curve can be parameterized by its arc length
s. This can be easily demonstrated by writing 𝜃 = s/a. This equation can be substituted into
Eq. 3 to obtain
[ ]T [ ]T
r = x y z = a cos (s∕a) a sin (s∕a) 0 (1.4)
Therefore, given any space curve, the curve parametric equation can be written in terms of
one parameter as (Do Carmo, 1976; Goetz, 1970; Kreyszig, 1991)
[ ]T
r (t) = x (t) y (t) z (t) (1.5)
where t is the curve parameter. While a curve can be parameterized using any scalar vari-
able, the curve arc length that measures the distance from the curve starting point to an
arbitrary point on the curve is often used as the curve parameter. Using the parametric form
of the curve expressed in terms of its arc length, a spatial curve can be uniquely defined in
terms of geometric invariants called the curvature and torsion, which appear in the curve
Serret–Frenet equations presented in Chapter 2. A curve can also be defined in its implicit
form by eliminating the parameter to obtain a single equation expressed in terms of the
curve coordinates, as previously demonstrated by the circular curve example.

s2

∂r
∂s2 s1

∂r
∂s1

Figure 1.6 Surface geometry.


8 Mathematical Foundation of Railroad Vehicle Systems

Theory of Surfaces Unlike curves, surface equations are defined in terms of two param-
eters, because on a surface, one has the freedom to move in two independent directions, as
shown in Figure 6. For example, in the case of a rail surface, a wheel can roll and/or slide on
the surface longitudinally or laterally. If the particle considered previously in this section
is not constrained to move in the plane on the circular curve, the coordinate z is no longer
constant, and such a coordinate can vary. In this case, there is only one constraint equation
on the motion of the particle. This constraint equation is defined as (x)2 + (y)2 = (a)2 ,
and therefore the vector r can be written in terms of two independent coordinates x and
z as
[ ]T [ √ ]T
r (x, z) = x y z = x ± (a)2 − (x)2 z (1.6)

It is clear that this equation, which defines the cylindrical surface shown in Figure 7,
depends on the two independent surface parameters (coordinates) x and z. Once these
two parameters are known, the coordinates of an arbitrary point on the surface can be
determined. As in the case of curves, the surface parameterization is not unique. That is,
other parameters such as 𝜃 and z can be used. In this case, the surface parametric equation
can be written as
[ ]T [ ]T
r (𝜃, z) = x y z = a cos 𝜃 a sin 𝜃 z (1.7)

a
Y
X

Figure 1.7 Cylindrical geometry.

In general, the surface equation can be written in its parametric form in terms of two
independent parameters s1 and s2 as (Do Carmo, 1976; Goetz, 1970; Kreyszig, 1991)
( ) [ ( ) ( ) ( )]T
r s1 , s2 = x s1 , s2 y s1 , s2 z s1 , s2 (1.8)
Therefore, for a surface, one can define two independent tangent vectors 𝜕r/𝜕s1 and 𝜕r/𝜕s2
which define what is called a tangent plane. The geometric properties of a surface are
defined using the first and second fundamental forms of surfaces, which are presented in
Introduction 9

Chapter 2. The coefficients of these fundamental forms are used to define the principal
curvatures and principal directions that enter into formulating the wheel/rail contact force
equations.

Computational Approach for Geometric Representations To use the theory


of differential geometry in practical applications, polynomial approximations are used to
describe curves and surfaces. In railroad vehicle dynamics, using polynomials allows for
defining arbitrary rail and wheel profile geometries. In Chapter 2, a finite element (FE)
approach called the absolute nodal coordinate formulation (ANCF) is used to define the rail
surface geometry. Starting with the polynomial interpolation, the ANCF finite elements
are developed by replacing the polynomial coefficients with position and position gradient
coordinates. This allows for describing the position of arbitrary points on a continuum
using the equation r(x, y, z, t) = S(x, y, z)e(t), where S is a shape function matrix, t is time,
and e is the vector of position and position gradient coordinates. If ANCF finite elements
are used to describe the geometry of fixed rigid rail, one has r(x, y, z) = S(x, y, z)e. Using this
approach, which enables integrating geometry and analysis, allows for defining a surface
systematically by writing an algebraic equation in which one coordinate (parameter) can
be expressed in terms of the other two coordinates. For example, one can write z = f (x, y)
and use this functional relationship to define the rail surface equation as
[ ]T
r (x, y) = x y f (x, y) (1.9)
In this surface equation, which can be conveniently defined using ANCF finite elements,
only two parameters can be varied. Therefore, ANCF elements can be used to describe
the geometries of curves and surfaces in their most general forms based on polynomial
interpolations. The use of the ANCF position-gradient coordinates allows for conveniently
describing complex shapes as well as the deformations in the case of flexible rails. The
approach described in Chapter 2 also allows for using numerical or tabulated data to
describe the surface geometry. The fact that one method can be used to define the geometry
correctly and to accurately predict the deformation of the rail in the case of flexible rails
allows for the systematic integration of geometry and the analysis of complex railroad
vehicle systems, as discussed in more detail in the following section.

1.2 INTEGRATION OF GEOMETRY AND MECHANICS


The integration of geometry and mechanics represents the foundation for formulating the
railroad vehicle system nonlinear dynamic equations of motion. The dynamic behavior
and stability of the rail vehicle depend on the wheel/rail contact forces. These forces are
functions of the geometry of the wheel and rail surfaces, which can be described using
the techniques of differential geometry as well as computational geometric methods based
on polynomial interpolations, as discussed in the preceding section. The track geometry
also has a significant impact on rail-vehicle motion and stability. Track irregularities can
influence vehicle dynamics and be a source of derailments and serious accidents when the
vehicle negotiates curved and straight tracks. Therefore, the geometries of these irregulari-
ties need to be accurately represented in the simulation models in order to be able to predict
their effect on overall vehicle behavior and nonlinear dynamics.
10 Mathematical Foundation of Railroad Vehicle Systems

When the vehicle negotiates a curved track, the effect of centrifugal forces must be taken
into account. To avoid derailments as the result of high centrifugal forces, the geometry
of the track is altered by providing a track elevation that results in a lateral gravity-force
component that opposes and balances the centrifugal forces, as discussed in this chapter.
Accurate prediction of the effect of the centrifugal forces requires an accurate represen-
tation of the track geometry. Curved track sections can consist of curves with constant
curvatures, and spirals that have curvatures that vary along the track. In the case of spirals,
the radius of curvature is not constant, and consequently, the centrifugal force does not
remain constant. Therefore, in railroad vehicle dynamics, geometry, motion descriptions,
and force formulations are interrelated and cannot be separated.

General Displacement In the general case of unconstrained motion, the displacement


of a rigid body in space can be described using six independent coordinates. Three coordi-
nates define the global position of a point on the body, called the body reference point, and
three coordinates define the orientation of the body with respect to the global coordinate
system. The global position of the body reference point can be defined using three Cartesian
coordinates. The orientation coordinates, on the other hand, can be introduced using three
independent parameters that can represent angles or can be parameters that do not have an
obvious physical meaning. Therefore, in spatial analysis, the orientation parameters are not
unique, and different sets of parameters have been used in the literature and in developing
computational multibody system (MBS) algorithms.
To define the configuration of a component (body) i in a vehicle system, two coordinate
systems are first introduced, as shown in Figure 8. The first coordinate system is the global
XYZ coordinate system, which is assumed fixed in time, while the second coordinate system
X i Y i Z i is the body coordinate system, which is assumed to be rigidly attached to the body
reference point Oi . Using these two coordinate systems, the global position vector ri of an
arbitrary point on the rigid body i in the vehicle system can be written as
ri = Ri + ui (1.10)
[ i ]T
where R = Rx Ry Rz is the global position vector of the body reference point Oi , and
i i i
[ ]T
ui = uix uiy uiz defines the location of the arbitrary point with respect to the origin of
the body coordinate system X i Y i Z i in the global system: that is,
[ ]T
ui = uix uiy uiz = uix i + uiy j + uiz k (1.11)
In this equation, i, j, and k are, respectively, unit vectors along the global axes X, Y , and
Z. As discussed in Chapter 3, the vector ui can be written in terms of constant compo-
nents defined in the body coordinate system X i Y i Z i . This can be achieved by developing a
transformation matrix that defines the body orientation. The columns of the transforma-
tion matrix define orthogonal unit vectors along the axes of the body coordinate system.
While the body transformation matrix can be expressed in terms of three independent ori-
entation parameters such as Euler angles or any other sets of parameters, the elements
of the transformation matrix must assume the same numerical values regardless of the
orientation parameters used. These elements of the transformation matrix, as discussed
in Chapter 3, are the direction cosines of unit vectors along the axes of the body coordinate
system X i Y i Z i .
Introduction 11

Zi
Yi

Oi
ui

Xi Pi
Z
Y Ri

ri

Figure 1.8 Coordinate systems.

Angular Velocity and Orientation Parameters By differentiating Eq. 10 once and


twice with respect to time, the absolute velocity and acceleration vectors of the arbitrary
point on the body can be defined. The derivative of the transformation matrix with respect
to time can be used to define the angular velocity vector, as discussed in Chapter 3. In spa-
tial analysis, the angular velocities are not exact differentials, and therefore they are not the
time derivatives of orientation parameters. That is, the angular velocities cannot be directly
integrated to determine the orientation parameters. Nonetheless, the angular velocities can
always be written as linear functions of the time derivatives of the orientation parameters
using a velocity transformation matrix. This velocity transformation matrix plays a funda-
mental role in determining the generalized forces associated with the orientation parameters
since these orientation parameters serve as generalized coordinates and are not directly asso-
ciated with the Cartesian moments applied to the bodies, as will be discussed in Chapter 6.
The kinematic description that will be used in this book to develop the equations of
motion of the components of railroad vehicle systems is introduced in Chapter 3. It is shown
in Chapter 3 that the use of three parameters, such as Euler angles (Greenwood, 1988; Hus-
ton, 1990; Roberson and Schwertassek, 1988; Rosenberg, 1977), to define the body orienta-
tion in space leads to kinematic singularities. Such singularities, however, can be avoided
by using the four Euler parameters at the expense of adding an algebraic constraint equation
that relates the four Euler parameters. Euler parameters, which are becoming more popular
in developing general MBS algorithms, have many identities that can be used to simplify
the kinematic and dynamic equations of the railroad vehicle system.

Euler Angles and Track Geometry In addition to using Euler angles to describe time-
dependent motion by defining the orientation of bodies in space, these angles have also been
used in railroad vehicle dynamics to define the geometry of the track based on given simple
industry inputs. For the most part, track is constructed using three main segments: tangent
(straight), curve, and spiral, as shown in Figure 9. The tangent segment has zero curvature,
12 Mathematical Foundation of Railroad Vehicle Systems

the curve segment has constant curvature, and the spiral segment, used to connect two seg-
ments with different curvature values, has a curvature that varies linearly along the track
to ensure a smooth transition between the two segments connected by the spiral. The track
geometry is often described using three inputs at points along the track at which the geome-
try changes. These three inputs are the horizontal curvature, superelevation, and grade, and
they can be expressed in terms of three Euler angles that are used to construct the track and
rail space curves. To this end, Euler angles are converted to field variables and used system-
atically to construct a curve with well-defined geometry based on the given simple track
inputs. Unlike the three Euler angles used to describe the time-dependent motion of an

(a)

(b)

Figure 1.9 Track segments. Sources: (a) Dinodia Photos/Alamy Stock Photo. (b) Jens
Teichmann/Adobe Stock.
Introduction 13

unconstrained body in space, the three Euler angles used to describe the geometry of a curve
are written in terms of one parameter that can be the arc length. Therefore, when Euler
angles are used to describe curve geometry, these angles are converted to field variables
expressed in terms of the curve arc length to ensure a unique definition of the geometry.
Therefore, it is important to recognize that Euler angles are used in this book for two
fundamentally different purposes: (i) as motion-generalized coordinates to describe rigid
body kinematics in space; and (ii) as geometric field variables to uniquely define the geom-
etry of the track and rail space curves. The analysis presented in Chapter 3 is used as the
basis for a computer procedure for developing the track geometry data required for non-
linear dynamic simulations of railroad vehicle systems. The data can be generated before
the dynamic simulation at a preprocessing stage in a track preprocessor computer program,
as will be explained in Chapter 4. The track preprocessor output file normally has data for
three different space curves: the track centerline space curve, the right rail space curve, and
the left rail space curve, as shown in Figure 10, in which RH is the radius of curvature of the
track centerline curve. These three curves can have different geometries. The right and left
rail space curves are used in formulating the wheel/rail contact conditions, while the track
space curve is used in the definition of the distance traveled and in the motion description
of the coordinate systems of the vehicle components.
ΔR

Br
B
Bl
/2
ΔR

Track centerline
– H
R

Left rail space curve


Δψ

R R Right rail space curve


H
H +
Δ
R/
2 Al
A

Ar

Figure 1.10 Track space curves.


14 Mathematical Foundation of Railroad Vehicle Systems

1.3 HUNTING OSCILLATIONS

A simple analysis based on pure kinematics and geometric considerations can be used to
shed light on the dynamics of railroad vehicle systems without consideration of the forces.
In most railroad vehicle systems, a wheelset consists of two wheels connected by a stiff
axle, as shown in Figure 11. The wheels are assumed to have conical profiles with the
larger diameter close to the flange in order to achieve self-centering and minimize flange
contact (Karnopp, 2004). Lateral wheelset oscillations with respect to the track centerline
are referred to as hunting. During hunting oscillations, there is a relationship between
the wheelset lateral displacement and the yaw angle, which represents the rotation of the
wheelset about an axis normal to the track structure. In this section, a simple analysis based
on pure geometry is used to demonstrate the relationship between the lateral displacement
and yaw angle of the wheelset when it exhibits hunting oscillations. Such oscillations play a
fundamental role in the stability of railroad vehicle systems. As will be demonstrated in this
section, the hunting frequency is a function of the forward velocity of the wheelset as well
as some other geometric parameters, including the wheel conicity, nominal rolling radius,
and distance between the two rails.

y γ

Rl Rr

Figure 1.11 Railroad wheelset.

To provide an example of this simple geometric analysis, the wheelset shown in Figure 11
is considered. As shown in the figure, the wheelset conicity is denoted as 𝛾, which defines
the slope of the wheel profile curve. The lateral displacement of the wheelset center of mass
is denoted as y. Before displacement, the wheelset is assumed to be centered and the dis-
placement y is assumed to be zero: that is, y = 0. At this initial configuration, the radii of the
two wheels at the points of contact with the rails are equal and denoted as Ro . As a result
of disturbances that can be attributed to initial conditions or rail irregularities, the rolling
radii of the two wheels will deviate from Ro as the wheelset starts to move forward. These
rolling radii are denoted as Rr and Rl for the right and left wheels, respectively. As a result
of a lateral displacement y, the rolling radii of the two wheels change, and such a change
in the rolling radii is defined by ΔR = y𝛾. It follows from simple geometry that Rr = Ro − y𝛾
and Rl = Ro + y𝛾. If the wheelset is assumed to rotate with a constant angular velocity 𝜔,
Introduction 15

Vl

Y
G
X y(x) dy
Ψ =
dx

Vr

Figure 1.12 Hunting oscillations.

the forward velocities of the right and left wheels can be written, respectively, as
( )}
Vr = 𝜔Rr = 𝜔 Ro − y𝛾
( ) (1.12)
Vl = 𝜔Rl = 𝜔 Ro + y𝛾
This equation shows that, during hunting oscillations, the two wheels have different for-
ward velocities, and this gives rise to a yaw angle 𝜓, as shown in Figure 12. Nonetheless,
the forward velocity of the wheelset center of mass remains constant and is always defined
by the following equation:
( )
V = Vr + Vl ∕2 = 𝜔Ro (1.13)

Using the small oscillation assumption, one can write tan 𝜓 = dy/dx ≈ 𝜓. Because one can
write, using Eq. 12, V r − V l = − 2y𝜔𝛾, it follows that
( ) }
𝜓̇ = Vr − Vl ∕G = −2y𝜔𝛾∕G,
(1.14)
̇
𝜓̈ = −2y𝜔𝛾∕G
where G is the distance between the two rails. Furthermore, one can write, using the
assumption of constant wheelset forward velocity V,

dy dy dx ⎫
ẏ = = = 𝜓 V = 𝜓 Ro 𝜔,⎪
dt dx dt ⎬ (1.15)
ÿ = 𝜓̇ V = 𝜓̇ Ro 𝜔 ⎪

Substituting the first equation of Eq. 14 in the second equation of Eq. 15; and substituting the
first equation of Eq. 15 in the second equation of Eq. 14, one obtains, respectively, the follow-
ing second-order homogeneous ordinary differential equations for the lateral displacement
and yaw angle, respectively:
( )2 ( )2
ÿ + 𝜔h y = 0, 𝜓̈ + 𝜔h 𝜓 = 0 (1.16)

where

𝜔h = 𝜔 2Ro 𝛾∕G (1.17)
16 Mathematical Foundation of Railroad Vehicle Systems

is the hunting frequency that can be defined only in the case of positive conicity. In the
case of positive conicity, solutions of the preceding equations can be assumed in the forms
y = Ay sin(𝜔h t + 𝜙y ) and 𝜓 = A𝜓 sin(𝜔h t + 𝜙𝜓 ), where Ay and A𝜓 are the amplitudes, and 𝜙y
and 𝜙𝜓 are phase angles that can be determined using the initial conditions. These solutions
for the lateral displacement and yaw angle show that the frequencies of oscillation√ of the
lateral and angular yaw displacements are the same and are defined by 𝜔h = 𝜔 2Ro 𝛾∕G.
Furthermore, by using these solutions, the first equation of Eq. 15, ẏ = 𝜓 Ro 𝜔, can be used
to prove that the amplitudes of the lateral displacement and yaw angles are related by the
equation Ay = Ro 𝜔A𝜓 /𝜔h = VA𝜓 /𝜔h , and there is a phase angle 𝜋/2 between the lateral
displacement y and the yaw angle 𝜓: that is, 𝜙y − 𝜙𝜓 = 𝜋/2. This difference in the phase
angle and the relationship ẏ = 𝜓 Ro 𝜔 show that the maximum and minimum values of the
yaw angle 𝜓 occur when the lateral displacement y is zero, and the maximum and minimum
y occur when 𝜓 = 0. The hunting oscillations in the case of positive conicity are shown in
Figure 13a.

y(x) y(x) y(x)

X X X

Y Y Y
(a) (b) (c)

Figure 1.13 Conicity effect.

If, on the other hand, the conicity is equal to zero, 𝛾 = 0, which is the case of a cylindrical
wheel, one has, from Eq. 15, ÿ = 0 and 𝜓̈ = 0. Integrating these two equations with respect
to time shows that the solution is represented by straight lines and the motion is not oscil-
latory. If the initial conditions are different from zero, the solution will increase with time,
leading to an unstable solution. In this case of a cylindrical wheel, the wheelset does not
tend to self-center, as shown in Figure 13b.
Introduction 17

In the case of negative conicity, 𝛾 < 0, the coefficient of y and 𝜓 in Eq. 16 is negative, and
the characteristic equations of the two equations have real roots that define an exponentially
unstable solution, as shown in Figure 13c. The simple analysis presented in this section
explains the effect of conicity on wheelset stability.
The hunting frequency can be written in terms of the forward velocity of the wheelset
using the equation V = 𝜔Ro , which upon substitution into Eq. 17 leads to

2𝛾
𝜔h = V (1.18)
Ro G
This equation defines the period of hunting oscillations T h as

2𝜋 2𝜋 Ro G
Th = = (1.19)
𝜔h V 2𝛾
The preceding two equations are called Klingel’s formulas (Klingel 1883). Klingel’s formula
in Eq. 18 shows that the frequency of hunting oscillations increases as the wheelset for-
ward velocity or the conicity increases, and the hunting frequency decreases as the nomi-
nal rolling radius or the distance between the two rails increases. While Klingel’s formula
is obtained in this section based on pure kinematic and geometric considerations with-
out regard to the forces acting on the wheelset, computer simulations based on nonlinear
dynamic formulations that account for all the forces acting on the system have demon-
strated the accuracy of the hunting frequency predicted by Klingel’s formula in Eq. 18.

1.4 WHEEL AND TRACK GEOMETRIES

The wheel and rail surface geometries enter into formulating the normal and tangential
contact forces that influence the nonlinear dynamics and stability of railroad vehicle sys-
tems as well as the integrity of the track structure. Therefore, an accurate description of
the geometry is necessary for developing credible virtual prototyping computer models for
railroad vehicle systems. The wheel and rail surface geometries can be described using the
theories of curves and surfaces that are discussed in more detail in Chapter 2. For example,
to determine the dimensions of the wheel/rail contact area, the principal curvatures and
principal directions of the wheel and rail surfaces in the contact region must be evaluated.
While the wheel geometry can be described using a surface of revolution, the rail geom-
etry can be defined by extrusion of the rail profile in the rail longitudinal direction. More
complex surface geometries can be described in the mathematical models using numerical
approximation methods in order to capture details that cannot be captured using analytical
techniques that are more suited for simple or idealized geometries.

Wheel Surface Geometry A railroad wheel normally consists of two sections: the
tread and the flange. The flange is used to limit the lateral motion of the wheel in order
to avoid derailments. The tread is designed to have conicity that provides the tendency of
self-centering and decreases the possibility of motion instability. Very high conicity can
also lead to higher-frequency hunting oscillations, as discussed in the preceding section.
The conicity 𝛾, which depends on the particular combination of the wheel and rail used,
18 Mathematical Foundation of Railroad Vehicle Systems

normally varies from 1/20 to 1/40, with some values more popular in particular regions and
countries. The conicity can also have an effect on the number of contact points between
the wheel and rail; the frequency of the occurrence of two-point contacts between the
wheel and rail depends on the conicity used.
The geometry of the unworn wheel can be described as a surface of revolution by rotating
the profile curve about the wheel axis, as shown in Figure 14. While the profile can assume
any shape, conical wheels are often used in order to improve vehicle stability and avoid
derailments, as explained in the preceding section. Different profiles with different conicity
values are used, depending on the type of vehicle, speed of operation, and loading condi-
tions. The functions that define the profiles are not simple straight-line functions, and in
most practical applications, a numerical description of the profile function is required for
accurate computer modeling and virtual prototyping.

sw1 Zwp
Ywp sw2

gwp(sw1 )
Xwp

(a) (b)

Figure 1.14 Wheel geometry.

The wheel surface can be described using two parameters, as discussed in the preceding
section. Because the unworn wheel surface is a surface of revolution, it is convenient to use
the parameterization sw1 = yws and sw2 = 𝜃sw , as shown in Figure 14, where subscript w refers
to the wheel. To develop a mathematical definition of the wheel surface, a profile frame
X wp Y wp Z wp is introduced, as shown in Figure 14, for the convenience of defining the profile
curve. The angular surface parameter sw2 is measured from the Z wp axis. The position of the
origin of the profile frame with respect to the wheel coordinate system X w Y w Z w is defined
wp wp wp wp
by the Cartesian coordinates xo , yo , and zo , which form the elements of the vector: R =
[ wp wp wp ]T
xo yo zo . The Y w axis is assumed to coincide with the wheel axis of rotation. A
profile function gwp can be used to define the wheel profile curve in the wheel profile frame.
In the case of unworn wheels, the profile function gwp does not depend on the wheel angular
( )
surface parameter sw2 = 𝜃sw , and in this special case, one has gwp = gwp sw1 . In this case of
Introduction 19

unworn wheels, the coordinates of an arbitrary point on the wheel surface can be defined
mathematically in the selected wheel coordinate system X w Y w Z w in terms of the surface
parameters sw1 and sw2 as
wp ( )
⎡xo + gwp sw1 sin sw2 ⎤
( ) wp ⎢ wp ⎥
uw sw1 , sw2 = R + uwp = ⎢ yo + sw1 ⎥ (1.20)
⎢ wp ( w) w⎥
⎣ zo − g wp s1 cos s2 ⎦
wp [ wp wp wp ]T
where, as previously defined, R = xo yo zo is the vector that defines the origin
of the wheel profile frame X wp Y wp Z wp with respect to the coordinate system X w Y w Z w of
[ ( ) ( ) ]T
the wheel or wheelset, and uwp = gwp sw1 sin sw2 sw1 −gwp sw1 cos sw2 is the vector
that defines the location of the arbitrary point in the profile frame. In the case of a single
wp wp wp
wheel, the coordinates xo , yo , and zo define the position of the origin of the profile frame
X wp Y wp Z wp in the wheel coordinate system X w Y w Z w . In the case of a rigid wheelset that has
two wheels rigidly connected by an axle, the coordinate systems of the rigidly connected
right and left wheels can be assumed the same and have origins located at the axle center
wp wp wp
point. Using the coordinates xo , yo , and zo of the origin of the wheel profile frame allows
the systematic generalization of the geometric description used in this book to the case of
deformable wheel axles or independent non-rigidly connected wheels. Furthermore, this
description allows for using a numerical spline function representation of the profile func-
tion gwp , and therefore, measured wheel profile data can be used. In the case of worn wheels,
( )
the profile function depends on both parameters sw1 and sw2 : that is, gwp = gwp sw1 , sw2 . In
Chapter 4, a more detailed description of the wheel surface geometry is presented.

Track Geometry Developing an accurate description of the track geometry is one of the
basic steps in formulating the rail vehicle nonlinear dynamic equations of motion and in the
numerical solution of these equations. As previously discussed in this chapter and shown
in Figure 9, for the most part, the track is constructed using three segment types with differ-
ent geometries: tangent, curve, and spiral segments. A tangent segment is a straight section
of track with zero curvature, which corresponds to a radius of curvature equal to infinity.
A curve segment is a circular section of track that has a constant radius of curvature and
constant curvature. To connect tangent and curve segments that have different curvatures,
a spiral segment is used to ensure a smooth transition. When a spiral segment is used to con-
nect tangent and curve sections of the track, the geometry of the spiral segment is designed
such that the spiral has zero curvature at the end connected to the tangent segment and
the value of the curve curvature at the end connected to the curve segment. This spiral
design allows for smoothly varying the curvature and ensuring smooth operation of the rail
vehicle during the transition between the tangent and curve sections and vice versa. These
simple track segments (straight, curve, and spiral) represent the basic geometric elements
for the construction of a track, but at intersections, the track can have a complex structure,
as shown in Figure 9. This complex structure, which can include switches (turnouts) and
guardrails, can still be constructed using basic rail segments.
The surface of each rail of the track can also be described using two surface parameters,
which are selected in this book to be the longitudinal surface parameter sr1 and the lateral
surface parameter sr2 , as shown in Figure 15, where superscript r refers to the rail. Using
20 Mathematical Foundation of Railroad Vehicle Systems

sr2

f(sr2)

sr1

Z
X
Y

Figure 1.15 Rail geometry.

these surface parameters, the location of an arbitrary point on the rail surface can be defined
in a selected rail coordinate system as explained in Chapter 4 as
( ) rp rp
ur sr1 , sr2 = R + Arp u (1.21)
rp rp ( ) [ rp ( ) rp ( ) rp ( )]T
where R = R sr1 = xo sr1 yo sr1 zo sr1 is the vector that defines the origin of
( )
the rail profile frame with respect to a rail coordinate system X r Y r Z r , Arp = Arp sr1 is a
matrix that defines the orientation of the profile frame X rp Y rp Z rp with respect to the rail
rp [ ( )]T
coordinate system X r Y r Z r , u = 0 sr2 grp sr2 is the vector that defines the location
of the arbitrary point in the profile frame X rp Y rp Z rp , and grp is a function that defines the
geometry of the rail profile. In Chapter 4, it is shown how the transformation matrix Arp =
( )
Arp sr1 is defined using the techniques of differential geometry presented in Chapter 2.

Track Design Important definitions and terminologies are used in railroad vehicle
dynamics, particularly in the layout of the track. These definitions are also important in
developing computational algorithms in which numerical representation of the track is
necessary. The gage G shown in Figure 16 is the lateral distance between two points on
the heads of the right and left rails. In North America, these two points are located at

Track plane
w
G
ϕ h

Horizontal plane

Figure 1.16 Gage and superelevation.


Introduction 21

a distance 14 mm (5/8 in.) from the top of the railhead. The standard gage value used
in North America varies from 56 to 57.25 in. The superelevation h, shown in Figure 16,
is defined as the vertical distance between the right and left rails. This vertical distance
defines the bank angle 𝜙 shown in the figure. In the case of a curved section of the track,
the curvature is different from zero. As shown in Figure 17, the curvature is defined using
the value of the angle 𝜓 encompassed by a 100 -length chord defined by the end points P1

and P2 . The chord is assumed to have a constant radius RH in the horizontal plan, as shown
in Figure 17. The grade is defined as the ratio (percentage) between the vertical elevation
and the longitudinal distance. The cant angle is defined as the rotation of each rail about
its longitudinal axis, as shown in Figure 18.

100

P1

RH

P2
ψ

Figure 1.17 Curvature.

Track irregularities can be the source of derailments and serious accidents, and therefore,
their effect must be evaluated. As discussed in Chapter 4, some standard track deviation
functions can be used in computational models to test the vehicle stability. Measured track
data are also often used in the computer simulations of railroad vehicle systems. Track devi-
ations can be classified as profile or alignment. The profile is defined as the vertical deviation
of the rail space curve, while the alignment is defined as the lateral deviation of the rail space
curve, as shown in Figure 19. In Chapter 4, examples of different standard track deviations
are provided.

Cant angle

Figure 1.18 Cant angle.


22 Mathematical Foundation of Railroad Vehicle Systems

Alignment

Chord length
Profile (Vertical deviation)
Z
X

X
Chord length

Y
(a) (b)

Figure 1.19 Track deviations.

Analytical and Computational Methods Chapter 4 discusses the geometries of the


wheel and rail surfaces, which are fundamental in defining the kinematic and dynamic
equations of the rail vehicle. The wheel and rail surface equations are defined in terms of
two independent surface parameters, as previously mentioned. These surface parameters,
which define the surface equations in their parametric form, can be used to determine
the locations of arbitrary points on the wheel and rail surfaces. The parametric surface
equations are necessary for developing the wheel/rail contact conditions, which are used to
determine online the wheel/rail contact points. To describe the rail geometry, two different
approaches are discussed in Chapter 4: the semi-analytical approach and the ANCF inter-
polation approach. The semi-analytical approach has two main disadvantages. The first is
the need to evaluate the derivatives of angles with respect to the rail longitudinal surface
parameter in order to determine the tangent, normal, and curvature vectors. The second is
the low order of interpolation used to determine the position coordinates of arbitrary points
on the rail with limited point data. These data, obtained as the output of a track preprocessor
computer program, include three position coordinates of and three angles at discrete nodal
points on the track. With these limited data at each point, the semi-analytical approach
cannot be used with a higher order of interpolation. The ANCF interpolation approach is
preferred because it does not have these disadvantages. Furthermore, the ANCF approach
does not require differentiation of angles and allows for higher-order interpolation for the
position coordinates of arbitrary points on the rail space curve.

1.5 CENTRIFUGAL FORCES AND BALANCE SPEED


When a rail vehicle negotiates a curve or a spiral section, the forces exerted on the vehicle
can be significantly different from the forces that arise when the vehicle negotiates a straight
segment. During curve and spiral negotiations, the centrifugal forces must be taken into
account in order to ensure safe operation of the rail vehicle. In order to avoid derailments,
the magnitude of the centrifugal force is used to enforce a limit on the vehicle speed; this
Introduction 23

limit is referred to as the balance speed. In the case of a curve which has a constant radius
of curvature, the variations in the centrifugal forces acting on the vehicle are not as signifi-
cant as in the case of a spiral which has variable curvature. When constructing a track, the
geometry of the spiral sections is often designed to have a linearly varying curvature.
To balance the effect of the centrifugal force F ce when the vehicle negotiates a curve, the
curve must be designed to have a super-elevation h in order to produce a lateral gravity
force component that balances the centrifugal force. The centrifugal force has the effect
of pushing the vehicle away from the center of the curve in the direction of the high rail,
while the lateral gravity force component resulting from the super-elevation has the effect of
pushing the vehicle in the opposite direction towards the center of the curve. The direction
of the centrifugal force depends on the direction of the normal to the curve traced by the
center of mass of the vehicle. In order to demonstrate the interrelationship between the
geometry and mechanics concepts, two different scenarios are considered in this section.
In the first scenario, it is assumed that the vehicle center of mass traces a circular curve that
lies in a plane parallel to the horizontal plane. In this case, the centrifugal force acting on the
vehicle lies in the horizontal plane regardless of the amount of the track superelevation. In
the second scenario, it is assumed that the vehicle experiences lateral motion, and therefore,
the motion-trajectory curve is not in general circular and does not lie in a plane parallel
to the horizontal plane. In this latter case, the vehicle center of mass can move vertically.
Therefore, distinction must be made between the track geometry and the geometry of the
curve the vehicle traces during its motion.

Circular Curve Equations If the curve is assumed to have a radius of curvature R, the
arc length s that encompasses an angle 𝜃 can be written as s = R𝜃, which implies that for
a constant radius of curvature R, the forward velocity V of a vehicle negotiating the curve
can be written as V = ṡ = R𝜃.̇ If the vehicle, for simplicity, is assumed to be represented
by a point mass, the position of the vehicle in the curve plane with respect to a coordi-
[ ]T
nate system located at the center of the curve can be written as r = R cos 𝜃 sin 𝜃 0 ,
where the first and second elements in this vector represent, respectively, the lateral and
forward components with the forward component in a direction tangent to the curve. Dif-
ferentiating this position vector r with respect to time, the velocity of the vehicle is defined
[ ]
as ṙ = 𝜃Ṙ − sin 𝜃 cos 𝜃 0 T , which shows that the magnitude of the velocity vector is
̇ Differentiating the velocity with respect to time, one obtains the absolute accel-
V = 𝜃R.
eration vector defined as
[ ] ( ) [ ]
̈ − sin 𝜃 cos 𝜃 0 T − 𝜃̇ 2 R cos 𝜃 sin 𝜃 0 T
r̈ = 𝜃R (1.22)

̇
If the vehicle is assumed to travel on the curve with a constant forward velocity V = 𝜃R,
𝜃̇ is constant, and 𝜃̈ = 0. Therefore, the preceding acceleration equation reduces to r̈ =
( )2 [ ]T
− 𝜃̇ R cos 𝜃 sin 𝜃 0 . Since 𝜃 = s/R and V = s,̇ where s is the arc length of the curve,
one has in the case of constant forward velocity V

(V)2 [ ]T
r̈ = − cos 𝜃 sin 𝜃 0 (1.23)
R
This equation is used to determine the centrifugal inertia force.
24 Mathematical Foundation of Railroad Vehicle Systems

Horizontal-Plane Curve During curve negotiation, if the vehicle is not allowed to


move vertically, the vehicle center of mass traces a circular arc that lies in a plane parallel
to the horizontal plane. The normal to this curve is in the horizontal direction, and
therefore, the centrifugal force is in the direction shown in Figure 20 regardless of the
amount of the superelevation defined by the bank angle 𝜙t . A simple force balance shows
that (mV 2 /R) cos 𝜙t = mg sin 𝜙t . This equation can be used to define the balance speed as

V = Rg tan 𝜙t . If the right and left rails are canted to have a gage value similar to the gage
value of the track before the super-elevation, one can show that the balance speed obtained
in this case does not differ significantly from the actual balance speed that is determined by
including the effect of the cant. Because of the hunting oscillations, rolls, and suspension
system, the assumption of zero vertical motion can be violated. Nonetheless, highway and
railroad super-elevations are designed to limit the vertical motion of the vehicle during
curve negotiations by proper vehicle steering or using flanged wheels as in the case of
railroad vehicle systems.

Figure 1.20 Centrifugal force in the case of a horizontal-plane curve.

General Motion-Trajectory Curve In the motion scenario considered above, the cen-
ter of mass of the vehicle is assumed to negotiate a curve that lies in a plane parallel to the
horizontal plane. In this case, the vector normal to the curve that defines the direction of the
centrifugal force mV 2 /R has no vertical component. In the case of railroad vehicle systems,
the vehicle components can move laterally and vertically, and therefore, the assumption of
a horizontal curve cannot be, in general, satisfied. In the case of more general motion sce-
nario, the vector normal to the motion-trajectory curve has non-zero vertical component.
Using the more general definition of the absolute acceleration r̈ , as will be explained in this
section, one can show that the centrifugal force has the direction shown in Figure 21 and
has magnitude m |̈r| = mV 2 ∕R, where m is the mass of the vehicle. In this figure, 𝜙t is the
angle that defines the superelevation of the track and 𝜙 is the angle that defines the eleva-
tion of the osculating plane of the motion-trajectory curve. The osculating plane is the plane
that contains the vectors tangent and normal to the motion-trajectory curve. In the case of
the horizontal circular curve previously discussed in this section, 𝜙 = 0. The centrifugal
Introduction 25

Figure 1.21 Centrifugal force in the case of a general motion-trajectory curve.

inertia force in the more general case in which 𝜙 ≠ 0 still tends to push the mass away from
the center of the curve which can have curvature and radius of curvature that vary with the
arc length. This centrifugal force can still be balanced by a component of the gravity force.
Using Figure 21 and a simple force balance, one has
mV 2 ∕R = mg sin 𝜙 (1.24)
where g is the gravity constant, and 𝜙 is the angle that defines the direction of the normal
to the curve. Because the angle 𝜙 is assumed to be small, one can write sin 𝜙 ≈ tan 𝜙 = h/G,
where G is the track gage. Using this approximation, Eq. 23 can be used to define an approx-
imation of the balance speed as
√ √ √
V = gR sin 𝜙 ≈ gRh∕G = gR tan 𝜙t (1.25)
It is clear from this equation that the exact definition of the balance speed is derived using
the angle 𝜙 and not the track superelevation angle 𝜙t . In practice, the difference between
the two angles is small. Because the motion-trajectory curve is not a priori known, the use
of the approximate value of the balance speed is justified.

Practical Considerations and Geometry If the vehicle is negotiating a curve with


a velocity higher or lower than the balance speed, the vehicle is said to have cant defi-
ciency or cant excess, respectively. Cant excess or cant deficiency is defined as the amount
of super-elevation need to be reduced or increased, respectively, such that the given vehicle
speed is equal to the balance speed. The super-elevation is normally kept below 7 inches,
and therefore, the bank angle 𝜙t is small in most practical applications. While the two cases
of general and horizontal-plane curves discussed in this section give results for the balance
speed that do not differ significantly, it is important to recognize that because of the lateral
oscillations of the railroad vehicle, the conditions of a horizontal-curve cannot be precisely
met in practice, and the motion trajectory of the center of mass cannot be described by an
exact circular curve. If the actual motion trajectory is defined by a curve r(s), where s is the
arc length, a mass m tracing this curve has a velocity ṙ (s) = rs s,̇ where rs = 𝜕r/𝜕s is the unit
tangent to the curve. The absolute acceleration of the mass is defined by the vector r̈ (s) =
̇ 2 . As will be discussed in Chapter 2, the curvature vector rss = 𝜕 2 r/𝜕s2 is along
rs s̈ + rss (s)
26 Mathematical Foundation of Railroad Vehicle Systems

the unit vector n normal to the curve, which is defined from the equation rss = 𝜅n, where
𝜅 = 1/R is the curvature of the curve and R is the radius of curvature. Therefore, the accel-
( 2 )
eration vector can be written as r̈ (s) = s̈ rs + (s)
̇ ∕R n, which shows that the centrifugal
acceleration (s) ̇ 2 ∕R = (V)2 ∕R is along the normal to the curve. Therefore, if the trajectory of
motion does not lie in the horizontal plane, the direction of the inertia force m(V)2 /R will
be the same as the direction of the unit vector normal to the curve. Understanding these
basic differential-geometry equations and concepts is important in properly defining the
direction of the centrifugal force. For example, regardless of how the high rail is elevated,
if a wheelset with flanged wheels is placed on a super-elevated track plane which has two
rails to support the two wheels, the center of the wheelset will follow the centerline curve of
the track, but not precisely because of any lateral motion. Therefore, the motion trajectory
of the wheelset deviates from the centerline curve of the track due to the lateral motion.
The motion trajectory of the center of the wheelset is represented by a curve which defines
the direction of the centrifugal force and lies in the osculating plane of the curve.
In order to provide an example for the difference between the balance speeds evaluated
using the exact and approximate definitions, it is important to recognize that 𝜙 ≤ 𝜙t . That
is, the maximum value of 𝜙 cannot exceed 𝜙t . For example, in the case of a bank angle
𝜙t = √6.370∘ , the maximum √ balance speed obtained using the exact definition cannot exceed
V = Rg sin 𝜙 = 0.3333 Rg, while using the assumption of a horizontal-plane curve leads
√ √
to a balance speed V = Rg tan 𝜙t = 0.3344 Rg. That is, the error is less than 0.11%, which
is not significant, as previously mentioned.

1.6 CONTACT FORMULATIONS


Another basic step in the study of railroad vehicle system dynamics and stability is formulat-
ing the dynamic interaction forces between the wheel and rail. Accurate prediction of these
forces is necessary in order to obtain credible results that shed light on the system dynamic
behavior. When the wheel is pressed against the rail, a contact region, referred to as the con-
tact area or contact patch, is formed. The shape of the region of contact between two solids
depends on many factors that include the material properties, the contact pressure, the
solid geometries in the contact area, etc. If the contact region covers a finite area that can-
not be approximated by a point or a line, one has the case of conformal contact. In the case of
wheel/rail contact, the dimensions of the contact area are small compared to the dimensions
of the wheel and rail, and therefore, a localized concentrated contact is often assumed when
the wheel/rail contact forces are evaluated. This assumption of non-conformal contact can
be justified because of the shapes of the wheel and rail surfaces in the contact area.

Creep Forces When two solids come into contact and are subjected to external pressure,
some points on the contact surfaces may slip while other points may stick, and therefore, the
contact region consists of slip and adhesion areas. The small relative slip and spin between
the two solids can be the result of the difference between the deformations in the contact
region, and they lead to creep forces and spin moments, respectively.
The wheel/rail contact forces are the result of a combination of relative sliding and rolling.
As the vehicle attains a certain speed, and because of the effect of friction forces, the motion
Introduction 27

of the wheel with respect to the rail becomes predominantly rolling with small amount of
slipping; this gives rise to tangential creep forces as well as spin moments that have a sig-
nificant effect on the vehicle dynamics and stability. In the case of traction and braking,
for example, the relative velocity between the wheel and rail increases, causing significant
sliding, a case known as full saturation in which the tangential forces can be approximated
using the Coulomb friction law. Below a certain relative velocity value, slipping as the result
of the creep phenomenon produces creep forces that can be expressed in terms of normal-
ized velocities called creepages. The creep phenomenon, which is due to the elasticity of the
two solids in the contact area, is the source of tangential creep forces that are functions of
the creepages. Different linear and nonlinear wheel/rail tangential contact force and spin
moment models expressed in terms of velocity creepages are used in the railroad vehicle
system literature, as discussed in Chapter 5.

Contact Formulations The accuracy of normal and tangential contact force calcula-
tions depends on the accuracy of predicting the locations of the wheel/rail contact points.
Predicting the locations of the contact points online is necessary for the generality of the
wheel/rail dynamic model and for accurately formulating the wheel/rail interaction forces
and moments. Fundamentally different approaches are used in formulating the wheel/rail
dynamic interaction. In some of these approaches, referred to as constraint contact formu-
lations (CCFs), the wheel is assumed to remain in contact with the rail: that is, wheel/rail
separation is not allowed. In other formulations, referred to in this book as elastic contact for-
mulations (ECFs), wheel/rail separation is allowed. Both formulations can be used to deter-
mine the normal contact force, which is the wheel/rail interaction force in the direction
normal to the surfaces of contact. To develop efficient computational algorithms for predict-
ing wheel/rail contact forces, it is often assumed that wheel/rail contact is non-conformal:
that is, the contact is assumed to cover a very small region such that the use of a point con-
tact can be justified. Once the normal contact force is determined, tangential creep forces
can be computed using linear or nonlinear models, as discussed in Chapter 5.
The CCF and ECF approaches discussed in Chapter 5 lead to different models with dif-
ferent numbers of degrees of freedom. Because the CCF approach uses contact-constraint
equations that must be satisfied at position, velocity, and acceleration levels, the CCF
solution procedure is more complex than the ECF solution procedure in which wheel/rail
contact force is represented using a compliant force model. The fact that assumptions of
non-conformal contact are used in both approaches (CCF and ECF) does not imply that
conformal contact is not encountered in railroad vehicle system dynamics. Developing an
efficient and accurate method for modeling conformal contact in railroad vehicle system
applications remains a challenging problem, and it has been the subject of several research
investigations.

Hertz Theory As previously mentioned, in the case of non-conformal contact, the


wheel/rail contact area is assumed small in comparison with the wheel and rail dimen-
sions. When the constraint contact formulation is used, the normal contact force is
determined as a reaction force. In the elastic contact formulation, on the other hand,
normal contact forces are determined using a compliant force model with assumed stiff-
ness and damping coefficients. Experimental observations have shown that the wheel/rail
28 Mathematical Foundation of Railroad Vehicle Systems

contact area can be approximated using an elliptical shape. The dimensions of the contact
ellipse, which enter into formulating wheel/rail tangential creep forces, can be determined
using Hertz contact theory (Hertz 1882). The contact ellipse dimensions, wheel and rail
material properties, creepages, and normal contact force predicted using the constraint or
elastic contact formulation can be used to formulate the tangential creep force and spin
moment. In Hertz contact theory, the geometry of the two surfaces in contact is used to
determine the principal curvatures, which are used to determine the dimensions of the
contact ellipse. Chapter 5 discusses Hertz contact theory, wheel/rail contact formulations,
and creep force models.

Maglev Trains While most of Chapter 5 is devoted to formulating the wheel/rail contact
problem, discussions of the forces used in magnetically levitated (maglev) trains are also
presented. In the case of maglev trains, in which there is no contact between the vehicle
and the guideway, the vehicle is lifted using magnetic forces. This allows for a significant
increase in train speeds because it eliminates the limitations that are the result of wheel/rail
and pantograph/catenary contacts.
Maglev trains are currently being used, for the most part, for short-distance transporta-
tion. Many issues need to be addressed before using this technology for long-distance
transportation. It is also not clear whether this technology will be useful for freight trains
that consist of a very large number of cars.

1.7 COMPUTATIONAL MBS APPROACHES

Railroad vehicles are complex systems that consist of many components, joints, and force
elements. Their motion is three-dimensional and is composed of both large translations
and finite rotations (Greenwood 1988; Huston 1990; Roberson and Schwertassek 1988;
Rosenberg 1977). Some railroad vehicle components, such as wheelsets, spin at very high
speeds. Wheel/rail contact must be formulated using a three-dimensional, fully nonlinear
approach in order to accurately represent vehicle motion. Simplified planar and linearized
approaches are not suited for the analysis of modern railroad vehicle systems, particularly
when such systems operate at speeds that do not justify simplifications or linearization.
For this reason, using a fully nonlinear MBS approach is necessary for the analysis, design,
and performance evaluation of modern railroad vehicle systems.

Complexity of Railroad Vehicles Railroad vehicles consist of interconnected com-


ponents that can experience large relative displacements with respect to each other. One
rail car has a large number of components that have distributed inertia that cannot be
modeled using discrete elements. Figure 22 shows an example of a rail car that consists
of several subsystems. The car body of this rail car, as shown in the figure, is mounted on
two bogies (also referred to as trucks). The car body has distributed inertia, and in accurate
simulation models, the car inertia cannot be represented by a concentrated mass. In many
investigations, the distributed elasticity of the car body must also be accounted for using a
continuum-based approach, such as the finite element (FE) method. Figure 23 shows an
Introduction 29

Figure 1.22 Rail car. Source: rruntsch/Adobe Stock.

example of a bogie on which the wheelsets are mounted. The bogie consists of several com-
ponents with distributed inertia that include two wheelsets, two equalizers, a frame, and a
bolster. The car body is mounted on the bogie at the center plate, and this connection is often
modeled using a pin (revolute) joint. In addition to this pin-joint connection, secondary sus-
pensions are used between the car body and the bogie bolster. This secondary suspension,
as shown in the figure, has springs and dampers in order to create a vibration-isolation
system and reduce the effect of the force transmitted from the bogie to the car body. The
primary suspensions, shown in the figure, are used to connect the wheelsets to the bogie
frame. The primary suspensions serve the purpose of reducing frame vibration that results
from the wheel/rail contact forces.

Secondary spring Center plate


Secondary damper

Bolster

Frame

Wheelset Primary spring


Equalizer
Bearing

Figure 1.23 Bogie components.


30 Mathematical Foundation of Railroad Vehicle Systems

The vehicle components described in this section represent the basic components that
must be included in a MBS railroad vehicle model. In addition to these basic components, a
large number of bushings and bearing elements must be included in order to develop real-
istic simulation models. As discussed in this book, a simple rigid body model of a railroad
vehicle system must also take into consideration the track geometry and three-dimensional
wheel/rail contact. In the case of high-speed trains, a pantograph/catenary system model
need also be included, as discussed in a later section. The flexibility of the vehicle compo-
nents can have a significant effect on vehicle dynamics and stability. Accounting for the
car body, track, and catenary deformations using a three-dimensional continuum-based
approach may be necessary in order to obtain accurate results in analysis, performance
evaluation, and accident investigations.

Computational Approaches The dynamics of complex mechanical systems such as


railroad vehicle systems is governed by a system of differential/algebraic equations (DAE).
The differential equations represent the second-order ordinary differential equations of
motion, while the algebraic equations describe the constraints imposed on the motion of
the system. The constraint equations represent mechanical joints and specified motion
trajectories. In this book, the motion of a multibody vehicle system is described using n
[ ]T
coordinates defined by the vector q = q1 q2 … qn . These coordinates are related by
kinematic constraints because of mechanical joints and specified motion trajectories; there-
fore, the coordinates are not independent. As explained in Chapter 6, using the principles
of mechanics, the system differential equations of motion can be written as

Mq̈ = Qe + Qv + Fc (1.26)

where M is the system mass matrix, Qe is the vector of applied forces, Qv is the vector of
Coriolis and centrifugal inertia forces, and Fc is the vector of constraint forces. As explained
in Chapter 6, the constraint equations that represent algebraic relationships between the
[ ]T
system coordinates q = q1 q2 … qn and describe the system mechanical joints and
specified motion trajectories can be written in a vector form as
[ ]T
C (q, t) = C1 C2 … Cnc = 𝟎 (1.27)

where nc is the number of algebraic constraint equations. Because each algebraic equation
can be used to write a dependent coordinate in terms of the independent coordinates, the
[ ]T
vector of coordinates can be written in a partitioned form as q = qTi qTd , where qi is
the vector of independent coordinates or system degrees of freedom, and qd is the vector of
dependent coordinates. The number of dependent coordinates is equal to the number of
algebraic constraint equations nc , while the number of degrees of freedom is nd = n − nc .
Each constraint equation in Eq. 27 introduces an independent constraint force, and there-
fore, for given applied forces, the number of unknowns in Eq. 26 is n + nc . These unknowns
are n accelerations q̈ and nc independent constraint forces that appear in vector Fc in Eq. 26.
Therefore, the following n + nc DAE system is sufficient for solving for all unknowns:
}
Mq̈ = Qe + Qv + Fc ,
[ ]T (1.28)
C (q, t) = C1 C2 … Cnc = 𝟎
Introduction 31

Chapter 6 presents procedures for solving the DAE system defined by the preceding
equations. These procedures lead to different forms of the dynamic equations of motion.
One of the procedures used for solving the DAE system is the augmented formulation,
which leads to a large system of equations that has a sparse matrix structure. Another
approach is to use the embedding technique, which leads to a number of equations
equal to the system number of degrees of freedom; these equations do not include any
constraint forces.

Coordinate Selection While different approaches can be used to formulate the


dynamic equations of motion of constrained dynamical systems, the Lagrangian approach
lends itself easily for developing general algorithms for the computer-aided analysis of
railroad vehicle systems. In the Lagrangian approach, the concepts of virtual displacement,
virtual work, generalized coordinates, and generalized forces are fundamental. As previously
discussed and explained in more detail in Chapter 3, the general unconstrained spatial
motion of a rigid body is described using six independent coordinates. Three of these coor-
dinates describe the translation of the body reference point, and three rotation coordinates
define the body orientation. The orientation coordinates can be three Euler angles or
four Euler parameters, as discussed in Chapter 3. In Chapter 6, the Cartesian translation
and orientation coordinates are referred to as the absolute generalized coordinates, which
define, respectively, the body translation and orientation with respect to a selected global
coordinate system. Another set of coordinates that is also used in formulating the dynamic
equations of railroad vehicle systems is the set of trajectory coordinates as which includes
coordinates that define the body translation and orientation with respect to a body-track
coordinate system that follows the body motion. This body-track coordinate system is
referred to in this book as the trajectory body coordinate system. The trajectory coordinates
will be briefly discussed later in this section and are discussed in more detail in Chapter 6.

Newtonian and Lagrangian Approaches Generalized coordinates are used in this


book to define the global position of the origin and the orientation of the body coordinate
system. Because the formulations presented in Chapter 6 are based on the Newton–Euler
equations of motion, the body reference point, which defines the origin of the body coor-
dinate system, is assumed to be attached to the body center of mass in order to eliminate
the inertia coupling between the body translation and rotation. The virtual work princi-
ple used in the Lagrangian formulation is introduced and used with expressions for the
absolute velocity and acceleration vectors of an arbitrary point on the body to formulate
the body equations of motion. Chapter 6 also explains using contact conditions to formu-
late the wheel/rail interaction forces and discusses hunting oscillations using a constrained
wheelset model that accounts for the coupling between the lateral and yaw displacements.
In formulating the MBS constrained dynamic equations of motion, two approaches are
commonly used: the Newtonian approach and the Lagrangian approach. The Newtonian
approach, which is based on vector mechanics, requires the use of free-body diagrams by
making cuts at the joints. On these free-body diagrams, the joint reaction forces, as well as
the inertia and applied forces, are shown. This approach is not well-suited for developing
general MBS algorithms and can be used for developing the dynamic equations of motion
of relatively simple systems. For the analysis of complex systems, such as railroad vehicles,
32 Mathematical Foundation of Railroad Vehicle Systems

the Lagrangian approach, which has its roots in D’Alembert’s principle and employs
scalar quantities such as virtual work and kinetic and potential energies, is often used
for developing general-purpose MBS algorithms. When the Lagrangian approach is used,
there is no need for free-body diagrams or for the study of the equilibrium of the bodies
separately. This is mainly because the Lagrangian approach is based on connectivity
conditions, which describe mechanical joints in the system and can be formulated using
nonlinear algebraic constraint equations. In the Lagrangian approach, joint forces take
a standard form expressed in terms of the constraint Jacobian matrix and multipliers
called Lagrange multipliers. The fact that algebraic constraint equations can be used to
systematically define joint reaction forces allows for developing general computational
procedures for the computer-aided analysis of a wide class of physics and engineering
systems without the need for using free-body diagrams.

Trajectory Coordinates Some specialized railroad vehicle formulations do not use the
absolute Cartesian coordinates adopted in the development of general MBS algorithms.
Instead, another set of coordinates is adopted: trajectory coordinates as described in
Chapter 3. The detailed formulation of the equations of motion in terms of trajectory
coordinates is also presented in Chapter 6. Trajectory coordinates are also used in devel-
oping longitudinal train dynamics (LTD) algorithms in which each vehicle is normally
represented by a single body with one degree of freedom that defines the distance traveled
along the track centerline.
As explained in Chapter 3, the trajectory and Cartesian coordinates are related by a veloc-
ity transformation, thereby allowing for systematically converting the equations of motion
expressed in terms of Cartesian coordinates to equations expressed in terms of trajectory
coordinates. As shown in Figure 24, the general displacement of a rigid body i in a vehicle
system can be described using six trajectory coordinates
[ ]T
pi = si yir zir 𝜓 ir 𝜙ir 𝜃 ir (1.29)

X
Y
O

Figure 1.24 Trajectory coordinates.


Introduction 33

where si is the arc length coordinate of the track space curve; yir and zir are, respectively,
the lateral and vertical displacements of the origin Oi of the body coordinate system X i Y i Z i
relative to the trajectory body coordinate system X ti Y ti Z ti that follows the body as shown in
Figure 24; and 𝜓 ir , 𝜙ir , and 𝜃 ir are, respectively, the yaw, roll, and pitch angles that define
the orientation of the body coordinate system X i Y i Z i with respect to the trajectory body
coordinate system X ti Y ti Z ti . As discussed in Chapter 3, if the arc length parameter si is
known, the location of the origin and the matrix that defines the orientation of the trajec-
tory body coordinate system X ti Y ti Z ti can be defined and written in terms of the arc length
parameter as Rti = Rti (si ) and Ati = Ati (si ), respectively. Using trajectory coordinates has
the advantage of simplifying the specified-motion constraints. However, it has the disad-
vantage of making the formulation and the computational algorithm more complex, less
user-friendly, and more difficult to generalize for the analysis of flexible bodies.

1.8 DERAILMENT CRITERIA


Railroad derailments involve complex three-dimensional dynamic behavior and forces
that cannot be captured using simplified planar analysis or linearized models. Three-
dimensional wheel/rail contact formulations must be adopted in order to accurately
predict the spatial motion of the wheel with respect to the rail. Using three-dimensional
analysis requires a full parameterization for the wheel and rail surfaces because the theory
of curves alone is not sufficient to capture the complexity of some derailment and wheel
climb scenarios. For example, wheel climb can occur at a low velocity with a relatively large
angle of attack (<3∘ ). The angle of attack 𝛼 a is defined as the angle between the forward
velocity of the wheel and the longitudinal tangent tr1 of the rail, as shown in Figure 25.

k)
Y of attac
V α(A ngle
X

tr1 (Rail longitudinal tangent)

Figure 1.25 Angle of attack.

A large number of investigations have been devoted to the development and validation
of derailment criteria (Blader 1989; Elkins and Wu 2000; Marquis and Grief 2011; Shust
et al. 1997; Wu and Elkins 1999; Wilson et al. 2004). These criteria have been widely used
for developing safety and operation guidelines as well as in accident investigations. Some of
these criteria are based on the ratio L/V, where L and V are, respectively, the lateral and ver-
tical forces acting on the wheel, as shown in Figure 26. Examples of derailment criteria are
34 Mathematical Foundation of Railroad Vehicle Systems

L
N
F
α

Figure 1.26 Lateral and vertical forces in Nadal’s criterion.

the Nadal single-wheel L/V limit criterion, Weinstock axle-sum L/V limit criterion, Federal
Railroad Administration (FRA) high-speed passenger distance limit (5 ft), Association of
American Railroads (AAR) Chapter 11 50-ms time limit, Japanese National Railway (JNR)
L/V time duration criterion, Electro-Motive Diesel (EMD) L/V time duration criterion, and
Transportation Technology Center, Inc. (TTCI) wheel climb distance criterion.
General computational MBS approaches can be used to investigate derailment scenarios
using more realistic railroad vehicle models that include significant details. These compu-
tational algorithms can be used to develop derailment criteria based on three-dimensional
analysis and check the accuracy of these criteria using fully nonlinear spatial models. The
ratio between the lateral force L and the vertical force V acting on the wheel, as shown
in Figure 26, is used as the basis for developing several existing derailment criteria, as
previously mentioned. In these criteria, it is assumed that if the lateral force exceeds a cer-
tain limit, derailment can occur. The L/V ratio can be more accurately predicted based on
fully nonlinear MBS models, and the results can be compared with those obtained using
simplified approaches such as Nadal’s formula (Nadal 1908), which is often adopted for
determining the limit of the L/V ratio.
In the simple analysis used to develop the L/V ratio of Nadal’s formula, it is assumed that
the lateral and vertical forces L and V acting on the rail are balanced by the forces N and
F that apply on the wheel, where N is the normal reaction force, F = 𝜇N is the tangential
friction force, and 𝜇 is the coefficient of friction. If the wheel flange angle is 𝛼, a simple force
analysis can be used to show that the L/V ratio can be written as
L tan 𝛼 − 𝜇
= (1.30)
V 1 + 𝜇 tan 𝛼
In some of the existing derailment criteria, if the L/V ratio exceeds the right side of Eq. 30,
wheel climb is assumed to occur. Because of the shape of the flange, the flange angle is
not constant as the wheel climbs the rail. In Nadal’s formula, defined by the preceding
equation, the maximum wheel-flange contact angle is used. Furthermore, in some criteria,
a positive angle of attack is assumed when applying the preceding equation, despite the fact
this equation is not a function of the angle of attack.
Because of the complexity of wheel/rail interaction forces, it is difficult to have a universal
limit for the L/V ratio in all motion scenarios. For example, Blader (1989) suggested that,
in order to avoid derailments when the wheel/rail contact point is located at the rail gage
point, the L/V ratio must be within the range 0.66–0.73, depending on the shape of the
rail profile. A high value for the lateral force acting on the track can be the source of other
problems that can lead to derailments. For example, a high-magnitude lateral force not only
Introduction 35

(a) (b)
V

hr

wr
(c)

Figure 1.27 Gage widening and rail rollover.

causes wheel climbs but also can lead to gage widening and rail rollover. Gage widening can
lead to wheel/rail separation, as shown in Figure 27a. Rail rollover, shown in Figure 27b, is
also a common source of accidents that occur when the vehicle negotiates a spiral section
of track. It is clear that if the L/V ratio is higher than the ratio wr /hr , where wr and hr are
shown in the figure, the moment L × hr generated by the lateral force L becomes higher
than the moment V × wr generated by the vertical force V, and this can lead to rail rollover
as the result of the rail rotation about its corner.
While a simple analysis can give insight on the issues encountered in the analy-
sis of railroad vehicle derailments and accident investigations, wheel/rail interaction
forces, as discussed in this book, are better understood by developing more detailed
three-dimensional models that account for displacement modes that cannot be repre-
sented using a simple planar analysis model. For instance, wheel climb at an angle of
attack can be better understood using a spatial analysis, despite the fact that the yaw angle
remains small compared with the pitch angle of rotation of the wheelset about its axis.
Because finite and infinitesimal rotations do not commute, using linearized equations
to investigate derailments can lead to wrong force prediction, inaccurate velocity results,
and wrong conclusions regarding the root causes of derailments of complex railroad
vehicles.
36 Mathematical Foundation of Railroad Vehicle Systems

Example 1.1 The derivation of Nadal’s formula of Eq. 30 is based on simple force
balance. Using Figure 26, one can write
L = N sin 𝛼 − F cos 𝛼
V = N cos 𝛼 + F sin 𝛼
The flange tangential force F can be expressed in terms of the normal force as F = 𝜇N.
Substituting this equation into the expressions of L and V, one obtains
L = N(sin 𝛼 − 𝜇 cos 𝛼)
V = N(cos 𝛼 + 𝜇 sin 𝛼)
It follows that
L N(sin 𝛼 − 𝜇 cos 𝛼) tan 𝛼 − 𝜇
= =
V N(cos 𝛼 + 𝜇 sin 𝛼) 1 + 𝜇 tan 𝛼
It is recommended to use a flange angle 𝛼 above 70∘ in order to increase the limiting
value of the L∕V ratio. Old rail systems used a wheel flange angle 𝛼 ≈ 63–65∘ , while for
modern rail system, the wheel flange angle 𝛼 ≈ 72–75∘ . In the case of dry wheel/rail
contact conditions, the friction coefficient 𝜇 can be assumed equal to 0.5, that is 𝜇 = 0.5.
Using this value of the friction coefficient, one can show that the limiting value for the
L∕V ratio is 1.1276 for 𝛼 = 72–75∘ ; and this L∕V ratio is 0.7382 for 𝛼 = 63∘ . This shows
that larger flange angle allows for supporting higher lateral force, and this can contribute
to the vehicle stability by avoiding derailments.

1.9 HIGH-SPEED RAIL SYSTEMS


As the demand for higher rail-transportation speed increases, more scientific challenges
emerge. Modern high-speed rail transportation systems are being designed to operate at
speeds of around 370 km/h (220 mph). Nonetheless, research and experimentation are cur-
rently focused on trains that can have speeds close to or above 500 km/h (298 mph). At
these high speeds, contact and friction introduce many challenging problems that cannot
be scientifically examined using some of the contact formulations currently being used.
In addition to the wheel/rail contact problem previously discussed in this chapter, which
is a subject of more detailed discussion in later chapters, there are challenging problems
that are particular to high-speed trains. High-speed trains are electrically powered using a
pantograph/catenary system like the one shown in Figure 28. Electrification is not normally
used for freight trains, which operate at much lower speeds, have higher axle loads, can
consist of a much larger number of cars, and use lower-quality tracks. Freight trains are
normally driven by diesel locomotives, which have a speed limit of approximately 238 km/h
(148 mph). As a result of this speed limitation, diesel engines cannot be used for high-speed
passenger trains. Extensive experimentation is currently under way to significantly increase
the train speed to a level that allows minimizing travel time, and at such high speeds,
the pantograph/catenary technology presents itself as the only viable power-supply option.
Electrification systems, which were used in the 1800s to provide the electric power sup-
ply for other transportation systems such as trams, trolleys, and buses, are quieter, more
reliable, and less costly, and can provide the power needed for higher speeds.
Introduction 37

Pantograph/catenary system

Figure 1.28 Pantograph/catenary system. Source: hpgruesen/Pixabay.

Electric Power Collection As shown in Figure 28, the pantograph is often mounted on
the top of one of the train cars. Multiple pantographs can be used for a single train in order
to draw sufficient power as the train speed increases. However, there has to be a minimum
distance between two pantographs. The voltage used for electric train operations, which can
reach 25 kV AC at 50–60 Hz, is generated using electric feeder stations at several locations
along the track. The pantograph mechanism collects power through contact with an over-
head power line, called the catenary, which receives electric power from feeder stations. The
pantograph consists of several components connected by joints, while the catenary consists
of several cables and can be treated as a structural system. Electric current is collected by
the train using the pantograph, which normally has a carbon strip that maintains contact
with a catenary wire called the contact wire. Contact is maintained with the catenary by
applying uplift force using pneumatic actuators attached to the pantograph. The combined
pantograph/catenary system is also called the current collection system. Current collection
systems allow electric current to flow through to the train and back to the feeder station
through wheel/rail contact. This arrangement for the flow of current represents a closed
electric circuit.

Pantograph/Catenary Contact To achieve an uninterrupted power supply to the


train, loss of contact between the pantograph carbon strip and catenary contact wire should
be minimized. This loss of contact can lead to electric arcing that can cause damage to
the carbon strip and contact wire. In addition to the pantograph uplift force, maintaining
continuous contact between the pantograph carbon strip and catenary contact wire can be
achieved by reducing contact wire vibration. This uninterrupted contact can be achieved
by increasing the bending stiffness of the catenary wires using pretension as well as using
other cables such as a messenger wire and droppers to support the contact wire and reduce
its oscillations resulting from contact and aerodynamic forces.
Pretension of the catenary wires, which leads to higher bending stiffness as a result of
the coupling between axial and bending deformations, is necessary to maintain catenary
38 Mathematical Foundation of Railroad Vehicle Systems

stability and ensure stable contact with the pantograph carbon strip. As the stiffness of the
contact wire increases, the speed of the propagation of elastic waves resulting from the pan-
tograph/catenary interaction also increases to a level that is higher than the vehicle oper-
ating speed. To avoid resonance, the frequency of the wire oscillations and the speed of
the elastic waves must be much higher than the train operating speed, and this can be
achieved using pretension of the catenary wires. The pretension can be made independent
of weather conditions by using balance weights or hydraulic tensioners, as described in
Chapter 7.

Pantograph/Catenary Mathematical Formulations A large number of investiga-


tions have been devoted to studying pantograph/catenary system dynamics and vibration.
Some of these investigations are based on linear models, while others propose nonlinear
models with varying degrees of complexity and assumptions. In some of these models, the
pantograph mechanism is modeled as a system of masses and springs, and the catenary
is modeled using simple discrete spring-damper elements. Such simplified models do not
account for the effect of pantograph joint articulation or distributed inertia and elasticity
of the catenary cables. For this reason, simplified models may not accurately predict the
system dynamics and stresses that are necessary for a credible evaluation of mechanical
pantograph design and catenary structural integrity and durability. Such credible assess-
ment can be made using MBS computational approaches for modeling pantograph dynam-
ics, and continuum-based approaches that account for the distributed inertia and elasticity
of catenary wires, as discussed in Chapter 7. MBS approaches for modeling a pantograph
system allow for the accurate representation of the rotations of pantograph components,
mathematical description of joints between these components, systematic computation of
joint and actuator forces, and accurate prediction of contact forces. Continuum-based mod-
els of the catenary wire allow for a more accurate evaluation of the speed of elastic wave
propagation and the implementation of different material models as well as accounting for
geometric nonlinearities that result from large-amplitude oscillations.

ANCF Catenary Model As previously mentioned, the train operating speed should
be kept lower than the speed of the propagation of catenary elastic waves due to safety
considerations (Kumaniecka and Snamina 2008; Pappalardo et al. 2016). Therefore, it is
recommended to use a continuum-based approach, such as the absolute nodal coordinate
formulation (ANCF) for catenary models, as discussed in Chapter 7. Discrete spring models
do not allow for modeling the propagation of elastic waves or assessing the wear and dura-
bility of catenary wires. ANCF finite elements can be systematically integrated with general
MBS computational algorithms to develop detailed models of the pantograph/catenary
system. Formulating the catenary equations of motion using two different ANCF finite
elements is discussed in Chapter 7. These elements are the gradient-deficient cable element
and the fully parameterized beam element. While these two elements use different numbers
of parameters, a unified approach for modeling contact with the pantograph is presented
in Chapter 7. Using ANCF finite elements avoids the use of incremental-rotation and
co-simulation approaches, since the resulting dynamic equations can be solved using a
non-incremental-rotation solution procedure.
Introduction 39

Pantograph/Catenary Contact Formulations As in the case of wheel/rail contact


previously discussed in this section, two fundamentally different formulations can be used
to develop pantograph/catenary interaction forces: the constraint contact formulation (CCF)
and the elastic contact formulation (ECF). In the constraint contact formulation, the panto-
graph carbon strip is assumed to remain in contact with the catenary wire, and therefore,
separations or penetrations are not allowed. In this case, pantograph/catenary loss of con-
tact, which is a source of arcing, cannot be modeled. Nonetheless, a constraint contact
formulation can be developed to model the longitudinal and lateral relative motion between
the pantograph and catenary. In the CCF approach, the normal contact force is determined
as a constraint (reaction) force. In the elastic contact formulation, on the other hand, panto-
graph/catenary separations and penetrations are allowed, and therefore, the loss of contact
can be modeled. In the ECF approach, no constraints are imposed on the motion of the
pantograph carbon strip with respect to the catenary wire, and the pantograph/catenary
normal contact force is described using a compliant force model with assumed stiffness and
damping coefficients. The constraint contact formulation is recommended in simulation
scenarios in which it is desirable to have a smoother solution by avoiding the oscillations
and high frequencies that result from using an elastic contact force formulation. Avoiding
these oscillations and high frequencies makes it easier to observe and assess some dynamic
behaviors and phenomena that can be difficult to observe and assess when high-frequency
oscillations are superimposed on solutions. Therefore, it is recommended to implement
both the constraint and elastic contact formulations in computer software developed for
nonlinear dynamic simulations of railroad vehicle systems, to be able to conduct research
investigations and perform computer simulations of practical motion scenarios.

Pantograph/Catenary Contact Force Control Maintaining pantograph/catenary


contact and ensuring system stability at high speeds are necessary for smooth train oper-
ations. This is a challenging problem because of disturbances that can result from aero-
dynamics forces, rail car vibrations, and track irregularities. These disturbances can have
an adverse effect on current collection quality as a result of variations in the contact force
between the pantograph strip and catenary wire. For this reason, investigations have been
conducted to study the effectiveness of developing control algorithms that ensure the sta-
bility of pantograph/catenary As previously mentioned, contact and ensure smooth train
operation, particularly at high speeds.
The design of effective control systems requires accurate modeling of train dynamics
(Poetsch et al. 1997). The stability of the current collection system depends on the dynamic
behavior of the pantograph/catenary system and its response to disturbances. As previously
mentioned, contact between the pantograph carbon strip and catenary contact wire is nor-
mally maintained using an uplift force exerted by actuators on the pantograph lower arm.
The magnitude of the uplift force should not be very high, to avoid a high contact force
between the pantograph strip and catenary wire that can lead to increased wear as a result
of high friction forces. Low uplift forces, on the other hand, can lead to loss of contact, which
increases the probability of electrical arcing. Arcing, which is the result of electric current
flow in an air gap between the catenary contact wire and pantograph contact strip, leads
to electric sparks with intense light that can cause damage to the contact wire and contact
strip (Hsiao 2010). One solution to this problem is to use an active control system by placing
40 Mathematical Foundation of Railroad Vehicle Systems

an actuator between the pantograph components, with the goal of reducing contact force
variations, as discussed in Chapter 7 and the literature (Pappalardo et al. 2016).

Effect of Aerodynamic Forces Environmental conditions such as temperature and


wind can have a significant impact on pantograph/catenary interaction forces. For example,
high temperature can alter the static equilibrium position and pretension in the catenary
wires, while cold weather conditions can result in the formation of ice on the wires, leading
to undesirable deformations and poor contact and current collection quality. Aerodynamic
forces, on the other hand, can result in severe vibrations of the catenary cables as well
as undesirable forces acting on the components of the pantograph system that negatively
influence its functional operation. Aerodynamic drag and lift forces can cause variations
in contact forces, wear, and loss of contact. Wear can generate asymmetric drag and lift
forces, leading to galloping catenary motion (Stickland Scanlon 2001; Stickland et al. 2003),
while high cross-wind loads can also cause severe vehicle vibrations that directly influence
pantograph/catenary interaction (Bacciolone et al. 2008b; Cheli et al. 2010). Aerodynamics
drag and lift force components alter the uplift force exerted on the pantograph mecha-
nism. For double-arm pantographs, aerodynamic forces can cause an imbalance, result-
ing in faster wear of one of the collector strips (Bacciolone et al. 2006a; Carnevale et al.
2016; Pombo et al. 2009). Furthermore, boundary layer turbulence near the car body roof
due to vortex shedding can excite pantograph components, generate high frequencies, and
increase the sparking level; this, in turn, can negatively impact the current collection qual-
ity (Bacciolone et al. 2006a,2006b; Ikeda et al. ). For these reasons, evaluating the effect
of aerodynamic forces on pantograph/catenary dynamics is an important design consid-
eration. As pointed out by Kulkarni et al. (2017), in general, two main approaches can be
used to evaluate the effect of aerodynamic forces on pantograph/catenary systems. In the
first approach, computational fluid dynamics (CFD) is used to calculate the effect of aerody-
namic forces on pantograph system components and to examine the contribution of these
forces to total uplift pantograph force (Bacciolone et al. 2007; Carnevale et al. 2016). In the
second approach, drag and lift coefficients are obtained from experimental studies, and aero-
dynamic forces are computed using simpler equations. Developing an aerodynamic force
model that can be applied to pantograph/catenary systems in which the catenary wires are
modeled using continuum-based ANCF finite elements is described in Chapter 7.

Wear Effect Developing a general computational MBS algorithm for predicting panto-
graph/catenary wear allows incorporating the effects of vehicle vibration, wheel/rail con-
tact forces, and track irregularities. Such an algorithm also allows for predicting the wear
rate in the case of different motion scenarios that require the use of nonlinear models,
including curve negotiations, accelerations, and braking (Daocharoenporn et al. 2019). In
these motion scenarios, the effects of contact forces resulting from vehicle dynamics on
the wear rates of the catenary wire can be significant. Severe environmental and operating
conditions may cause arcing, high wear rates, or even failure of the pantograph/catenary
system to provide desired uninterrupted electric power necessary for train operation. Arc-
ing, for example, may increase significantly the wear rate of the pantograph contact strip,
leading to variations of contact forces and a deterioration in pantograph/catenary system
performance.
Continuous localized contact between the pantograph carbon strip and catenary contact
wire can cause significant wear to the carbon strip inserted on the pantograph top and used
Introduction 41

for electric current collection. Between two overhead line supports, the contact wire seg-
ment is straight; therefore, during curve negotiations, the contact wire sweeps laterally over
the whole carbon strip, resulting in uniform wear. When the vehicle negotiates a tangent
track, it is also desirable to obtain a pattern of contact in which the contact wires sweep
laterally over the surface of the carbon strip to achieve uniform (instead of localized and
more severe) wear. Therefore, in the case of a tangent track, the contact catenary wire is
slightly zigzagged around the centerline of the track to produce a lateral sweep that leads
to uniform wear.
As discussed in Chapter 7, several important factors can have a direct effect on panto-
graph/catenary wear, including the design of the contact wire, pretension in the catenary
cable, and uplift force of the pantograph mechanism. For example, using a high uplift force
can lead to a significant increase in the magnitude of the contact force, and this in turn
can lead to an increased wear rate. Therefore, the wear rate can be reduced by control-
ling this uplift force of the pantograph mechanism, as previously mentioned. The wear rate
can also be reduced by properly designing the pantograph/catenary system to better handle
aerodynamic forces, staggering the contact wire, controlling the intensity of the collection
current, using the proper materials for the contact wire and contact strip to reduce friction,
and properly adjusting the pretension of the catenary cables (Bucca and Collina 2009; Bucca
and Collina 2015; Daocharoenporn et al. 2019). A wear model that accounts for electric and
mechanical effects such as the electric arcing effect, Joule effect of the electrical current, and
effect of contact friction forces is presented in Chapter 7 based on the model developed by
Bucca and Collina (2015).

1.10 LINEAR ALGEBRA AND BOOK NOTATIONS


Vector, matrix, and tensor algebra are used to develop static, kinematic, and dynamic
equations of physics and engineering systems in compact forms. They are also used to
perform addition and multiplication operations required for the proofs of many important
identities. Therefore, vector, matrix, and tensor algebra are integral parts of engineering
and physics courses, including courses at the undergraduate level. Therefore, it is assumed
that the reader has some familiarity with the subject of linear algebra. This section reviews
some specific topics of linear algebra that are frequently used in this book. The notations
used in the book are also described.

Cross Product and Skew-Symmetric Matrices The relationship between the cross
product and skew-symmetric matrix representation is used in this book to define the angu-
̃ is said to be skew-symmetric if its elements aij , i, j = 1,
lar velocity vector. A square matrix A
2, …, n, where n is the order of the matrix, satisfy the relationship aij = − aji . This implies
that all the diagonal elements of a skew-symmetric matrix are equal to zero, since zero is
the only number that is equal to its negative. An example of a 3 × 3 skew-symmetric matrix
̃ can be written as
A
⎡ 0 −a3 a2 ⎤
̃ = ⎢⎢ a3
A 0 −a1 ⎥

(1.31)
⎢ ⎥
⎣−a2 a1 0 ⎦
It is clear that this form of skew-symmetric matrix can be constructed using the elements
[ ]T
of the vector a = a1 a2 a3 . Therefore, given any three-dimensional vector a, a unique
42 Mathematical Foundation of Railroad Vehicle Systems

skew-symmetric matrix A ̃ is associated with the vector. In this book, the skew-symmetric
[ ]
̃ associated with the vector a = a1 a2 a3 T is assumed to be in the form given
matrix A
by the preceding equation. It is also important to note that a matrix that is equal to the
negative of its transpose must be skew-symmetric: that is, if A = − AT , then A must be a
skew-symmetric matrix.
[ ]T
The cross product between two three-dimensional vectors a = a1 a2 a3 and
[ ]T
b = b1 b2 b3 is defined as
| i j k|
| |
| |
a × b = ||a1 a2 a3 ||
| |
|b b b |
| 1 2 3|
( ) ( ) ( )
= a2 b3 − a3 b2 i + a3 b1 − a1 b3 j + a1 b2 − a2 b1 k (1.32)
where i, j, and k are unit vectors along the axes of the coordinate system in which the
components of vectors a and b are defined. It is clear that the cross product of the preceding
equation can be written as
⎡ a 2 b3 − a3 b2 ⎤
⎢ ⎥
a × b = ⎢ a3 b1 − a1 b3 ⎥ (1.33)
⎢ ⎥
⎣ a1 b2 − a2 b1 ⎦
This equation can also be written as
⎡a2 b3 − a3 b2 ⎤ ⎡ 0 −a3 a2 ⎤ ⎡b1 ⎤
⎢ ⎥ ⎢ ⎥⎢ ⎥ ̃
a × b = ⎢ a3 b1 − a1 b3 ⎥ = ⎢ a3 0 −a1 ⎥ ⎢b2 ⎥ = Ab (1.34)
⎢ ⎥ ⎢ ⎥⎢ ⎥
⎣a1 b2 − a2 b1 ⎦ ⎣−a2 a1 0 ⎦ ⎣b3 ⎦
This equation shows that the cross product of two vectors a and b can be written in terms of
[ ]
the skew-symmetric matrix A ̃ associated with the vector a = a1 a2 a3 T . Similarly, one
can show that a × b = −b × a = −Ba,̃ where B ̃ is the skew-symmetric matrix associated
with vector b. The skew-symmetric form of the cross product is repeatedly used in this
book and is used to obtain a general definition of the angular velocity vector in the spatial
analysis.
[ ]T
Orthogonal Matrices Two n-dimensional vectors a = a1 a2 … an and
[ ]T
b = b1 b2 … bn are said to be orthogonal if their dot product is equal to zero:
that is,

n
a • b = a T b = a1 b1 + a2 b2 + · · · + an bn = ai bi = 0 (1.35)
i=1
A square matrix A is said to be orthogonal if
AAT = AT A = I (1.36)
where I is the identity matrix. It follows that the transpose of an orthogonal matrix is equal
to its inverse: that is, AT = A−1 . Orthogonal matrices are encountered in formulating the
dynamic equations of motion of mechanical systems. The transformation matrices that
define the orientations of coordinate systems in space are orthogonal matrices. The columns
of these transformation matrices define orthogonal unit vectors along the axes of coordinate
Introduction 43

systems. Therefore, for such orthogonal matrices, no element of the transformation matrix
should be greater than one.
The orthogonality condition of Eq. 36 is used in this book to obtain a general definition of
the angular velocity vector. In this case, the orthogonal matrix that defines the orientation
of the body coordinate system is written in terms of a set of orientation parameters that
depend on time. If A is an orthogonal matrix, and the identity AAT = I is differentiated
with respect to time, one obtains
d ( T) ̇ T + AȦ T = 𝟎
AA = AA (1.37)
dt
Keeping in mind that (AB)T = BT AT , then for any arbitrary matrices A and B that have the
right number of rows and columns for a valid matrix product, the preceding equation can
be written as
( T )T
̇ T = −AȦ T = − AA
AA ̇ (1.38)
This equation shows that AA ̇ T is equal to the negative of its transpose, and therefore, it is
a skew-symmetric matrix that can be written as AA ̇ T =𝛚̃ , where 𝛚̃ can be used to define a
[ ]T
vector 𝛚 = 𝜔1 𝜔2 𝜔3 . Following a similar procedure, one can use the identity AT A = I
( )T
to show that Ȧ T A + AT Ȧ = 𝟎, which leads to Ȧ T A = −AT Ȧ = − Ȧ T A = 𝛚, ̃ where 𝛚̃ can
[ ]T
be used to define a vector 𝛚 = 𝜔1 𝜔2 𝜔3 . The procedure used to define the two vectors
[ ]T [ ]T
𝛚 = 𝜔1 𝜔2 𝜔3 and 𝛚 = 𝜔1 𝜔2 𝜔3 is followed in Chapter 3 to develop a general
expression for the angular velocity vector.

Differentiation of Vector Functions As previously discussed in this chapter, the con-


figuration of a railroad vehicle system is defined using the generalized coordinates q =
[ ]T
q1 q2 … qn . These coordinates are not, in general, independent, because of mechani-
cal joints and specified motion trajectories. The kinematic constraint equations imposed on
[ ]T
system motion can be written in a vector form as C (q, t) = C1 C2 … Cnc = 𝟎, where C
is the vector of constraint functions and nc is the number of algebraic constraint equations.
In formulating the equations of motion presented in this book, it is necessary to evaluate
the derivatives of the constraint equations with respect to time. For a given constraint func-
tion Ck (q, t) = Ck (q1 , q2 , …, qn , t), the total derivative of the function can be written using
the chain rule of differentiation as
dCk 𝜕Ck 𝜕Ck 𝜕Ck 𝜕Ck
= Ċ k = q̇ 1 + q̇ 2 + · · · + q̇ n +
dt 𝜕q1 𝜕q2 𝜕qn 𝜕t
∑n
𝜕Ck 𝜕Ck
= q̇ j + (1.39)
j=1
𝜕q j 𝜕t
This equation can be written as
⎡q̇ 1 ⎤
[ ]⎢ ⎥
dCk 𝜕Ck 𝜕Ck 𝜕Ck ⎢q̇ 2 ⎥ 𝜕Ck
···
𝜕qn ⎢⎢ ⋮ ⎥⎥
= +
dt 𝜕q1 𝜕q2 𝜕t
⎢q̇ ⎥
⎣ n⎦
𝜕Ck 𝜕Ck
= q̇ + (1.40)
𝜕q 𝜕t
44 Mathematical Foundation of Railroad Vehicle Systems

where
[ ]
𝜕Ck 𝜕Ck 𝜕Ck 𝜕Ck
= ··· (1.41)
𝜕q 𝜕q1 𝜕q2 𝜕qn
and 𝜕Ck /𝜕t is the partial derivative of Ck (q, t) with respect to time. The preceding equation
shows that the partial derivative of a scalar function with respect to the coordinates is a
[ ]T
row vector. Therefore, in the case of the vector functions C (q, t) = C1 C2 … Cnc = 𝟎,
one has
⎡ dC1 ∕dt ⎤ ⎡ 𝜕C1 ∕𝜕q1 𝜕C1 ∕𝜕q2 ··· 𝜕C1 ∕𝜕qn ⎤ ⎡q̇ 1 ⎤ ⎡ 𝜕C1 ∕𝜕t ⎤
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
dC ⎢ dC2 ∕dt ⎥ ⎢ 𝜕C2 ∕𝜕q1 𝜕C2 ∕𝜕q2 ··· 𝜕C2 ∕𝜕qn ⎥ ⎢q̇ 2 ⎥ ⎢ 𝜕C2 ∕𝜕t ⎥
=⎢ ⎥=⎢ ⎥⎢ ⎥ + ⎢ ⎥
dt ⎢ ⋮ ⎥ ⎢ ⋮ ⋮ ⋮ ⋱ ⋮ ⎥⎢ ⋮ ⎥ ⎢ ⋮ ⎥
⎢dC ∕dt⎥ ⎢𝜕C ∕𝜕q 𝜕Cnc ∕𝜕q2 ··· 𝜕Cnc ∕𝜕qn ⎥⎦ ⎢⎣q̇ n ⎥⎦ ⎢⎣𝜕Cnc ∕𝜕t⎥⎦
⎣ nc ⎦ ⎣ nc 1

= Cq q̇ + Ct (1.42)
where
⎡ 𝜕C1 ∕𝜕q1 𝜕C1 ∕𝜕q2 · · · 𝜕C1 ∕𝜕qn ⎤ ⎡ 𝜕C1 ∕𝜕t ⎤
⎢ ⎥ ⎢ ⎥
⎢ 𝜕C2 ∕𝜕q1 𝜕C2 ∕𝜕q2 · · · 𝜕C2 ∕𝜕qn ⎥ 𝜕C ⎢ 𝜕C2 ∕𝜕t ⎥
Cq = ⎢ ⎥, Ct =
𝜕t
=⎢ ⎥ (1.43)
⎢ ⋮ ⋮ ⋮ ⋮ ⎥ ⎢ ⋮ ⎥
⎢𝜕C ∕𝜕q 𝜕C ∕𝜕q · · · 𝜕C ∕𝜕q ⎥ ⎢𝜕C ∕𝜕t⎥
⎣ nc 1 nc 2 nc n⎦ ⎣ nc ⎦
The nc × n matrix Cq is the Jacobian matrix, and the vector Ct is the partial derivative of
the vector functions C with respect to time. It is clear from the preceding equation that
each row in the Jacobian matrix corresponds to a function, and each column corresponds
to a coordinate. The Jacobian matrix of the constraint functions plays a fundamental role
in the dynamic formulations considered in this book. This matrix can be used to identify
the degrees of freedom in the case of complex systems, and it can also be used to define the
constraint forces in the Lagrangian formulation discussed in this book.

Book Notations In this book, scalar, vectors, and matrices are used. Italic letters are
used for scalar symbols, while bold-face letters are used for vectors and matrices. Some of
these scalars, vectors, and matrices are associated with certain bodies in multibody vehicle
systems. A scalar a, a vector a, and a matrix A associated with a body i are referred to as ai ,
ai , and Ai , respectively. That is, superscript i is used to refer to the body number. If a scalar is
raised to a certain power, parentheses are used. For example, if ai is raised to the power 3, it
is written as (ai )3 . Using this notation makes clear the difference between the body number
and the power. Scalars, vectors, and matrices can also be defined in different coordinate
systems. If scalar a, vector a, and matrix A are defined in the coordinate system of body i, a
i
bar is used, and the scalar, vector, and matrix are written as ai , ai , and A , respectively.
45

Chapter 2

DIFFERENTIAL GEOMETRY

The geometric description is a fundamental step in the formulation of the equations that
govern the mechanics of railroad vehicle systems. The track and wheel geometries must
be accurately described in order to correctly predict the wheel/rail contact forces. Both the
differential geometry theory of curves and theory of surfaces are required in order to be able
to formulate the wheel and rail kinematic and force equations. These theories are used to
define nonlinear geometric equations that can be solved for the location of the wheel/rail
contact points, as will be explained in a later chapter of this book. Furthermore, the spatial
wheel and rail geometric representations are crucial in the study of railroad vehicle non-
linear dynamics in different motion scenarios, including curve negotiations and travel on
tracks with irregularities and worn profiles that influence ride comfort, vehicle stability,
safe operations, and the possibility of train derailments.

Theory of Curves Curve equations are defined in terms of only one parameter (Do
Carmo 1976; Goetz 1970; Kreyszig 1991). To trace a curve, one has to move in only one
direction along the curve because motion is not allowed in any other directions perpen-
dicular to the tangent to the curve at an arbitrary point. Because of this restriction, the
location of an arbitrary point on the curve in a three-dimensional space can be defined
completely in terms of one parameter. If the value of this parameter is given at an arbitrary
point, the three Cartesian coordinates that define the location of the point can be deter-
mined. While a curve can be parameterized using any scalar variable, the curve arc length
that measures the distance from the curve starting point is often used as the curve param-
eter. Using the parametric form of the curve, a spatial curve can be uniquely defined in
terms of geometric invariants called the curvature and torsion, which appear in the curve
Serret–Frenet equations. A curve can also be given in an implicit form by eliminating the
parameter to obtain a single equation expressed in terms of the curve coordinates.

Theory of Surfaces Unlike curve equations, surface equations are defined in terms of
two parameters because, on a surface, one has the freedom to move in two independent
directions. For example, in the case of a rail surface, a wheel can slide and/or rotate on the
surface longitudinally or laterally. Therefore, for a surface, one can define two independent
tangent vectors that define what is called a tangent plane. The geometric properties of a sur-
face are defined using the first and second fundamental forms of surfaces. The coefficients

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
46 Mathematical Foundation of Railroad Vehicle Systems

of these fundamental forms are used to define the principal curvatures that enter into the
formulation of the contact force equations.

Numerical Description of Geometry To use the theory of differential geometry in


practical applications, polynomial approximations are used for curves and surfaces. In rail-
road vehicle dynamics, using polynomials allows for defining arbitrary rail and wheel pro-
file geometries. In this chapter, a finite element (FE) approach, called the absolute nodal
coordinate formulation (ANCF), is used to define the rail geometry. Using this approach
enables integrating geometry and analysis, particularly if the rail deformation is of concern
and must be predicted. ANCF elements can be used to conveniently describe the geome-
tries of both curves and surfaces. Using the ANCF position-gradient coordinates allows for
describing complex shapes as well as the deformations in the case of flexible rails. The first
two sections of this chapter follow, for the most part, the first two sections of Chapter 3 of a
previous book on railroad vehicle dynamics (Shabana et al. 2008).

2.1 CURVE GEOMETRY


A curve can be described using a single implicit-form equation or in a parametric form
in which the coordinates of an arbitrary point on the curve are written in terms of one
parameter. In the implicit form, the single equation relates the coordinates of an arbitrary
point on the curve. For example, in the case of a planar parabolic curve, one can write
y = (x)2 , where x and y are the Cartesian coordinates of a point on the curve. When using
the parametric form of a spatial curve, on the other hand, the three coordinates of a point
on the curve in a Cartesian coordinate system XYZ, as shown in Figure 1, can be written as
r = [x y z]T , where x = x (t), y = y (t), and z = z (t) are the Cartesian coordinates of the point
on the curve, and t is the parameter. Therefore, a curve defined over the interval a ≤ t ≤ b
can, in general, be written in the parametric form
[ ]T
r (t) = x (t) y (t) z (t) (2.1)
Given the value of the parameter t, the location of a point on the curve can be determined.
The coordinates x = x(t), y = y(t), and z = z(t) must be differentiable over the interval on

Z r

Figure 2.1 Spatial curve.


Differential Geometry 47

which t is defined. In the case of a curve, one can trace points in only one direction, and
motion in a plane perpendicular to this direction is not permitted; so, Eq. 1 is completely
defined once the parameter t is given.

Example 2.1 The simple equation of a parabolic curve, shown in Figure 2, defined
in its implicit form, is y = (x)2 . The parametric form of the parabolic curve can be
written as
x (t) = t, y (t) = (t)2
where t is the parameter. Note that by eliminating the parameter t from the parametric
representation, the single equation y = (x)2 , which defines the implicit curve form, can
be obtained. Note also that in this case, the parameter t = x : that is, the coordinate x is
used as the curve parameter. This choice of the curve parameter is not unique.

y = (x)2

Figure 2.2 Parabolic curve.

Example 2.2 The equation of a circle, shown in Figure 3, with radius a and center at
the origin, is defined by the parametric equations
x (t) = a cos t, y (t) = a sin t
These equations, in which t is the parameter, can be used to obtain the implicit form
(x)2 + (y)2 = (a)2 by simply eliminating t by squaring and adding the two parametric
equations and using the trigonometric identity cos2 t + sin2 t = 1. The parameter t rep-
resents the angle that defines the point on the circle, as shown in Figure 3.

t=θ
X

Figure 2.3 Circle geometry.

(Continued)
48 Mathematical Foundation of Railroad Vehicle Systems

If the center of the circle is not at the origin of the Cartesian coordinate system and is
defined by the coordinates xc and yc , the parametric representation of the curve can be
written as
x (t) = xc + a cos t, y (t) = yc + a sin t
These two equations can be used to define the implicit form of the circle, by eliminating
the parameter t, as (x − xc )2 + (y − yc )2 = a2 .
In this book, polynomials will be used to define the curve and surface geometries.
Non-rational polynomials can lead to an accurate representation of the geometry despite
the fact that some geometries cannot be described exactly using non-rational polyno-
mials. A circle is an example of a geometry that can only be described exactly using
rational polynomials. One can show that the parametric curve equations x(t) = a cos t
and y(t) = a sin t of a circle with center at the origin can be written using the following
rational polynomials:
( ) ( )
1 − (t)2 2t
x (t) = a , y (t) = a
1 + (t)2 1 + (t)2

Example 2.3 The parametric equation of a helix of radius a like the one shown in
[ ]T
Figure 4 is r (t) = a cos t a sin t 𝛼t , where a and 𝛼 are constants. One can show that
the trace of this curve is a helix of pitch 2𝜋 𝛼 on the cylinder (x)2 + (y)2 = (a)2 (Do Carmo
1976).

Y
a

Figure 2.4 Helix geometry.

Tangent Vector Given two points on the curve defined by the parameter values t1 and t2
as shown in Figure 5, one can define the vector difference Δr = r(t2 ) − r(t1 ). In the limit as
t1 approaches t2 : that is, Δt = t2 − t1 approaches zero, Δr/Δt becomes tangent to the curve.
Therefore, for a given t, the tangent vector to the curve at t is defined as
[ ]
dr dx (t) dy (t) dz (t) T
= (2.2)
dt dt dt dt
Differential Geometry 49

If |dr(t)/dt| = 0 at a point t, the point is called a singular point. The arc length s of a curve
between two arbitrary points to and t can be defined using the norm of the tangent vector as
t | dr |
s= | | dt (2.3)
∫to || dt ||
If a curve is parameterized by its arc length s – that is, r = r(s) – one can show using Eq. 3
that the tangent vector, t(s) = dr/ds, is a unit vector: that is,
| dr |
| | = |t (s)| = 1 (2.4)
| ds |
| |
The proof of Eq. 3 is straightforward and can be established by considering two
points on the curve defined by the parameters t1 and t2 , as shown in Figure 5. As t1
approaches t2 , the infinitesimal arc length ds can be written as (ds)2 = drT dr, where
dr = r(t 2 T 2
√ 2 ) − r(t1 ). Since dr = (dr/dt)dt, one has (ds) = (dr/dt) (dr/dt)(dt) , which yields
T
ds = (dr∕dt) (dr∕dt) dt = |dr∕dt| dt. Equation 3 follows by integrating this equation
from to to t. It is also clear from Eq. 3 that if t = s, then |dr/ds| = 1, which provides a
proof that t(s) = dr/ds is a unit vector. This result is also intuitively obvious since as t1
approaches t2 , it is clear that Δs = |Δr|, which shows that (Δs)2 = ΔrT Δr. This equation
yields (Δr/Δs)T (Δr/Δs) = 1, which shows that the tangent to the curve evaluated by
differentiation with respect to the arc length is indeed a unit vector.

Δr
Δt
t2

t1

r(t1)
r(t2)

Figure 2.5 Parametric derivative.

While any parameter can be used to define the curve, the tangent vector is a unit vector
only when the curve is parameterized by its arc length and the differentiation is carried out
with respect to the arc length. For an arbitrary parameter t, one can write the relationship
50 Mathematical Foundation of Railroad Vehicle Systems

dr/dt = (dr/ds)(ds/dt). Therefore, without any loss of generality, it is assumed in the


remainder of this section that the curve is parameterized by its arc length s.

Example 2.4 The equation of a circle with radius a and center at the origin is defined
[ ]T
in Example 2 by the parametric equation r (t) = a cos t a sin t . The tangent vector is
defined as
dr (t) [ ]T
= −a sin t a cos t
dt
It is clear that the length of this tangent vector is a, and therefore, this vector is not a unit
vector. It is also clear, in this example, that the parameter t represents an angle 𝜃 : that
is, t = 𝜃, and 𝜃 varies from 0 to 2𝜋. Therefore, the length of the circle can be evaluated
using Eq. 3 as
| dr |
t 2𝜋
s= | | dt = ad𝜃 = 2𝜋a
∫to || dt || ∫0
which defines the expected length of the circle perimeter. Furthermore, because
[ ]T
dr/d𝜃 = (dr/ds)(ds/d𝜃), dr∕d𝜃 = −a sin 𝜃 a cos 𝜃 , and dr/ds is the unit tangent
[ ]T
vector dr∕ds = − sin 𝜃 cos 𝜃 , one has (ds/d𝜃) = a or, equivalently, ds = ad𝜃 = adt,
which implies that t = s/a. Substituting this equation in the curve equations and
differentiating with respect to s, one can show that the resulting tangent vector is
indeed a unit vector. The equation ds = ad𝜃, which defines the relationship between
the infinitesimal arc length ds and the angle d𝜃 encompassed by a curve segment
with a radius of curvature a, is general and is applicable to circular and noncircular
curves.

Curvature Vector and Serret–Frenet Equations Assuming that a curve is parame-


terized by its arc length, the curvature vector is obtained by differentiating the unit tangent
vector with respect to the arc length as
d2 r dt
r′′ (s) = = (2.5)
ds2 ds
In this equation, r′′ = d2 r/ds2 denotes twice differentiation with respect to s. Since the tan-
gent vector resulting from differentiation with respect to the arc length is a unit vector, the
tangent and curvature vectors are orthogonal: that is, r′′T t = 0. This can be proven by differ-
entiating the equation r′T r′ = 1 with respect to s, leading to 2r′T r′′ = 0, which demonstrates
the orthogonality of the unit tangent vector and the curvature vector.
The curvature of the curve 𝜅(s) at a point s is a scalar defined as the magnitude of the
curvature vector: that is,

𝜅 (s) = ||r′′ (s)|| = ||t′ (s)|| (2.6)

Because the tangent vector t(s) has unit length, the norm of its derivative defined by
the scalar curvature 𝜅(s) measures the rate of change of the tangent vector orientation.
The radius of curvature at a point s is defined as R(s) = 1/𝜅(s). Because the tangent and
Differential Geometry 51

curvature vectors are orthogonal vectors, the normal to the curve n is defined as a unit
vector along the curvature vector: that is,
r′′ (s) t′ (s)
n (s) = = (2.7)
𝜅 (s) 𝜅 (s)
The osculating plane at a given point s is defined as the plane formed by the unit tangent
and normal vectors. The binormal vector at a point s on the curve is a vector normal to the
osculating plane and is defined by the cross product between the orthogonal tangent and
normal unit vectors as
b (s) = t (s) × n (s) (2.8)
Because t, n, and b are orthogonal unit vectors, they form a frame called the Frenet frame,
[ ]
whose orientation is defined by the orthogonal matrix Af = t n b . The t-b plane is
called the rectifying plane, and the n-b plane is called the normal plane.
Keeping in mind that the two vectors t′ (s) and n(s) are parallel vectors, the derivative of
the binormal vector b with respect to s can be written as
b′ (s) = t′ (s) × n (s) + t (s) × n′ (s) = t (s) × n′ (s) (2.9)
This equation demonstrates that vector b′ (s) is normal to the tangent vector t(s), and
because b(s) is a unit vector, b′ (s) is also normal to b(s); consequently, b′ (s) is parallel
to n. The parallelism of b′ (s) and n can be used to write b′ (s) in the following form, which
defines the torsion 𝜏 of the curve as
b′ (s) = −𝜏 (s) n (s) (2.10)
The curve geometry is completely defined by the curvature and torsion in the neighborhood
of s. In summary, one can write the following equations (Kreyszig 1991; Shabana et al. 2008):

t′ = 𝜅n ⎫

n = −𝜅t + 𝜏b⎬
′ (2.11)
b′ = −𝜏n ⎪

These equations are the Serret–Frenet formulas.

Example 2.5 In Example 4, a circle with radius a and center at the origin defined
[ ]T
by the parametric equations r (t) = a cos t a sin t was considered. It was shown in
Example 4 that t = s/a, where s is the arc length. The circle equation can then be writ-
[ ]T
ten, using s as a parameter, as r (s) = a cos (s∕a) a sin (s∕a) . The tangent vector is
defined as
dr (s) [ ]T
t (s) = = − sin (s∕a) cos (s∕a)
ds
It is clear from this equation that the tangent vector obtained by differentiation with
respect to the arc length s is a unit vector. The curvature vector is defined as
d2 r (s) 1[ ]T
r′′ (s) = t′ (s) = = − cos (s∕a) − sin (s∕a)
ds2 a

(Continued)
52 Mathematical Foundation of Railroad Vehicle Systems

This equation defines the curvature as 𝜅(s) = |r′′ (s)| = 1/a and the normal vector as
[ ]T
n (s) = − cos (s∕a) − sin (s∕a)
The binormal vector is defined as
[ ]T
b (s) = t (s) × n (s) = 0 0 1
This equation shows that b′ (s) = 0; therefore, the torsion 𝜏 = 0 since the circular curve
is planar and is not twisted. The matrix that defines the orientation of the Serret–Frenet
frame at a given point s is defined as

] ⎡
− sin (s∕a) − cos (s∕a) 0⎤
[
Af = t n b = ⎢ cos (s∕a) − sin (s∕a) 0⎥
⎢ ⎥
⎣ 0 0 1⎦
One can show that this matrix is an orthogonal matrix: that is, ATf Af = I, where I is the
3 × 3 identity matrix.

Example 2.6 The parametric equation of the helix in Example 3 is r (t) =


[ ]T
a cos t a sin t 𝛼t , where a and 𝛼 are constants. The tangent vector for this curve
[ ]T √
is r′ (t) = dr∕dt = −a sin t a cos t 𝛼 , which shows that |r′ (t)| = (a)2 + (𝛼)2 .
( √ )
It follows that dt = 1∕ (a)2 + (𝛼)2 ds. Since a and 𝛼 are constants, one has
( √ )
t = 1∕ (a)2 + (𝛼)2 s. The unit tangent vector is, therefore, defined as r′ (s) = dr∕ds =
( √ )[ ]T ( √ )
1∕ (a)2 + (𝛼)2 −a sin t a cos t 𝛼 . Using the equation t = 1∕ (a)2 + (𝛼)2 s,
one can then write
[ ( √ ) ( √ ) ( √ )]T
r (s) = a cos s∕ (a)2 + (𝛼)2 a sin s∕ (a)2 + (𝛼)2 𝛼 s∕ (a)2 + (𝛼)2

which defines the unit tangent vector obtained by differentiation with respect to the arc
length s as
[ ( ) ( ) ]T
1 s s
t (s) = dr∕ds = √ −a sin √ a cos √ 𝛼
(a)2 + (𝛼)2 (a)2 + (𝛼)2 (a)2 + (𝛼)2
The curvature vector is defined as
[ ( ) ( ) ]T
2 a s s
r (s) = d r∕ds = ( 2
′′ 2
) − cos √ − sin √ 0
(a) + (𝛼)2 (a)2 + (𝛼)2 (a)2 + (𝛼)2
a [ ]T
= ( 2 2
) − cos t − sin t 0
(a) + (𝛼)
This equation defines the helix curvature at s as 𝜅(s) = |r′′ (s)| = |a|/((a)2 + (𝛼)2 ), which
shows that the helix has a constant curvature. It follows that the normal vector is
Differential Geometry 53

defined as
[ ( ) ( ) ]T
s s
n (s) = − cos √ − sin √ 0
(a)2 + (𝛼)2 (a)2 + (𝛼)2
[ ]T
= − cos t − sin t 0
The binormal vector is
( )
1
b (s) = t (s) × n (s) = √
(a)2 + (𝛼)2
[ ( ) ( ) ]T
s s
× 𝛼 sin √ −𝛼 cos √ a
(a)2 + (𝛼)2 (a)2 + (𝛼)2
( )
1 [ ]T
= √ 𝛼 sin t −𝛼 cos t a
2 2
(a) + (𝛼)
Differentiating the binormal vector with respect to the arc length, one obtains
( )[ ( ) ( ) ]T
𝛼 s s

b (s) = cos √ sin √ 0
(a)2 + (𝛼)2 (a)2 + (𝛼)2 (a)2 + (𝛼)2
( )
𝛼 [ ]T
= cos t sin t 0
(a)2 + (𝛼)2
which shows that the helix has a constant torsion defined as 𝜏(s) = 𝛼/((a)2 + (𝛼)2 ): that
is, the curvature of the helix can be written in terms of the torsion as 𝜅(s) = (|a|/𝛼)𝜏. The
matrix that defines the orientation of the Serret–Frenet frame at a given point defined
by s can then be written in terms of the curvature and torsion as
√ (√ ) (√ ) √ (√ )
⎡− 𝜅a sin s 𝜏∕𝛼 − cos s 𝜏∕𝛼 𝜏𝛼 sin s 𝜏∕𝛼 ⎤
⎢√ (√ ) (√ ) √ (√ )⎥
Af = ⎢ 𝜅a cos s 𝜏∕𝛼 − sin s 𝜏∕𝛼 − 𝜏𝛼 cos s 𝜏∕𝛼 ⎥
⎢ ⎥
⎢ √ √ ⎥
⎣ 𝜏𝛼 0 a 𝜏∕𝛼 ⎦

Railroad Application of the Theory of Curves The theory of curves is extensively


used in developing the equations that govern the dynamics of railroad vehicle systems. The
profiles of the wheel and rail are first defined using planar curves, as shown in Figure 6.
Using the planar wheel profile curve, the wheel surface of revolution can be constructed, as
will be discussed in a later chapter of this book. Similarly, the rail surface can be extruded
using the planar rail profile. The wheel and rail profiles must be measured frequently in
order to determine the influence of the wear and irregularities on the profile curve charac-
teristics as the result of the wheel/rail dynamic interaction.
54 Mathematical Foundation of Railroad Vehicle Systems

Wheel profile

Rail profile

Figure 2.6 Wheel and rail profiles.

Additionally, as shown in Figure 7, the track is constructed using three different segments
for the longitudinal rail space curve. The first is a straight segment called a tangent, which
has zero curvature; the second is a circular segment called a curve, which has constant cur-
vature; and the third segment, called a spiral, connects the tangent segment to the curve or
connects two curves with different curvature values. The spiral segment, when connecting
tangent and curve segments, is designed such that it starts with zero (or constant) curva-
ture and ends with constant (or zero curvature). The spiral curvature is assumed to vary
linearly, as will be discussed in Chapter 4, in order to ensure continuity of the curvature.
Rail derailments occur when a vehicle negotiates spiral sections that have varying geomet-
ric characteristics and, consequently, the balance speed is not constant during the spiral
negotiation.

Curve

Spiral
tangent-to-curve
Tangent

Figure 2.7 Rail space curve.

2.2 SURFACE GEOMETRY


While in the case of a curve, one can only trace points along one direction, in the case of
a surface, the coordinates of the points can be defined in terms of two independent surface
Differential Geometry 55

s2

∂r
s1
∂s2

∂r
∂s1

Figure 2.8 Surface geometry.

parameters s1 and s2 , as shown in Figure 8. Using these parameters and the Cartesian
coordinates x, y, and z, an arbitrary point on the surface has a unique position defined by
vector r, which can be written in terms of the two independent parameters s1 and s2 as
follows:
( ) [ ( ) ( ) ( )]T
r s1 , s2 = x s1 , s2 y s1 , s2 z s1 , s2 (2.12)

That is, the parameters s1 and s2 completely describe the surface geometry. The parameters
s1 and s2 are called the surface parameters. The surface representation of Eq. 12 is assumed
to satisfy continuity and differentiability requirements. Specifically, the following two
conditions must be met (Kreyszig 1991; Goetz 1970):

1. The mapping in Eq. 12 must be one-to-one: that is, each point on the surface corresponds
to a unique set of the surface parameters s1 and s2 .
2. For a given set of s1 and s2 , the Jacobian matrix

⎡ 𝜕x 𝜕x ⎤
⎢ ⎥
[ ] ⎢ 𝜕s1 𝜕s2 ⎥
𝜕r 𝜕r 𝜕y 𝜕y
J= =⎢ ⎥ (2.13)
𝜕s1 𝜕s2 ⎢ 𝜕s1 𝜕s2 ⎥
⎢ ⎥
⎢ 𝜕z 𝜕z ⎥
⎣ 𝜕s 𝜕s ⎦
1 2

must have a rank of two, which means the two tangent vectors 𝜕r/𝜕s1 and 𝜕r/𝜕s2 are
linearly independent. This implies mathematically that (𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 ) ≠ 0 : that is,
the two tangent vectors are not parallel.

The condition that (𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 ) ≠ 0 is necessary in order to uniquely define the nor-
mal to the surface at an arbitrary point defined by the parameters s1 and s2 . While the two
tangent vectors 𝜕r/𝜕s1 and 𝜕r/𝜕s2 must be linearly independent, these tangent vectors are
56 Mathematical Foundation of Railroad Vehicle Systems

not necessarily orthogonal or unit vectors. The definitions of the two independent tangents
to the coordinate lines s1 and s2 are required for the formulation of the surface fundamental
forms.

Example 2.7 The parametric equations of the cylinder shown in Figure 9, which has
an axis along the X axis and a radius a, are defined in terms of the two parameters
[ ]T [ ]T
x and 𝜃 as r (x, 𝜃) = x y z = x a cos 𝜃 a sin 𝜃 . The two tangent vectors are
defined as
⎡1⎤ ⎡ 0 ⎤
𝜕r ⎢ ⎥ 𝜕r
= 0 , = a ⎢ − sin 𝜃 ⎥
𝜕x ⎢ ⎥ 𝜕𝜃 ⎢ ⎥
⎣0⎦ ⎣ cos 𝜃 ⎦
These two tangent vectors are orthogonal vectors, and therefore, they are independent
for any values of the two parameters x and 𝜃. Note that for the cylinder and for a given x,
one has (y)2 + (z)2 = (a)2 .

r
a X
Y
θ

Figure 2.9 Cylindrical surface.

Example 2.8 The parametric equations of the conic surface shown in Figure 10,
which has an apex at the origin, has an aperture 2𝛾, and opens along the X axis,
[ ]T
are defined in terms of the two parameters x and 𝜃 as r (x, 𝜃) = x y z =
[ ]T
x 1 tan 𝛾 cos 𝜃 tan 𝛾 sin 𝜃 . The two tangent vectors are defined as

⎡ 1 ⎤ ⎡ 0 ⎤
𝜕r ⎢ 𝜕r
= tan 𝛾 cos 𝜃 ⎥ , = x tan 𝛾 ⎢ − sin 𝜃 ⎥
𝜕x ⎢ ⎥ 𝜕𝜃 ⎢ ⎥
⎣ tan 𝛾 sin 𝜃 ⎦ ⎣ cos 𝜃 ⎦
These two tangent vectors are not orthogonal vectors, and singularity occurs at the cone
apex when x = 0. For the conic surface, one has (y)2 + (z)2 = (x tan 𝛾)2 .
Differential Geometry 57

θ
X

Figure 2.10 Conic geometry.

2.3 APPLICATION TO RAILROAD GEOMETRY


In railroad vehicle system dynamics, one of the main steps in solving the wheel/rail
contact problem is to determine online the locations of the contact points on the wheel
and rail surfaces. The surface parameters s1 and s2 are used to define the locations of
these contact points. Nonlinear algebraic contact equations can be formulated and solved
for the surface parameters of the wheel and rail, as will be discussed in this book. These
surface parameters are used in the surface parametric equations to define the location
of the contact points. Knowing the locations of the contact points, the velocities at these
points can be evaluated and used to define velocity variables called creepages that enter
into the formulation of the wheel/rail tangential contact forces. Knowing the surface
parameters also allows for defining the tangent and normal vectors as well as the curva-
tures at the contact points. The principal curvatures and principal directions are determined
using the surface parameters in order to determine the dimensions of the wheel/rail
contact area.
To explain the use of the surface parameterization in railroad vehicle dynamics, the rail
shown in Figure 11a is considered. In railroad vehicle dynamics, the surface geometry
of the rail is defined in a profile coordinate system X rp Y rp Z rp , shown in Figure 11a,
where superscript r refers to the rail and superscript p refers to the profile. For the
purpose of demonstration, the rail is assumed to have a cylindrical surface with radius
a. Since the surface parameterization is not unique, one can use the parameters s1 = x
and s2 = ys , as shown in Figure 11 where ys measures the lateral distance from the origin
of the rail profile frame X rp Y rp Z rp . This x − ys parameterization, often used in railroad
58 Mathematical Foundation of Railroad Vehicle Systems

ys

R
Zrp Zrp
Xrp
θ Xrp
Yrp
Yrp
sr2
sr1

(a) (b)

Figure 2.11 Rail surface.

vehicle dynamics, is slightly different from the parameterization used in Example 7 for
the cylindrical surface. The position of points on a cylindrical surface, translated by the
distance d in the lateral direction, can be defined in the coordinate system X r Y r Z r as
[ ]T [ √ ( )2 T
] [ ]T
r = x y z = x −d + ys (a)2 − ys , which can be written as r = x y z =
[ ]T √ ( )2
x −d + ys f r (y) , where f r (y) = (a)2 − ys defines the profile geometry in the pro-
file coordinate system X rp Y rp Z rp . The rail profile can also change as a function of the param-
eter x that represents the distance traveled. In this case, one can use the function f r = f r (x, ys )
to account for the change in the profile shape as a function of the parameter x. Therefore,
for a rail with arbitrary geometry, the following general representation can be used:
[ ]T [ ( )]T
r = x y z = x y f r x, ys (2.14)
Keeping in mind that y = −d + ys , in the general case of Eq. 14, the tangent vectors are
defined as
⎡ 1 ⎤ ⎡ 0 ⎤
𝜕r 𝜕r ⎢ 𝜕r 𝜕r
= = 0 ⎥, = =⎢ 1 ⎥ (2.15)
𝜕s1 𝜕x ⎢ r ⎥ 𝜕s2 𝜕ys ⎢ r ⎥
⎣𝜕f ∕𝜕x⎦ ⎣𝜕f ∕𝜕ys ⎦
It is clear from this equation that if the profile function f r does not depend on the parameter
x, using x and ys as surface parameters leads to orthogonal tangents. It is also clear that in the
r
( ) √ 2 ( )2
special case of a cylindrical rail in which f ys = (a) − ys , the tangent and normal
[ ( √ )]
[ ]T ( )2 T
vectors are defined as 𝜕r∕𝜕x = 1 0 0 , 𝜕r∕𝜕ys = 0 1 −ys ∕ (a)2 − ys , and
[ ( √ ) ]T
( )2
n = 0 ys ∕ (a)2 − ys 1 .
In the case of curved rail, the longitudinal parameter, shown in Figure 11b, can be con-
sidered as the arc length s of the rail space curve, and therefore, the two parameters s and
ys can be used instead of x and ys . In this case, one can write Eq. 14 as
[ ]T [ ( ) ( ) ( )]T
r = x y z = g1r s, ys g2r s, ys f r s, ys (2.16)
Differential Geometry 59

( ) ( )
In this equation, g1r s, ys and g2r s, ys are functions that, respectively, define the Cartesian
coordinates x and y of points on the rail surface in terms of the surface parameters s
( )
and ys . For example, in the case of a circular rail space curve, the functions g1r s, ys
( )
and g2r s, ys are defined, respectively, as g1r (s) = ds sin (s∕R) and g2r (s) = −ds cos (s∕R),
where ds = − R + ys and R is the radius of curvature of the rail space curve. Using simple
geometry, the angle 𝜃 shown in Figure 11b is defined as 𝜃 = s/R. Therefore, for a circular
rail space curve with a radius of curvature R and a cylindrical profile with a radius a,
[ ]T [ √ ( )2 T
]
one has r = x y z = ds sin (s∕R) −ds cos (s∕R) (a)2 − ys . It follows that
[( ) ( ) ]T
the tangent vectors are 𝜕r∕𝜕s = ds ∕R cos (s∕R) ds ∕R sin (s∕R) 0 and 𝜕r∕𝜕ys =
[ ( √ )]
( )2 T
sin (s∕R) − cos (s∕R) − ys ∕ (a)2 − ys . Using these two tangent vectors, the
normal vector can be defined using the cross product (𝜕r/𝜕s) × (𝜕r/𝜕ys ).
A similar procedure can be used for the definition of the parametric form of the wheel
surface. The surface of an unworn wheel is a surface of revolution generated by rotating the
wheel profile curve about the wheel axle. To define the wheel surface geometry in a para-
metric form, a coordinate system X w Y w Z w is assigned to the wheel, where superscript w
refers to the wheel. This coordinate system is rigidly attached to the wheel axle, and there-
fore, it rotates with the wheel. The wheel surface parametric form can be defined by the
lateral surface parameter sw1 = ys and the angular surface parameter sw2 = 𝜃. The angle 𝜃,
measured here from the Z w axis, as shown in Figure 12, is a surface parameter and has noth-
ing to do with the angle that defines the wheel rotation as it rolls. As shown in Figure 12a,
the wheel profile curve can, in general, be defined by the function f w (ys , 𝜃) in a wheel profile
frame X wp Y wp Z wp using the surface parameters ys and 𝜃. Using this function, and assum-
ing that the origin of the wheel profile coordinate system X wp Y wp Z wp is defined in a wheel
coordinate system by the coordinates xo , yo , and zo , the parametric form that defines the
coordinates of the points on the wheel surface can be written as
[ ]T [ ( ) ( )]T
r = x y z = g1w ys , 𝜃 yo + ys g2w ys , 𝜃 (2.17)
( ) ( ) ( ) ( )
In this equation, g1w ys , 𝜃 = xo + f w ys , 𝜃 sin 𝜃 and g2w ys , 𝜃 = zo + f w ys , 𝜃 cos 𝜃
are functions that, respectively, define the Cartesian coordinates x and z of points on
the wheel surface in terms of the surface parameters ys and 𝜃. The parametric form

Zwp Zwp
a
Ywp
θ

ys
Xwp

R
Xwp

(a) (b)

Figure 2.12 Wheel surface geometry.


60 Mathematical Foundation of Railroad Vehicle Systems

of Eq. 17 is general and allows for changing the profile geometry as a function of the
angular surface parameter 𝜃. In the special case of a wheel with radius R and cylindri-
cal profile of radius a, as shown in Figure 12b, one can select the origin of the profile
( ) ( )
coordinate system such that xo = R sin 𝜃, yo = 0, zo = R cos 𝜃, and f w ys , 𝜃 = f w ys =
√ ( )2 [ ]T [ ]T
(a)2 − ys . It follows that r = x y z = (R + f w ) sin 𝜃 ys (R + f w ) cos 𝜃 .
[( ) ( ) ]T
The two tangent vectors are 𝜕r∕𝜕ys = 𝜕f w ∕𝜕ys sin 𝜃 1 𝜕f w ∕𝜕ys cos 𝜃 and
[ ] √ ( )
T 2
𝜕r∕𝜕𝜃 = (R + f w ) cos 𝜃 0 − sin 𝜃 , where 𝜕f w ∕𝜕ys = −ys ∕ (a)2 − ys .
The profiles of the wheel and rail in railroad vehicle system applications are designed
using more complex geometry as compared to the simple cylindrical profiles considered
in the examples used in this section. The actual profiles can be measured using a device
called a MiniProf, which generates the ys − z table of spline data that can be used in
dynamic simulations. Using spline functions to define the wheel and rail profiles, which
allows for describing arbitrary profile geometries, and using the absolute nodal coordinate
formulation for the definition of the rail geometry, will be discussed in later sections of this
chapter.

2.4 SURFACE TANGENT PLANE AND NORMAL


VECTOR
The tangent plane to the surface at a point P is the plane defined by the two tangent vectors
t1 = 𝜕r/𝜕s1 and t2 = 𝜕r/𝜕s2 , where s1 and s2 are the surface parameters that define point P.
A curve can be defined on the surface, as shown in Figure 13, by assuming that the two
surface parameters s1 and s2 can be written in terms of one parameter t. In this case, one
can write s1 = s1 (t) and s2 = s2 (t), where t is the curve parameter: that is, s1 and s2 are related
and are no longer independent. Using this curve parameterization, Eq. 12 can be written as
( )
r s1 (t) , s2 (t) = y (t) (2.18)
where y(t) is a regular curve on the surface that has a tangent vector defined by the equation
dy 𝜕r ds1 𝜕r ds2
= + (2.19)
dt 𝜕s1 dt 𝜕s2 dt

t2
dy
dt
P
t1

y = y(t)

Figure 2.13 Tangent plane.


Differential Geometry 61

If (𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 ) ≠ 0 and ds1 /dt and ds2 /dt are assumed to be different from zero – that is,
(ds1 /dt) ≠ 0 and/or (ds2 /dt) ≠ 0 – the curve tangent vector of Eq. 19 is well defined. It is clear
from Eq. 19 that the curve tangent vector dy(t)/dt is a linear combination of the two surface
tangent vectors 𝜕r/𝜕s1 and 𝜕r/𝜕s2 , and therefore, dy(t)/dt lies in the tangent plane and is
also tangent to the surface, as shown in Figure 13. The unit vector normal to the surface at
point P is defined as the normal to the tangent plane and can be written as follows:
(( ) ( ))
t1 × t 2 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2
n= | | = |( ) ( )| (2.20)
| t1 × t2 | | 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2 |
| |
This vector is, therefore, normal to the tangent vectors 𝜕r/𝜕s1 and 𝜕r/𝜕s2 as well as dy(t)/dt
at point P. For example, for general rail geometry, the parametric representation of the
[ ]T [ ( )]T
surface was found to be r = x y z = x y f r x, ys (Eq. 14). The two tangent
[ ]T
vectors are defined in Eq. 15 as 𝜕r∕𝜕s1 = 𝜕r∕𝜕x = 1 0 𝜕f ∕𝜕x and 𝜕r∕𝜕s2 = 𝜕r∕𝜕ys =
r
[ ]T ( ) ( ) [
0 1 𝜕f r ∕𝜕ys . In this case, the normal vector is 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2 = −𝜕f r ∕𝜕x
]T
−𝜕f r ∕𝜕ys 1 . Therefore, the unit normal to the surface at a point defined by the surface
( ) ( ) |( ) ( )|
parameters s1 = x and s2 = ys is given by n = 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2 ∕ | 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2 | =
√ | |
[ ]T 2 ( )2
(1∕𝛼) −𝜕f ∕𝜕x −𝜕f ∕𝜕ys 1 , where 𝛼 = |n| = (𝜕f ∕𝜕x) + 𝜕f ∕𝜕ys + 1.
r r r r

Example 2.9 The parametric equation of the cylindrical surface in Example 7, which
has an axis along the X axis and radius a, is defined in terms of the two parameters
[ ]T [ ]T
x and 𝜃 as r (x, 𝜃) = x y z = x a cos 𝜃 a sin 𝜃 . The two tangent vectors were
[ ]T [ ]T
defined as 𝜕r∕𝜕x = 1 0 0 and 𝜕r∕𝜕𝜃 = a 0 − sin 𝜃 cos 𝜃 . The unit normal
[ ]T
vector is defined as n = 0 − cos 𝜃 − sin 𝜃 .

Example 2.10 It was shown in Example 8 that the parametric equation of the conic
surface shown in Figure 10, which opens along the X axis, is defined in terms of
[ ]T [ ]T
the two parameters x and 𝜃 by r (x, 𝜃) = x y z = x 1 tan 𝛾 cos 𝜃 tan 𝛾 sin 𝜃 .
[ ]T
The two tangent vectors were defined as 𝜕r∕𝜕x = 1 tan 𝛾 cos 𝜃 tan 𝛾 sin 𝜃 and
[ ]T
𝜕r∕𝜕𝜃 = x 0 − tan 𝛾 sin 𝜃 tan 𝛾 cos 𝜃 . The normal vector is defined as (𝜕r∕𝜕x) ×
[ ]T
(𝜕r∕𝜕𝜃) = x tan 𝛾 tan 𝛾 − cos 𝜃 − sin 𝜃 . The norm of this vector is x tan 𝛾/ cos 𝛾,
[ ]T
which shows that the unit normal is given by n = cos 𝛾 tan 𝛾 − cos 𝜃 − sin 𝜃 . It is
clear that there is a singularity at x = 0, where the normal vector is not defined.

The definition of the unit normal vector to the wheel and rail surfaces at the point of
contact is necessary in order to be able to determine the direction of the normal contact
force. The normal contact force, relative velocity, and wheel and rail material properties
and surface geometry at the contact point are used to determine the tangential creep and
friction forces, which play a detrimental role in railroad vehicle dynamics and stability.
A surface, as in the case of a curve, can be defined using local properties like those deter-
mined by differentiating the curve equations with respect to the surface parameters. Gauss
62 Mathematical Foundation of Railroad Vehicle Systems

introduced two forms that can be used to determine the local surface properties, which are
discussed next: the first and second fundamental forms of the surfaces.

2.5 SURFACE FUNDAMENTAL FORMS

This section introduces the first and second fundamental forms of surfaces. These funda-
mental forms allow measuring distances and defining surface curvatures. The coefficients
of the surface fundamental forms are used to define the principal curvatures and principal
directions that enter into the formulation of wheel/rail interaction forces.

First Fundamental Form The first fundamental form of a surface is defined as


I = dr ⋅ dr = drT dr (2.21)
Because dr = (𝜕r/𝜕s1 )ds1 + (𝜕r/𝜕s2 )ds2 , the first fundamental form I can be written as
(( ) ( ) )T (( ) ( ) )
I = 𝜕r∕𝜕s1 ds1 + 𝜕r∕𝜕s2 ds2 𝜕r∕𝜕s1 ds1 + 𝜕r∕𝜕s2 ds2 (2.22)
which can be written as
( )2 ( )2
I = E ds1 + 2Fds1 ds2 + G ds2
[ ][ ]
[ ] E F ds1
= ds1 ds2 (2.23)
F G ds2
where E, F, and G are called the coefficients of the first fundamental form and are defined as
( )T ( )
E = tT1 t1 = 𝜕r∕𝜕s1 𝜕r∕𝜕s1 ,⎫
( )T ( ) ⎪
F = tT1 t2 = 𝜕r∕𝜕s1 𝜕r∕𝜕s2 ,⎬ (2.24)
( )T ( )⎪
G = tT2 t2 = 𝜕r∕𝜕s2 𝜕r∕𝜕s2 ⎭
The first fundamental form of Eq. 23, which is a homogenous function of second degree
in ds1 and ds2 , can be used to measure distances, angles, and areas on the surface. To
demonstrate this, consider a curve on the surface defined by the parametric equation
y(t) = r(s1 (t), s2 (t)), where t is the curve parameter. The length of the curve y(t) = r(s1 (t),
t t √
s2 (t)) over the domain t1 ≤ t ≤ t2 is given by l = ∫t 2 |dy∕dt| dt = ∫t 2 |(dr∕dt) ⋅ (dr∕dt)|dt,
1 1
which√ can(
be written in terms of the coefficients of the first fundamental form I as l =
) ( ) ( ) (( ))
t 2 2
∫t 2 E ds1 ∕dt + 2F ds1 ∕dt ds2 ∕dt + G ds2 ∕dt dt, which shows that the length of
1
a curve can be evaluated using the square root of the first fundamental form. Furthermore,
the angle between two tangent vectors t1 = 𝜕r/𝜕s1 and t2 = 𝜕r/𝜕s2 is defined by the equation
( ) ( )T ( ) ( )
cos 𝛽 = tT1 t2 ∕ ||t1 || ||t2 || = 𝜕r∕𝜕s1 𝜕r∕𝜕s2 ∕ ||𝜕r∕𝜕s1 || ||𝜕r∕𝜕s2 || , which can be written in

terms of the coefficients of the first fundamental form as cos 𝛽 = F∕ EG. Clearly, the two
tangent vectors t1 = 𝜕r/𝜕s1 and t2 = 𝜕r/𝜕s2 are orthogonal at a point if F = 0 at that point.
It was previously shown in this chapter that in the general case in which the rail
profile changes as a function of the longitudinal surface parameter x that represents
the distance traveled, the rail profile curve can be written as f r = f r (x, ys ), where ys is
the lateral rail surface parameter. In this case, the surface parametric equation of a rail
[ ]T [ ( )]T
with arbitrary geometry can, in general, be written as r = x y z = x y f r x, ys .
While it is assumed here that y = ys , one can also assume that y = yo + ys , where yo is a
Differential Geometry 63

constant shift. The longitudinal surface parameter x is often chosen to be the arc length
of the rail space curve. Using the parametric form of the rail surface, it was shown
[ ]T
that the two tangent vectors are given by t1 = 𝜕r∕𝜕s1 = 𝜕r∕𝜕x = 1 0 𝜕f r ∕𝜕x and
[ ]T
t2 = 𝜕r∕𝜕s2 = 𝜕r∕𝜕ys = 0 1 𝜕f r ∕𝜕ys . One can then show that for this rail surface
description, the coefficients of the first fundamental form are
E = tT1 t1 = (𝜕r∕𝜕x)T (𝜕r∕𝜕x) = 1 + (𝜕f r ∕𝜕x)2 , ⎫
( ) ( ) ⎪
F = tT1 t2 = (𝜕r∕𝜕x) 𝜕r∕𝜕ys = (𝜕f r ∕𝜕x) 𝜕f r ∕𝜕ys ,⎬
T (2.25)
( )T ( ) ( )2 ⎪
G = tT2 t2 = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + 𝜕f r ∕𝜕ys ⎭
If the profile function f r does not vary as a function of the longitudinal surface parameter
x – that is, f r = f r (ys ) – the two tangent vectors are orthogonal vectors since F = 0, which
( ) ( )T ( ) ( )
shows that cos 𝛽 = tT1 t2 ∕ ||t1 || ||t2 || = 𝜕r∕𝜕s1 𝜕r∕𝜕s2 ∕ ||𝜕r∕𝜕s1 || ||𝜕r∕𝜕s2 || = 0.

( )
Example 2.11 In the special case of a straight cylindrical rail in which f r ys =
√ ( )2
(a)2 − ys , where a is the cylinder radius, the two tangent vectors are defined as
[ ( √ )]
[ ]T ( )2 T
𝜕r∕𝜕x = 1 0 0 , and 𝜕r∕𝜕ys = 0 1 2
−ys ∕ (a) − ys . In this case, the
coefficients of the first fundamental form are given by
( )
E = tT1 t1 = (𝜕r∕𝜕x)T (𝜕r∕𝜕x) = 1, F = tT1 t2 = (𝜕r∕𝜕x)T 𝜕r∕𝜕ys = 0,
( )T ( ) (( )2 ( ( )2 ))
G = tT2 t2 = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + ys ∕ (a)2 − ys

As previously discussed, the surface of the wheel is described by the profile curve
f w (ys , 𝜃), defined in the profile frame X wp Y wp Z wp , where ys and 𝜃 are, respectively,
the wheel lateral and angular surface parameters. Assuming that the origin of the
wheel profile coordinate system X wp Y wp Z wp is defined in the X w Y w Z w coordinate sys-
tem by the coordinates xo , yo , and zo , the parametric form of the wheel surface is r =
[ ]T [ ( ) ( )]T ( ) ( )
x y z = g1w ys , 𝜃 yo + ys g2w ys , 𝜃 , where g1w ys , 𝜃 = xo + f w ys , 𝜃 sin 𝜃 and
( ) ( ) [
g2w ys , 𝜃 = zo + f w ys , 𝜃 cos 𝜃. The two tangent vectors are defined as 𝜕r∕𝜕ys = 𝜕g1w ∕𝜕ys
]T [ ]T
1 𝜕g2w ∕𝜕ys and 𝜕r∕𝜕𝜃 = 𝜕g1w ∕𝜕𝜃 0 𝜕g2w ∕𝜕𝜃 . The coefficients of the first fundamental
form are defined in this case as
( )T ( ) ( )2 ( )2
E = tT1 t1 = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + 𝜕g1w ∕𝜕ys + 𝜕g2w ∕𝜕ys , ⎫
( )T ( ) ( ) ( )( w ) ⎪
F = t1 t2 = 𝜕r∕𝜕ys (𝜕r∕𝜕𝜃) = 𝜕g1 ∕𝜕ys 𝜕g1 ∕𝜕𝜃 + 𝜕g2 ∕𝜕ys 𝜕g2 ∕𝜕𝜃 ,⎬ (2.26)
T w w w
( )2 ( )2 ⎪
G = tT2 t2 = (𝜕r∕𝜕𝜃)T (𝜕r∕𝜕𝜃) = 𝜕g1w ∕𝜕𝜃 + 𝜕g2w ∕𝜕𝜃 ⎭
The condition of the orthogonality of the two vectors tangent to the wheel surface is F = 0,
( )( ) ( )( ) ( )
which implies that 𝜕g1w ∕𝜕ys 𝜕g1w ∕𝜕𝜃 + 𝜕g2w ∕𝜕ys 𝜕g2w ∕𝜕𝜃 = 0: that is, 𝜕g1w ∕𝜕ys
( w ) ( w )( w )
𝜕g1 ∕𝜕𝜃 = − 𝜕g2 ∕𝜕ys 𝜕g2 ∕𝜕𝜃 . If the wheel profile is not a function of the angular
( )( )
surface parameter 𝜃, one has f w = f w (ys ). In this special case, one has 𝜕g1w ∕𝜕ys 𝜕g1w ∕𝜕𝜃 =
( ) ( )( ) ( )
f w 𝜕f w ∕𝜕ys cos 𝜃 sin 𝜃 and 𝜕g2w ∕𝜕ys 𝜕g2w ∕𝜕𝜃 = −f w 𝜕f w ∕𝜕ys cos 𝜃 sin 𝜃, demonstrat-
ing that the two tangent vectors at an arbitrary point defined by the parameters ys and 𝜃
remain orthogonal vectors. This is, in general, the case for unworn wheels. In this case,
the coefficients of the first fundamental form reduce to
( )2 ( )2
E = 1 + 𝜕f w ∕𝜕ys , F = 0, G = f w (2.27)
64 Mathematical Foundation of Railroad Vehicle Systems

These coefficients have much simpler forms as compared to the more general expressions
given by Eq. 26.

Example 2.12 In the special case of the wheel shown in Figure 12b, which
has radius R and cylindrical profile of radius a, the origin of the profile coordi-
nate system X wp Y wp Z wp is defined in the wheel coordinate system X w Y w Z w by
( ) ( )
xo = R sin 𝜃, yo = 0, zo = R cos 𝜃. The profile function is defined as f w ys , 𝜃 = f w ys =
√ ( )2 [ ]T
(a)2 − ys , and the parametric form of the surface is defined as r = x y z =
[ ]T
(R + f w ) sin 𝜃 ys (R + f w ) cos 𝜃 . The two tangent vectors are t1 = 𝜕r∕𝜕ys =
[( w ) ( w ) ]T [ ]T
𝜕f ∕𝜕ys sin 𝜃 1 𝜕f ∕𝜕ys cos 𝜃 and t2 = 𝜕r∕𝜕𝜃 = (R + f w ) cos 𝜃 0 − sin 𝜃 ,
√ ( )2
where 𝜕f w ∕𝜕ys = −ys ∕ (a)2 − ys . The coefficients of the first fundamental form are
defined as
( )T ( ) ( )2 ( )2 ( ( )2 )
E = tT1 t1 = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + 𝜕f w ∕𝜕ys = 1 + ys ∕ (a)2 − ys ,
( ) T
F = tT1 t2 = 𝜕r∕𝜕ys (𝜕r∕𝜕𝜃) = 0,
( )2
G = tT2 t2 = (𝜕r∕𝜕𝜃)T (𝜕r∕𝜕𝜃) = R + f w

Example 2.13 The parametric equation of the conical wheel surface shown in
Figure 14, which is assumed to have aperture 2𝛾 and opens along the Y , axis
( ) [ ]T
are defined in terms of the two parameters ys and 𝜃 as r ys , 𝜃 = x y z =
[ ]T
ys tan 𝛾 sin 𝜃 ys ys tan 𝛾 cos 𝜃 . The two tangent vectors are defined as t1 = 𝜕r∕𝜕ys =
[ ]T [ ]T
tan 𝛾 sin 𝜃 1 tan 𝛾 cos 𝜃 and t2 = 𝜕r∕𝜕𝜃 = ys tan 𝛾 cos 𝜃 0 − tan 𝛾 sin 𝜃 . It is
( ) T
clear that these two tangent vectors are orthogonal because tT1 t2 = 𝜕r∕𝜕ys (𝜕r∕𝜕𝜃) =
0. The coefficients of the first fundamental form of this conical surface are given by
( )T ( )
E = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + tan2 𝛾,
( )T
F = 𝜕r∕𝜕ys (𝜕r∕𝜕𝜃) = 0,
( )2
G = (𝜕r∕𝜕𝜃)T (𝜕r∕𝜕𝜃) = ys tan 𝛾

θ
X

Y
ys

Figure 2.14 Conical wheel surface.


Differential Geometry 65

Second Fundamental Form The vector n normal to the surface was defined as
n = t1 × t2 /|t1 × t2 | = ((𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 ))/|(𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 )|. Therefore, the unit normal
vector n is a function in the surface parameters s1 and s2 , and its differential can be written as
( ) ( )
dn = 𝜕n∕𝜕s1 ds1 + 𝜕n∕𝜕s2 ds2 (2.28)

Because dr = (𝜕r/𝜕s1 )ds1 + (𝜕r/𝜕s2 )ds2 is tangent to the surface, one has dr ⋅ n = 0, which
( ) ( )2
upon differentiation yields d2 r ⋅ n = − dr ⋅ dn. In this equation, d2 r = 𝜕 2 r∕𝜕s21 ds1 +
( ) ( ) ( )2
2 𝜕 2 r∕𝜕s1 𝜕s2 ds1 ds2 + 𝜕 2 r∕𝜕s22 ds2 . The second fundamental form of a surface is
defined as
II = d2 r ⋅ n = −dr ⋅ dn
(( ) ( ) )T (( ) ( ) )
= − 𝜕r∕𝜕s1 ds1 + 𝜕r∕𝜕s2 ds2 𝜕n∕𝜕s1 ds1 + 𝜕n∕𝜕s2 ds2 (2.29)

This equation shows that the second fundamental form defines the projection of the second
derivative of the vector that represents the relative position of two neighboring points on
the surface normal vector. The second fundamental form of Eq. 29 can also be written as
( )2 ( )2
II = L ds1 + 2Mds1 ds2 + N ds2
[ ][ ]
[ ] L M ds1
= ds1 ds2 (2.30)
M N ds2
where L, M, and N are called the coefficients of the second fundamental form, defined as
( )T ( ) ⎫
L = − 𝜕r∕𝜕s1 𝜕n∕𝜕s1 , ⎪
(( )T ( ) ( )T ( )
) ⎪
1
M=− 𝜕r∕𝜕s1 𝜕n∕𝜕s2 + 𝜕r∕𝜕s2 𝜕n∕𝜕s1 ,⎬ (2.31)
2 ⎪
( )T ( )
N = − 𝜕r∕𝜕s2 𝜕n∕𝜕s2 ⎪

Equation 29 or 30 shows that the second fundamental form II is a homogenous function
of second degree in ds1 and ds2 . Because the normal vector n is orthogonal to the tangent
plane formed by vectors t1 = 𝜕r/𝜕s1 and t2 = 𝜕r/𝜕s2 , the following identities can be written:

𝜕 (T ) ( 2 )T ( )T ( ) ⎫
t1 n = 𝜕 r∕𝜕s21 n + 𝜕r∕𝜕s1 𝜕n∕𝜕s1 = 0 ⎪
𝜕s1 ⎪
𝜕 (T ) ( 2 )T ( )T ( )
t1 n = 𝜕 r∕𝜕s1 𝜕s2 n + 𝜕r∕𝜕s1 𝜕n∕𝜕s2 = 0⎪
𝜕s2 ⎪
𝜕 (T ) ( 2 )T ( )T ( ) ⎬ (2.32)
t2 n = 𝜕 r∕𝜕s2 𝜕s1 n + 𝜕r∕𝜕s2 𝜕n∕𝜕s1 = 0⎪
𝜕s1 ⎪
𝜕 (T ) ( 2 )
2 T
( )T ( ) ⎪
t2 n = 𝜕 r∕𝜕s2 n + 𝜕r∕𝜕s2 𝜕n∕𝜕s2 = 0 ⎪
𝜕s2 ⎭
These identities can also be written in the following forms:
( 2 )T ( )T ( )
𝜕 r∕𝜕s21 n = − 𝜕r∕𝜕s1 𝜕n∕𝜕s1 , ⎫
( 2 )T ( )T ( ) ⎪
𝜕 r∕𝜕s1 𝜕s2 n = − 𝜕r∕𝜕s1 𝜕n∕𝜕s2 ,⎪
( 2 )T ( )T ( ) ⎬ (2.33)
𝜕 r∕𝜕s1 𝜕s2 n = − 𝜕r∕𝜕s2 𝜕n∕𝜕s1 ,⎪
( 2 )T ( )T ( ) ⎪
𝜕 r∕𝜕s22 n = − 𝜕r∕𝜕s2 𝜕n∕𝜕s2 ⎭
66 Mathematical Foundation of Railroad Vehicle Systems

Using these identities, the coefficients of the second fundamental form can be written in
alternate forms in terms of second derivatives as
( )T ( )T ( )T
L = 𝜕 2 r∕𝜕s21 n, M = 𝜕 2 r∕𝜕s1 𝜕s2 n, N = 𝜕 2 r∕𝜕s22 n (2.34)
As will be explained in this chapter, 𝜕 2 r/𝜕si 𝜕sj , i, j = 1, 2 represent curvature vectors, and
therefore, the coefficients of the second fundamental form represent the projections or
components of these curvature vectors at a point along the normal to the surface at this
point. These coefficients of the second fundamental form determine the nature of the
surface in the neighborhood of a point. Based on the values of these coefficients, one has
the following surface classifications:
(1) If LN − (M)2 > 0 at a point, the surface is called elliptic at that point.
(2) If LN − (M)2 < 0 at a point, the surface is called hyperbolic at that point.
(3) If LN − (M)2 = 0 at a point, the surface is called parabolic at that point.
(4) If L = N = M = 0 at a point, the surface is called planar at that point.
Figure 15 shows examples of elliptic, hyperbolic, parabolic, and planar surfaces.

Elliptic Hyperbolic
(a) (b)

Parabolic Planar
(c) (d)

Figure 2.15 Second fundamental form of surfaces.

( )
Example 2.14 In the special case of a straight cylindrical rail in which f r ys =
√ ( )2
(a)2 − ys , where a is the cylinder radius, the two tangent vectors are defined in
[ ( √ )]
[ ]T ( )2 T
Example 11 as t1 = 𝜕r∕𝜕x = 1 0 0 and t2 = 𝜕r∕𝜕ys = 0 1 −ys ∕ (a)2 − ys .
( √ )
( )2 [ ]T
In this case, the normal vector is defined as n = 1∕ 1 + fyrs 0 −fyrs 1 , where
Differential Geometry 67

√ ( )2
fyrs = −ys ∕ (a)2 − ys . The following vectors can be defined:
[ ]T
𝜕 2 r∕𝜕s21 = 𝜕 2 r∕𝜕x2 = 0 0 0
[ ]T
𝜕 2 r∕𝜕s1 𝜕s2 = 𝜕 2 r∕𝜕x𝜕ys = 0 0 0
[ ]T
𝜕 2 r∕𝜕s22 = 𝜕 2 r∕𝜕y2s = 0 0 𝜕fyrs ∕𝜕ys
( ( )2 )3∕2
where 𝜕fyrs ∕𝜕ys = −(a)2 ∕ (a)2 − ys . One can, therefore, define the coefficients of
the second fundamental form as
( )T
L = 𝜕 2 r∕𝜕s21 n = 0,
( )T
M = 𝜕 2 r∕𝜕s1 𝜕s2 n = 0,

( 2 ) ( ) ( )2
2 T
N = 𝜕 r∕𝜕s2 n = 𝜕fys ∕𝜕ys ∕ 1 + fyrs
r

To give a geometric interpretation of the coefficients of the second fundamental form, the
straight cylindrical rail of the preceding example is considered. The surface parametriza-
tion can be changed from x and ys to x and s, where s is the arc length of the profile curve,
as shown in Figure 16. Using this parametrization, the equation of the cylindrical rail in its
[ ]T
parametric form can be written as r = x a cos (s∕a) a sin (s∕a) . The two tangent vec-
[ ]T [ ]T
tors are defined as t1 = 𝜕r∕𝜕x = 1 0 0 and t2 = 𝜕r∕𝜕s = 0 − sin (s∕a) cos (s∕a) .
[ ]T
The normal vector is defined as n = (𝜕r∕𝜕x) × (𝜕r∕𝜕s) = 0 − cos (s∕a) − sin (s∕a) .
The second derivatives of vector r with respect to the new parameters are 𝜕 2 r∕𝜕x2 =
[ ]T [ ]T [ ]T
0 0 0 , 𝜕 2 r∕𝜕x𝜕s = 0 0 0 , and 𝜕 2 r∕𝜕s2 = (1∕a) 0 − cos (s∕a) − sin (s∕a) .
Using these definitions, the coefficients of the second fundamental form, when using the
arc length parameter for the profile curve, are L = (𝜕 2 r/𝜕x2 )T n = 0, M = (𝜕 2 r/𝜕x𝜕s)T n = 0,
and N = (𝜕 2 r/𝜕s2 )T n = 1/a, which shows that in this case, the coefficient N represents
the constant curvature of the rail profile circular curve. Note also that in this example,
[ ]T
the curvature vector 𝜕 2 r∕𝜕s2 = (1∕a) 0 − cos (s∕a) − sin (s∕a) is parallel to the unit
[ ]T
normal vector to the surface n = 0 − cos (s∕a) − sin (s∕a) .

s
x

Zrp
Xrp

Yrp

Figure 2.16 Rail surface.


68 Mathematical Foundation of Railroad Vehicle Systems

Example 2.15 In the special case of the wheel shown in Figure 12b, which
has radius R and cylindrical profile of radius a, the origin of the profile coor-
dinate system X wp Y wp Z wp is defined in the wheel coordinate system X w Y w Z w
( )
by xo = R sin 𝜃, yo = 0, zo = R cos 𝜃. The profile function is defined as f w ys , 𝜃 =
( ) √ ( )2
f w ys = (a)2 − ys , and the parametric form of the surface is defined as
[ ]T [ ]T
r = x y z = (R + f w ) sin 𝜃 ys (R + f w ) cos 𝜃 . The two tangent vectors were
[( ) ( ) ]T
defined in Example 12 as t1 = 𝜕r∕𝜕ys = 𝜕f w ∕𝜕ys sin 𝜃 1 𝜕f w ∕𝜕ys cos 𝜃 and
[ ]T √ ( )2
t2 = 𝜕r∕𝜕𝜃 = (R + f w ) cos 𝜃 0 − sin 𝜃 , where fyws = 𝜕f w ∕𝜕ys = −ys ∕ (a)2 − ys .
( ) |( ) |
The normal vector is defined as n = 𝜕r∕𝜕ys × (𝜕r∕𝜕𝜃) ∕ | 𝜕r∕𝜕ys × (𝜕r∕𝜕𝜃)| =
√ | |
[ ]T ( )2
𝛽 − sin 𝜃 −fys − cos 𝜃 , where 𝛽 = (R + f ) ∕ 1 + fys . The second derivatives of
w w w
( )[ ]T ( )[ ]T
r are 𝜕 2 r∕𝜕y2s = 𝜕fyws ∕𝜕ys sin 𝜃 0 cos 𝜃 , 𝜕 2 r∕𝜕ys 𝜕𝜃 = − 𝜕f w ∕𝜕ys cos 𝜃 0 − sin 𝜃 ,
[ ]T
and 𝜕 2 r∕𝜕𝜃 2 = − (R + f w ) sin 𝜃 0 cos 𝜃 . Therefore, the coefficients of the second
fundamental form are
( ) ( )
L = 𝜕 2 r∕𝜕y2s ⋅ n = −𝛽 𝜕fyws ∕𝜕ys ,
( )
M = 𝜕 2 r∕𝜕ys 𝜕𝜃 ⋅ n = 0,
( ) ( )
N = 𝜕 2 r∕𝜕𝜃 2 ⋅ n = 𝛽 R + f w

As in the case of the cylindrical rail, one can parameterize the cylindrical wheel profile
using the arc length s. In this case, the parametric equation of the cylindrical wheel
[ ]T [
surface can be written in terms of the surface parameters s and 𝜃 as r = x y z = (R+
]T
a sin (s∕a)) sin 𝜃 a cos (s∕a) (R + a sin (s∕a)) cos 𝜃 . The two tangent vectors can be
[ ]T
defined in this case as t1 = 𝜕r∕𝜕s = cos (s∕a) sin 𝜃 − sin (s∕a) cos (s∕a) cos 𝜃 and t2 =
[ ]T
𝜕r∕𝜕𝜃 = (R + a sin (s∕a)) cos 𝜃 0 − sin 𝜃 . The normal vector is defined as n = ((𝜕r∕𝜕s)
[ ]T
× (𝜕r∕𝜕𝜃)) ∕ |(𝜕r∕𝜕s) × (𝜕r∕𝜕𝜃)| = 𝛽 sin (s∕a) sin 𝜃 cos (s∕a) sin (s∕a) cos 𝜃 , where 𝛽 =
[
R + a sin(s/a). The second derivatives of r are 𝜕 2 r∕𝜕s2 = − (1∕a) sin (s∕a) sin 𝜃 cos (s∕a)
]T [ ]T
sin (s∕a) cos 𝜃 , 𝜕 2 r∕𝜕s𝜕𝜃 = cos (s∕a) cos 𝜃 0 − cos (s∕a) sin 𝜃 , and 𝜕 2 r∕𝜕𝜃 2 =
[ ]T
−𝛽 sin 𝜃 0 cos 𝜃 . Therefore, the coefficients of the second fundamental form are
L = (𝜕 2 r/𝜕s2 ) ⋅ n = − 𝛽/a, M = (𝜕 2 r/𝜕s𝜕𝜃) ⋅ n = 0, and N = (𝜕 2 r/𝜕𝜃 2 ) ⋅ n = − 𝛽 sin(s/a).

Example 2.16 The parametric equation of the conical wheel surface shown in
Figure 14, which opens along the Y axis, is defined in Example 13 in terms of the
( ) [ ]T [ ]T
two parameters ys and 𝜃 as r ys , 𝜃 = x y z = ys tan 𝛾 sin 𝜃 ys ys tan 𝛾 cos 𝜃 .
[ ]T
The two tangent vectors are defined as t1 = 𝜕r∕𝜕ys = tan 𝛾 sin 𝜃 1 tan 𝛾 cos 𝜃
[ ]T
and t2 = 𝜕r∕𝜕𝜃 = ys tan 𝛾 cos 𝜃 0 − tan 𝛾 sin 𝜃 . The unit normal vector is
[ ]T
defined as n = cos 𝛾 − sin 𝜃 tan 𝛾 − cos 𝜃 . The second derivatives of vector
[ ] T [ ]T
r are 𝜕 2 r∕𝜕y2s = 0 0 0 , 𝜕 2 r∕𝜕ys 𝜕𝜃 = tan 𝛾 cos 𝜃 0 − sin 𝜃 , and 𝜕 2 r∕𝜕𝜃 2 =
[ ]T
−ys tan 𝛾 sin 𝜃 0 cos 𝜃 . The coefficients of the second fundamental form can then be
( 2 )
defined as L = 𝜕 r∕𝜕y2s ⋅ n = 0, M = (𝜕 2 r/𝜕ys 𝜕𝜃) ⋅ n = 0, and N = (𝜕 2 r/𝜕𝜃 2 ) ⋅ n = ys sin 𝛾.
Differential Geometry 69

2.6 NORMAL CURVATURE


As explained in Section 1, the normal to a curve is a unit vector along the curvature vector. In
general, the normal of a regular curve drawn on a surface is not the same as the surface nor-
mal n. This fact can be demonstrated by considering the simple example of a planar surface
parameterized by the Cartesian coordinates x and y, as shown in Figure 17. The paramet-
ric equation of the planar surface in terms of these Cartesian parameters can be written
[ ]T [ ]T
as r = x y 0 . The two tangents to the planar surface are t1 = 𝜕r∕𝜕x = 1 0 0 and
[ ]T
t2 = 𝜕r∕𝜕y = 0 1 0 . The normal to the surface is defined as n = (𝜕r∕𝜕x) × (𝜕r∕𝜕y) =
[ ]T
0 0 1 . Consider a circular curve C drawn on the planar surface. The circular curve
is assumed to have radius a, and it is parameterized by its arc length s. By tracing the
curve on the surface, the surface parameters x and y are related because x = a cos(s/a) and
y = a sin(s/a). Therefore, the parametric equation of the circular curve on the planar surface
[ ]T
can be written as C = C (s) = a cos (s∕a) a sin (s∕a) 0 . The tangent to the curve and
[ ]T
the curvature vector are defined, respectively, as 𝜕C (s) ∕𝜕s = − sin (s∕a) cos (s∕a) 0
[ ]T
and 𝜕 2 C (s) ∕𝜕s2 = − (1∕a) cos (s∕a) sin (s∕a) 0 , which shows the curve normal vector
( ) [ ]T
nc is 𝜕 2 C (s) ∕𝜕s2 ∕ ||𝜕 2 C (s) ∕𝜕s2 || = − cos (s∕a) sin (s∕a) 0 , which is a vector that lies
on the surface; therefore, in this example, the vectors normal to the curve and the surface
are orthogonal vectors.

Y
n

t2 s

nC C
s/a
O t1 X

Figure 2.17 Surface normal versus curve normal.

While for the planar surface, the vectors normal to the curves on the surface are
orthogonal to the surface normal, for other surfaces, the vectors normal to the curves
do not lie in the same plane. For example, consider the cylindrical surface previously
discussed in this chapter in Example 9 and shown in Figure 18. The parametric equation
of the cylindrical surface, which has an axis along the X axis and radius a, are defined
[ ]T [ ]T
in terms of the two parameters x and 𝜃 as r (x, 𝜃) = x y z = x a cos 𝜃 a sin 𝜃 .
[ ]T
The two tangent vectors were defined as t1 = 𝜕r∕𝜕x = 1 0 0 and t2 = 𝜕r∕𝜕𝜃 =
[ ]T
a 0 − sin 𝜃 cos 𝜃 . The unit vector normal to the surface is defined as (𝜕r∕𝜕x) ×
70 Mathematical Foundation of Railroad Vehicle Systems

[ ]T
(𝜕r∕𝜕𝜃) ∕ |(𝜕r∕𝜕x) × (𝜕r∕𝜕𝜃)| = − 0 cos 𝜃 sin 𝜃 . For a given x = xs , and keeping in
mind that 𝜃 = s/a where s is the curve arc length, one has a profile curve defined by the
[ ]T
parametric equation C (s) = xs a cos (s∕a) a sin (s∕a) . The tangent and curvature
[ ]T
vectors of this curve are, respectively, given by 𝜕C (s) ∕𝜕s = 0 − sin (s∕a) cos (s∕a)
[ ] T
and 𝜕 2 C (s) ∕𝜕s2 = − (1∕a) 0 cos (s∕a) sin (s∕a) , which defines the unit normal to
( 2 ) [ ]T
the curve as 𝜕 C (s) ∕𝜕s2 ∕ ||𝜕 2 C (s) ∕𝜕s2 || = − 0 cos (s∕a) sin (s∕a) ; this can also be
( ) [ ]T
written in terms of 𝜃 as 𝜕 2 C (s) ∕𝜕s2 ∕ ||𝜕 2 C (s) ∕𝜕s2 || = − 0 cos 𝜃 sin 𝜃 , which is the
same as the normal to the surface defined by vector n. Therefore, on a surface with general
geometry, different curves can be constructed. These curves may intersect at a given point
and can have curvature vectors that have different orientations with respect to the normal
to the surface at this point.

t2

Z t1

P
r
a
Y x
θ X

Figure 2.18 Normals of cylindrical curves.

Let C = C(s1 (t), s2 (t)) be a regular curve defined on the surface r = r(s1 , s2 ), and let K be the
curvature vector of the curve that defines the curve normal direction. The normal curvature
vector Kn at a point on the curve is defined as the projection of the curvature vector K on
the normal to the surface n at this point. That is,
Kn = (K ⋅ n) n (2.35)
The norm of this vector, called the normal curvature, is defined as
𝜅n = K ⋅ n (2.36)
The sign of 𝜅 n can be positive or negative, depending on the direction of the normal n, with
a positive sign implying that the normal vector n is directed toward the center of the radius
of curvature. Furthermore, because the surface normal n is determined using the cross
product, n = ((𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 ))/|(𝜕r/𝜕s1 ) × (𝜕r/𝜕s2 )|, the effect of the choice of the order
of multiplication must be observed.
If t is the tangent vector to curve C and s and t are two different parameters, with s rep-
resenting the curve arc length, then the curvature vector of curve C at a point on surface r
can be written as
d2 C dt (s) dt dt (dt∕dt)
K= 2 = = = (2.37)
ds ds dt ds |dr∕dt|
Differential Geometry 71

In this equation, dr is an infinitesimal line element between two points on the curve that
also correspond to two points on the surface. Furthermore, the fact that ds = |dr/dt|dt
was utilized. Because t and n are orthogonal vectors (tT n = 0), one has d(tT n)/dt = 0,
which leads to (dt/dt) ⋅ n = − t ⋅ (dn/dt). Using this equation in Eq. 37 after mul-
tiplying by n, and given the fact that t = (dr/dt)(dt/ds) = (dr/dt)/|dr/dt| because
the curve lies on the surface, one obtains the following expression for the normal
curvature 𝜅 n :
(dt∕dt) ⋅ n t ⋅ (dn∕dt) (dr∕dt) ⋅ (dn∕dt)
𝜅n = K ⋅ n = =− =− (2.38)
|dr∕dt| |dr∕dt| (|dr∕dt|)2
It is important to note that by dividing by (dt)2 , this equation can also be written as
𝜅 n = K ⋅ n = − (dr ⋅ dn)/(|dr|)2 . One can show that the normal curvature given in Eq. 38
depends on the ratio ds1 /ds2 , which defines direction. Because the tangent and normal
vectors to the surface at a given point are independent of the curve drawn on the surface,
all curves that intersect at a given point on the surface have the same normal curva-
ture 𝜅 n = K ⋅ n = − (dr ⋅ dn)/(|dr|)2 at the intersection point for the same ratio ds1 /ds2 .
Equation 38 for the normal curvature 𝜅 n can also be written in terms of the coefficients of
the first and second fundamental forms of surfaces as
( )2 ( )( ) ( )2
L ds1 ∕dt + 2M ds1 ∕dt ds2 ∕dt + N ds2 ∕dt
𝜅n = ( )2 ( )( ) ( )2 (2.39)
E ds1 ∕dt + 2F ds1 ∕dt ds2 ∕dt + G ds2 ∕dt
This equation or the equation 𝜅 n = K ⋅ n = − (dr ⋅ dn)/(|dr|)2 can also be used to write the
normal curvature as the ratio of the second and first fundamental forms as
( )2 ( )2
L ds1 + 2Mds1 ds2 + N ds2 II
𝜅n = ( )2 ( )2 = (2.40)
E ds + 2Fds ds + G ds I
1 1 2 2

In deriving Eqs. 39 and 40, it is assumed that (ds1 )2 + (ds2 )2 ≠ 0. Furthermore, because the
first fundamental form I measures the square of the length of a line element dr and con-
sequently is positive definite, the sign of the normal curvature 𝜅 n depends on the sign of
the second fundamental form II. For a planar surface, 𝜅 n is zero everywhere; for an elliptic
point on the surface, 𝜅 n ≠ 0 and assumes the sign as ds1 /ds2 ; for a hyperbolic point, 𝜅 n can
be positive, negative, or zero, depending on the value and sign of ds1 /ds2 ; and for a parabolic
point, 𝜅 n has the same sign and is zero for a zero value of II.

Example 2.17 A cylindrical surface that has an axis along the X axis and
radius a was previously discussed in Example 9 and shown in Figure 18. The
parametric equation of this surface is defined in terms of the two parameters
[ ]T [ ]T
x and 𝜃 by r (x, 𝜃) = x y z = x a cos 𝜃 a sin 𝜃 . The two tangent vectors
[ ]T [ ]T
were defined as t1 = 𝜕r∕𝜕x = 1 0 0 and t2 = 𝜕r∕𝜕𝜃 = a 0 − sin 𝜃 cos 𝜃 .
The unit vector normal to the surface is defined using the cross product n =
[ ]T
(𝜕r∕𝜕x) × (𝜕r∕𝜕𝜃) ∕ |(𝜕r∕𝜕x) × (𝜕r∕𝜕𝜃)| = − 0 cos 𝜃 sin 𝜃 . As previously dis-
cussed in this section, for a given x = xs , keeping in mind that 𝜃 = s/a where s is
the curve arc length, one has a profile curve defined by the parametric equation
[ ]T
C (s) = xs a cos (s∕a) a sin (s∕a) . The tangent and curvature vector of this curve

(Continued)
72 Mathematical Foundation of Railroad Vehicle Systems

are, respectively, given by


[ ]T
𝜕C (s) ∕𝜕s = 0 − sin (s∕a) cos (s∕a)
[ ]T
𝜕 2 C (s) ∕𝜕s2 = − (1∕a) 0 cos (s∕a) sin (s∕a)
The normal curvature is given by

]⎡
0 ⎤
( ) [
𝜅n = 𝜕 2 C (s) ∕𝜕s2 ⋅ n = − (1∕a) 0 cos (s∕a) sin (s∕a) ⎢− cos 𝜃 ⎥ = (1∕a)
⎢ ⎥
⎣ − sin 𝜃 ⎦

2.7 PRINCIPAL CURVATURES AND DIRECTIONS


The fact that the expression of the normal curvature at a point depends on a direction
defined by the ratio ds1 /ds2 suggests that there are directions in which the normal
curvature assumes maximum and minimum values. The normal curvature is called the
principal curvature if its value is this maximum or minimum value. From differential
calculus, the maximum and minimum values of the normal curvature can be obtained
by equating its derivatives with respect to s1 and s2 to zero: that is, 𝜕𝜅 n /𝜕(ds1 ) = 0 and
𝜕𝜅 n /𝜕(ds2 ) = 0. Using these two equations and the expression of the normal curvature
given by Eq. 40, one can write 𝜕𝜅 n /𝜕(dsk ) = (I(𝜕II/𝜕sk ) − II(𝜕I/𝜕sk ))/(I)2 = 0, k = 1, 2.
This equation, upon using the definition 𝜅 n = II/I, leads to (𝜕II/𝜕sk ) − 𝜅 n (𝜕I/𝜕sk ) = 0,
where 𝜕I/𝜕s1 = 2Eds1 + 2Fds2 , 𝜕I/𝜕s2 = 2Fds1 + 2Gds2 , 𝜕II/𝜕s1 = 2Lds1 + 2Mds2 , and
𝜕II/𝜕s2 = 2Mds1 + 2Nds2 . Therefore, the derivatives of 𝜅 n with respect to s1 and s2 lead
to the two equations Lds1 + Mds2 = 𝜅 n (Eds1 + Fds2 ) and Mds1 + Nds2 = 𝜅 n (Fds1 + Gds2 ),
respectively. These two scalar equations can be written in a matrix form as
[ ][ ] [ ][ ]
L M ds1 E F ds1
= 𝜅n (2.41)
M N ds2 F G ds2
This is an eigenvalue problem with symmetric coefficient matrices. In this equation, 𝜅 n
[ ]T
is the eigenvalue and y = ds1 ds2 is the eigenvector. This eigenvalue problem can be
written as
[ ][ ] [ ]
L − 𝜅n E M − 𝜅n F ds1 0
= (2.42)
M − 𝜅n F N − 𝜅n G ds2 0
This system of homogenous equations in ds1 and ds2 has a nontrivial solution if and only if
the determinant of the coefficient matrix is equal to zero. This yields the condition
| L − 𝜅 E M − 𝜅 F| ( ) ( )2
| n |
| n
| = EG − (F)2 𝜅n − (EN + GL − 2FM) 𝜅n + LN − (M)2 = 0
|M − 𝜅n F N − 𝜅n G |
| |
(2.43)
This is the characteristic equation, which is quadratic in the normal curvature 𝜅 n . This
equation has the following two roots:
√( )
2
−b ± b − 4a c
𝜅1,2 = (2.44)
2a
Differential Geometry 73

In this equation, a = EG − (F)2 , b = − (EN + GL − 2FM), and c = LN − (M)2 . The two


roots 𝜅 1 and 𝜅 2 determine the principal curvatures that define the maximum and mini-
[( ) ( ) ]T
mum curvatures. The principal directions ds1 i ds2 i , i = 1, 2 are determined using
Eq. 42 as
[ ] [( ) ] [ ]
L − 𝜅i E M − 𝜅i F ds 0
( 1 )i = , i = 1, 2 (2.45)
M − 𝜅i F N − 𝜅i G ds2 i 0
Because Eq. 45 is a homogeneous system of algebraic equations with a singular coefficient
[( ) ( ) ]T
matrix, the principal directions ds1 i ds2 i , i = 1, 2 can be determined to within an
arbitrary constant. Furthermore, because of the symmetry of the coefficient matrices of
Eq. 41, the eigenvalues that represent the principal curvatures and the eigenvectors that
represent the principal directions are guaranteed to be real non-complex numbers.
Two important definitions are often made in the differential geometry of surfaces. These
are the mean curvature 𝜅 m and the Gaussian curvature 𝜅 G . These two curvatures are defined
using the principal curvatures as
1( )
𝜅m = 𝜅 + 𝜅2 , 𝜅G = 𝜅1 𝜅2 (2.46)
2 1
It can be proven that the mean curvature is half the trace of a matrix H, while the Gaussian
curvature is the determinant of the same matrix H. Matrix H, formed from the coefficient
matrices of Eq. 41, is a function of the coefficients of the first and second fundamental forms
and can be written as follows:
[ ]−1 [ ] [ ]
E F L M 1 (GL − FM) (GM − FN)
H= = (2.47)
F G M N EG − (F)2 (EM − FL) (EN − FM)
That is, the mean curvature 𝜅 m , which represents the average of the principal curvatures,
and the Gaussian curvature 𝜅 G , which is the product of the principal curvatures, are invari-
ants of matrix H. Note that the eigenvalue problem in Eq. 42 can be simply written as
H − 𝜅 n I = 0, where I is the 2 × 2 identity matrix.

Example 2.18 Examples 11 and 14 considered the special case of a straight cylindrical
[ ]T [ ( )]T
rail that has the parametric surface equation r = x y z = x ys f r x, ys , where
( )
x and ys are, respectively, the longitudinal and lateral surface parameters, and f r ys =
√ ( )2
(a)2 − ys , where a is the cylinder radius. The coefficients of the first fundamental
form were determined in Example 11 as
( )
E = (𝜕r∕𝜕x)T (𝜕r∕𝜕x) = 1, F = (𝜕r∕𝜕x)T 𝜕r∕𝜕ys = 0,
( )T ( ) (( ) ( ( )2 ))
2
G = 𝜕r∕𝜕ys 𝜕r∕𝜕ys = 1 + ys ∕ (a)2 − ys
( √ )
( )2 [ ]T
In Example 14, the normal vector was defined as n = 1∕ 1 + fyrs 0 −fyrs 1 ,

and the coefficients of the second fundamental form were determined as


( )T ( )T
L = 𝜕 2 r∕𝜕s21 n = 0, M = 𝜕 2 r∕𝜕s1 𝜕s2 n = 0,
( √ )
( 2 ) ( )2 ( )
2 T
N = 𝜕 r∕𝜕s2 n = 1∕ 1 + fys r 𝜕fyrs ∕𝜕ys

( ( )2 )3∕2
where 𝜕fyrs ∕𝜕ys = −(a)2 ∕ (a)2 − ys .

(Continued)
74 Mathematical Foundation of Railroad Vehicle Systems

Consider first the case in which a = 0.05 m and ys = 0 m. Using these values, one has
f r = 0.05, fyrs = 0, and 𝜕fyrs ∕𝜕ys = −20; the coefficients of the first fundamental form
are E = 1, F = 0, and G = 1; and the coefficients of the second fundamental form are
L = 0, M = 0, and N = − 20. In this case, the characteristic equation (EG − (F)2 )(𝜅 n )2 −
(EN + GL − 2FM)𝜅 n + LN − (M)2 = 0 reduces to (𝜅 n )2 + 20𝜅 n = 0, which has the
roots 𝜅 1 = 0 and 𝜅 2 = − 20. The principal directions can be determined using
Eq. 45, which shows that (ds2 )i /(ds1 )i = − (L − 𝜅 i E)/(M − 𝜅 i F) or (ds2 )i /(ds1 )i =
− (M − 𝜅 i F)/(N − 𝜅 i G). For 𝜅 1 = 0, and considering s1 = x and s2 = ys , one has
(dys )1 /(dx)1 = 0, which shows that the first principal direction associated with
𝜅 1 , which can be determined to within an arbitrary constant, can be written as
[ ( ) ]T [ ]T
(dx)1 dys 1 = 1 0 . For 𝜅 2 = − 20, one has (dx)2 /(dys )2 = 0, which shows that
[ ( ) ]T [ ]T
the second principal direction can be written as (dx)1 dys 1 = 0 1 . One can
also show that matrix H is given in this case as
[ ] [ ]
1 (GL − FM) (GM − FN) 0 0
H= =
EG − (F)2 (EM − FL) (EN − FM) 0 −20
This matrix has a trace equal to −20, which is equal to 𝜅 1 + 𝜅 2 , and a zero determinant
that is equal to 𝜅 1 𝜅 2 . Recall that the mean curvature is defined as 𝜅 m = (𝜅 1 + 𝜅 2 )/2, and
the Gaussian curvature is defined as 𝜅 G = 𝜅 1 𝜅 2 . Note also that because LN − (M)2 < 0,
the surface is hyperbolic at the given point.
Consider second the case in which a = 0.05 m and ys = 0.025 m. Using these values,
one has f r = 0.0433, fyrs = −0.5774, and 𝜕fyrs ∕𝜕ys = −30.7920; the coefficients of the
first fundamental form are E = 1, F = 0, and G = 1.3333; and the coefficients of the
second fundamental form are L = 0, M = 0, and N = − 26.6667. In this case, the char-
acteristic equation (EG − (F)2 )(𝜅 n )2 − (EN + GL − 2FM)𝜅 n + LN − (M)2 = 0 reduces to
1.3333(𝜅 n )2 + 26.6667𝜅 n = 0, which has the roots 𝜅 1 = 0 and 𝜅 2 = − 20. These are the
same results obtained in the case in which ys = 0. Furthermore, because LN − (M)2 < 0,
the surface is hyperbolic at the point defined by ys = 0.025 m.

Example 2.19 A surface is defined in its parametric form by the equation


[ ]T
r = x 4xy y
Determine the type of the surface and the principal curvatures and directions at the
surface point whose coordinates are x = 1 and y = 1.
Solution Using the surface parametric equation, one can write
⎡1⎤ ⎡0⎤ ⎡ 4y ⎤
1
t1 = 𝜕r∕𝜕x = ⎢4y⎥ , t2 = 𝜕r∕𝜕y = ⎢4x⎥ , n= √ ( 2
⎢ ⎥
) −1⎥ ,
⎢ ⎥ ⎢ ⎥ 2 ⎢
⎣0⎦ ⎣1⎦ 1 + 16 (x) + (y) ⎣ 4x ⎦
⎡0⎤
𝜕 2 r∕𝜕x2 = 𝜕 2 r∕𝜕y2 = 𝟎, 𝜕 2 r∕𝜕xy = ⎢4⎥
⎢ ⎥
⎣0⎦
Differential Geometry 75

Using these definitions for the surface, the coefficients of the first fundamental form are
given by
E = (𝜕r∕𝜕x) ⋅ (𝜕r∕𝜕x) = 1 + 16(y)2
F = (𝜕r∕𝜕x) ⋅ (𝜕r∕𝜕y) = 16xy
G = (𝜕r∕𝜕y) ⋅ (𝜕r∕𝜕y) = 1 + 16(x)2
The coefficients of the second fundamental form are
( ) ( )
L = 𝜕 2 r∕𝜕x2 ⋅ n = 0, N = 𝜕 2 r∕𝜕y2 ⋅ n = 0,
( ) −4
M = 𝜕 2 r∕𝜕x𝜕y ⋅ n = √ ( 2 )
1 + 16 (x) + (y)2

At the point defined by the parameters x = 1 and y = 1, the coefficients of the first and
second fundamental forms are given by

E = 17, F = 16, G = 17, L = 0, M = −4∕ 33, N = 0
Because LN − (M)2 < 0 at the given point, the surface is hyperbolic at this point. In fact,
the surface is hyperbolic everywhere because the condition LN − (M)2 < 0 is satisfied at
every point on the surface. The principal curvatures at the given point can be defined
using Eq. 44 as
√( )
2
−b ± b − 4a c
𝜅1,2 =
2a

2
where a = √EG − (F) = 33, b = − (EN + GL − 2FM) = −128∕ 33 and c = LN −
(M)2 = −16∕ 33. Using these values, the principal curvatures can be evaluated as
𝜅1 = 0.0211, 𝜅2 = −0.6963
The principal directions can be determined using Eq. 45, which shows that
(ds2 )i /(ds1 )i = − (L − 𝜅 i E)/(M − 𝜅 i F) or (ds2 )i /(ds1 )i = − (M − 𝜅 i F)/(N − 𝜅 i G). Using
these ratios, and considering s1 = x and s2 = y, the principal directions are defined, to
within an arbitrary constant, as
[ ] [ ] [ ] [ ]
(dx)1 1 (dx)2 1
= , =
(dy)1 0.3469 (dy)2 1.1712
To check the results of the principal curvatures, one can write matrix H as
[ ] [ ]
1 (GL − FM) (GM − FN) 0.3376 −0.3587
H= =
EG − (F)2 (EM − FL) (EN − FM) −0.3587 0.3376
This equation shows that the trace of H is given by tr(H) = − 0.6752 = 𝜅 1 + 𝜅 2 = 2𝜅 m ,
and the determinant of H is |H| = − 0.01469 = 𝜅 G .
76 Mathematical Foundation of Railroad Vehicle Systems

2.8 NUMERICAL REPRESENTATION OF THE PROFILE


GEOMETRY

The wheel and rail profiles that define the geometry of the contact surfaces play a signifi-
cant role in railroad vehicle dynamics and stability. The wheel and rail profile geometries,
like the ones shown in Figure 19, are designed to ensure the stability of the railroad vehicle
as it negotiates both tangent and curved tracks. These profiles are not described using a
single analytical function. The rail profile, for example, is constructed using a series of arcs
that are combined together to form the profile geometry. For both new and worn profiles,
the profile geometry can be measured using a simple device called a MiniProf , like the one
shown in Figure 20. By sweeping the MiniProf over the wheel and rail profiles, it gener-
ates x − y tabulated data that can be used with interpolating functions to define the profile
numerically. The tabulated data can be used in numerical computer simulations to define
the wheel and rail contact surface geometries.

Wheelset

Track

Figure 2.19 Wheel and rail profiles.

Cubic Spline Interpolation Given the wheel or rail profile tabulated data in the form
yi = y(xi ), i = 0, 1, …, n, where x0 , x1 , …, xn are not necessarily equally spaced and are

(a) (b)

Figure 2.20 MiniProf profile measurements; (a) Wheel, Source: Greenwood Engineering (b) Rail.
Differential Geometry 77

not overlapping – that is, x0 ≤ x1 ≤ · · · ≤ xn – different interpolation functions can be used


to describe the profile geometry. One of the most popular methods for the interpolation
functions used to approximate the profile geometry is the cubic spline interpolation. Cubic
spline functions are considered the most popular spline functions because they are smooth
functions and do not exhibit the oscillatory behavior that characterizes higher-order inter-
polation. They are simple to develop and use and can lead to very efficient implementation.
Consider a profile segment over the interval defined by the two coordinates xi − 1 and xi .
The goal is to be able to describe the profile geometry over this interval using the cubic
interpolation (Atkinson 1978)
y (x) = ai + bi x + ci (x)2 + di (x)3 , xi−1 ≤ x ≤ xi , i = 1, … , n (2.48)
Because there are n profile segments, there are 4n unknown coefficients when the cubic
interpolation is used for each segment. Therefore, one needs 4n conditions in order to be
able to determine these unknown coefficients and uniquely define the profile geometry.
To develop the conditions that are required to determine the 4n unknown coefficients
in Eq. 48, one may require that the cubic spline interpolation must pass through the data
points x0 , x1 , · · ·, xn and must satisfy the continuity of the coordinates, first derivatives, and
second derivatives: that is,
( ) ⎫
yi = y xi , i = 0, 1, … , n ⎪
( ) ( ) ⎪
yi+1 xi = yi xi , i = 1, … , n − 1⎪
( ) ( ) ⎪
𝜕yi+1 𝜕yi
= , i = 1, … , n − 1 ⎬ (2.49)
𝜕x xi 𝜕x xi ⎪
( 2 ) ( 2 ) ⎪
𝜕 yi+1 𝜕 yi
= , i = 1, … , n − 1 ⎪ ⎪
𝜕x2 xi 𝜕x2 xi ⎭
where yi refers to the interpolating function over the segment defined by the non-
overlapping coordinates xi − 1 and xi . It is clear that Eq. 49 defines n + 1 + 3(n − 1) = 4n − 2
conditions; therefore, two coefficients can be varied arbitrarily, thereby defining two fam-
ilies of solutions. To have a unique cubic spline representation, two additional conditions
must be enforced. Typically, these two additional conditions, which are not unique, are
enforced as boundary conditions at x0 and xn . Among the most common choices for these
boundary conditions are the following:
1. Assuming that the derivatives y′0 and y′n are known at the endpoints x0 and xn , respec-
tively, one can require that y(x) satisfies the two endpoint conditions (𝜕y∕𝜕x)x0 = y′0 and
(𝜕y∕𝜕x)xn = y′n . The resulting cubic spline, in this case, is called the complete cubic spline
interpolation (Atkinson 1978).
2. If the derivatives at the two endpoints are not known, one may set the second derivatives
( ) ( )
equal to zero at the endpoints: that is, 𝜕 2 y∕𝜕x2 x = 0 and 𝜕 2 y∕𝜕x2 x = 0. The resulting
0 n
interpolation, in this case, is called the natural cubic spline.

Construction of the Cubic (Spline Interpolation


)
To explain how the cubic spline is
constructed, the notation 𝜅i = 𝜕 2 y∕𝜕x2 x , i = 0, 1, … , n is used, where the constants 𝜅 i ,
i
i = 0, 1, …, n will be determined. It is clear that in the case of the cubic interpolation of
78 Mathematical Foundation of Railroad Vehicle Systems

Eq. 48, one has the following linear interpolation for the second derivative over the interval
xi ≤ x ≤ xi + 1 :
( ) ( )
𝜕2 y xi+1 − x 𝜅i + x − xi 𝜅i+1
= , i = 0, 1, … , n − 1 (2.50)
𝜕x2 Δxi
where Δxi = xi + 1 − xi . Using Eq. 50 ensures that 𝜕 2 y/𝜕x2 is continuous over the entire
domain of the spline function. Integrating Eq. 50 twice, one can show that the cubic spline
function can be written as (Atkinson 1978)
( )3 ( )3
xi+1 − x 𝜅i + x − xi 𝜅i+1 ( ) ( )
y (x) = + Ci xi+1 − x + Di x − xi (2.51)
6Δxi
where the arbitrary constants of integration Ci and Di can be determined using the first
condition in Eq. 49, yi = y(xi ), i = 0, 1, …, n, as
( ) ( )
6yi − Δxi 𝜅i 6yi+1 − Δxi 𝜅i+1
Ci = , Di = (2.52)
6Δxi 6Δxi
Substituting these constants into Eq. 51 yields
( )3 ( )3 ( ) ( )
xi+1 − x 𝜅i + x − xi 𝜅i+1 xi+1 − x yi + x − xi yi+1
y (x) = +
6Δxi Δxi
Δxi [( ) ( ) ]
− xi+1 − x 𝜅i + x − xi 𝜅i+1 , xi ≤ x ≤ xi+1 , 0 ≤ i ≤ n − 1 (2.53)
6
This interpolation ensures the continuity of y(x) over the entire domain of the spline
function and also ensures that the spline function passes by all the data points: that is,
yi = y(xi ), i = 0, 1, …, n. To determine the constants 𝜅 0 , …, 𝜅 n , the third condition of
( ) ( )
Eq. 49, 𝜕yi+1 ∕𝜕x x = 𝜕yi ∕𝜕x x , i = 1, … , n − 1, which implies the continuity of the first
i i
derivatives at x1 , …, xn − 1 , can be applied. Enforcing this continuity condition and using
some manipulations, by equating the derivatives of the two cubic functions over the two
segments xi ≤ x ≤ xi + 1 and xi − 1 ≤ x ≤ xi , one obtains
Δxi−1 Δxi + Δxi−1 Δxi y − yi yi − yi−1
𝜅i−1 + 𝜅i + 𝜅i+1 = i+1 − , i = 1, … , n − 1
6 3 6 Δxi Δxi−1
(2.54)
which defines n − 1 equations, while there are n + 1 unknown constants 𝜅 0 , …, 𝜅 n . There-
fore, two additional end conditions, as previously discussed, need to be introduced. One can
specify the first derivatives or assume zero curvatures at the endpoints x0 and xn in order to
be able to determine all the constants 𝜅 0 , …, 𝜅 n . Using either of the two choices described
previously in this section for the end conditions at x0 and xn ensures obtaining a system of
algebraic equations with a non-singular coefficient matrix that is diagonally dominant. This
system of algebraic equations can be efficiently solved in order to determine the coefficients
𝜅 0 , …, 𝜅 n and have a unique cubic spline representation (Atkinson 1978).

2.9 NUMERICAL REPRESENTATION OF SURFACE


GEOMETRY
One effective and general method that can be used to describe surface geometry in
railroad vehicle system applications is to use the absolute nodal coordinate formulation
Differential Geometry 79

(ANCF) finite elements (Shabana 2018). ANCF finite elements have been used effectively
to integrate the geometry and analysis in railroad vehicle system applications (Berzeri
et al. 2000). These elements, which have been used to describe both rail and catenary
geometries, are well-suited for the description of arbitrary geometry, and their displace-
ment field is related to computer-aided design (CAD) computational geometry methods
by a linear mapping. Therefore, such ANCF elements allow for developing a solid model
geometry and using it in the analysis without any adjustments. For example, ANCF fully
parameterized beam elements can be used to develop the geometry of the rail space curve
and superimpose on the curve geometry the profile geometry to create the desired surface
geometry, as will be described in this section.

ANCF Element Kinematics In the case of a fully parameterized three-dimensional


ANCF element, the displacement field can be written as
r (x, y, z, t) = S (x, y, z) e (t) (2.55)
In this equation, x, y, and z are the element volume parameters, t is time, r is the position
vector of an arbitrary point on the element, S is the element shape function matrix, and e
is the element vector of the nodal coordinates. If ANCF elements are used to describe only
the geometry without considering deformations or displacements, as in the case of a fixed
rigid rail, then the preceding equation becomes independent of time, and the vector of nodal
coordinates becomes constant. In this special case, the preceding equation reduces to
r (x, y, z) = S (x, y, z) e (2.56)
Since the focus in this chapter is on the numerical description of the geometry, Eq. 56 will
be used. In the case of the fully parameterized ANCF beam element shown in Figure 21,
the shape function matrix S and vector of nodal coordinates e can be defined by starting
with the following interpolating polynomials:
2 3
⎡r1 ⎤ ⎡a0 + a1 x + a2 y + a3 z + a4 xy + a5 xz + a6 (x) + a7 (x) ⎤
r = ⎢r2 ⎥ = ⎢ b0 + b1 x + b2 y + b3 z + b4 xy + b5 xz + b6 (x) + b7 (x) ⎥
2 3
(2.57)
⎢ ⎥ ⎢ ⎥
⎣r3 ⎦ ⎣ c0 + c1 x + c2 y + c3 z + c4 xy + c5 xz + c6 (x)2 + c7 (x)3 ⎦
In this equation, ai , bi , ci , i = 0, 1, …, 7 are the coefficients of the interpolating polynomials.
The interpolation in the preceding equation is cubic in x and linear in y and z. Using
polynomials, which are cubic in x, allows for accurately describing the curvature of the
rail space curve. The 24 coefficients ai , bi , ci , i = 0, 1, …, 7 can be replaced by coordinates
that have physical meaning. For ANCF elements, these are position and position-gradient
coordinates. As shown in Figure 22, the position and gradient coordinates are introduced
at two endpoints called nodes. For a node k, k = 1, 2, the following 12 coordinates are

Rail geometry Catenary geometry

Figure 2.21 ANCF geometry.


80 Mathematical Foundation of Railroad Vehicle Systems

r2z
r2x

r2y

r2

Z
Y
O
r1 r1z
X
r1x
r1y

Figure 2.22 ANCF elements.

introduced: rk , rkx = 𝜕rk ∕𝜕x, rky = 𝜕rk ∕𝜕y, and rkz = 𝜕rk ∕𝜕z. Therefore, there are 24 new
coordinates that can be used to replace the polynomial coefficients. To this end, the
following conditions are applied at the first node, k = 1:
⎡e1 ⎤ ⎡e4 ⎤ ⎫
𝜕r1 ⎪
r1 = r (0, 0, 0) = ⎢e2 ⎥ , = rx (0, 0, 0) = ⎢e5 ⎥ ,
⎢ ⎥ 𝜕x ⎢ ⎥ ⎪
⎣e3 ⎦ ⎣e6 ⎦ ⎪
⎬ (2.58)
⎡e7 ⎤ ⎡ 10 ⎤⎪
e
𝜕r1 𝜕r1
= ry (0, 0, 0) = ⎢e8 ⎥ , = rz (0, 0, 0) = ⎢e11 ⎥⎪
𝜕y ⎢ ⎥ 𝜕z ⎢ ⎥⎪
⎣e9 ⎦ ⎣e12 ⎦⎭
and for the second node, k = 2, one has
⎡e13 ⎤ ⎡e16 ⎤ ⎫
𝜕r2 ⎪
r2 = r (l, 0, 0) = ⎢e14 ⎥ , = rx (l, 0, 0) = ⎢e17 ⎥ ,
⎢ ⎥ 𝜕x ⎢ ⎥ ⎪
⎣e15 ⎦ ⎣e18 ⎦ ⎪
⎬ (2.59)
⎡e19 ⎤ ⎡e22 ⎤⎪
𝜕r2 𝜕r2
= ry (l, 0, 0) = ⎢e20 ⎥ , = rz (l, 0, 0) = ⎢e23 ⎥⎪
𝜕y ⎢ ⎥ 𝜕z ⎢ ⎥⎪
⎣e21 ⎦ ⎣e24 ⎦⎭
where l is the length of the element. The nodal coordinates defined in the preceding two
equations can be used to define a system of algebraic equations. These algebraic equations
can be used to write the polynomial coefficients in terms of the ANCF coordinate vector
[ ]T
e = e1 e2 · · · e24 . This leads to the equation r(x, y, z) = S(x, y, z)e, where
[ ]
S = s1 I s2 I s3 I s4 I s5 I s6 I s7 I s8 I (2.60)
and the shape functions si , i = 1, 2, … , 8 are defined as (Yakoub and Shabana 2001)
( ) }
s1 = 1 − 3𝜉 2 + 2𝜉 3 , s2 = l 𝜉 − 2𝜉 2 + 𝜉 3 , s3 = l (𝜂 − 𝜉𝜂) , s4 = l (𝜍 − 𝜉𝜍) ,
( )
s5 = 3𝜉 2 − 2𝜉 3 , s6 = l −𝜉 2 + 𝜉 3 , s7 = l𝜉𝜂, s8 = l𝜉𝜍
(2.61)
Differential Geometry 81

where 𝜉 = x/l, 𝜂 = y/l and 𝜍 = z/l. Using the position vector gradients as nodal coordinates
allows for representing complex railroad geometries using a small number of ANCF
elements.

Surface Parameterization The surface of an element can assume an arbitrary shape by


writing the coordinate z as a function of the other two coordinates x and y. To this end, one
can write z = f (x, y), where f is a function that defines the shape of the element surface. In
the case of the rail surface, for example, one has z = f (y) if the rail profile does not depend on
the longitudinal parameter x. Using the equation z = f (x, y) implies that the element surface
is defined using the two parameters s1 = x and s2 = y. That is, on the surface of the element,
one has
( ) ( )
r (x, y, z) = r (x, y, f (x, y)) = r (x, y) = r s1 , s2 = S s1 , s2 e (2.62)
The function z = f (x, y) can assume any form and can also be represented numerically in a
tabulated form. In the special and important case z = f (y), the cubic spline function can be
used to describe this function based on tabulated data that can be obtained, in the case of a
rail profile, using the MiniProf device.

Tangent and Normal Vectors Using Eq. 62, one can write dr = (𝜕r/𝜕x)dx + (𝜕r/𝜕y)dy
+ (𝜕r/𝜕z)dz. Using the functional relationship z = f (x, y), one has dz = (𝜕f /𝜕x)dx + (𝜕f /𝜕y)dy.
It follows that
dr = (𝜕r∕𝜕x) dx + (𝜕r∕𝜕y) dy + (𝜕r∕𝜕z) dz
= ((𝜕r∕𝜕x) + (𝜕r∕𝜕z) (𝜕f ∕𝜕x)) dx + ((𝜕r∕𝜕y) + (𝜕r∕𝜕z) (𝜕f ∕𝜕y)) dy (2.63)
This equation defines the two tangent vectors at an arbitrary point on the surface of the
element as
( ) }
𝜕r∕𝜕s1 = (𝜕r∕𝜕x) + (𝜕r∕𝜕z) (𝜕f ∕𝜕x)
( ) (2.64)
𝜕r∕𝜕s2 = (𝜕r∕𝜕y) + (𝜕r∕𝜕z) (𝜕f ∕𝜕y)
It is important to recognize the difference between x and y when used without the relation-
ship z = f (x, y). In this case, the volume parameters x, y, and z are independent, and the
three tangent vectors at any point x, y, and z of the element (𝜕r/𝜕x), (𝜕r/𝜕y), and (𝜕r/𝜕z)
are independent vectors. However, if the surface of the element is specified by the function
z = f (x, y), one has only two independent parameters s1 = x and s2 = y, and the interpretation
of parameters x and y in this case is different from their interpretation when z is not speci-
fied on the surface. This is clear from the definition of the tangent vectors given by Eq. 64,
in which (𝜕r/𝜕x) = (𝜕r/𝜕s1 ) with x = s1 and (𝜕r/𝜕y) = (𝜕r/𝜕s2 ) with y = s2 have different
directions, magnitudes, and interpretations as compared to the case when the relationship
z = f (x, y) is not used.
Using Eq. 64, the normal vector to the surface of the element can be defined as
( ) ( )
𝜕r∕𝜕s1 × 𝜕r∕𝜕s2
n = |( ) ( )|
| 𝜕r∕𝜕s1 × 𝜕r∕𝜕s2 |
| |
((𝜕r∕𝜕x) + (𝜕r∕𝜕z) (𝜕f ∕𝜕x)) × ((𝜕r∕𝜕y) + (𝜕r∕𝜕z) (𝜕f ∕𝜕y))
= (2.65)
|((𝜕r∕𝜕x) + (𝜕r∕𝜕z) (𝜕f ∕𝜕x)) × ((𝜕r∕𝜕y) + (𝜕r∕𝜕z) (𝜕f ∕𝜕y))|
82 Mathematical Foundation of Railroad Vehicle Systems

The tangent vectors (𝜕r/𝜕x) = (𝜕r/𝜕s1 ) and (𝜕r/𝜕y) = (𝜕r/𝜕s2 ) with x = s1 and y = s2 can be
used to define the coefficients of the first fundamental form. The definition of the normal
vector given by Eq. 65 and the second derivatives of vector r with respect to x = s1 and y = s2
can be used to define the coefficients of the second fundamental form. These derivatives
can be evaluated conveniently using the element shape function matrix and vector of nodal
coordinates in Eq. 62. The coefficients of the first and second fundamental forms of surfaces
obtained using the ANCF geometry can be used to systematically define the principal cur-
vatures and principal directions of the surfaces at the wheel/rail contact points using the
procedure previously discussed in this chapter.
83

Chapter 3

MOTION AND GEOMETRY DESCRIPTIONS

In railroad vehicle dynamics, the motion of vehicle components strongly depends on the
wheel and rail geometries. The dynamics and vibration behavior, ride comfort, noise, and
forces of a passenger or freight train are determined by the track quality and geometry
and are also heavily influenced by the wheel and rail profile geometries. In general, in the
case of unconstrained motion, the displacement of a rigid body in space can be described
using six independent coordinates. Three coordinates define the global position of a point
on the body, called the body reference point; and three coordinates define the orientation of
the body with respect to the global coordinate system. The global position of the body ref-
erence point can be defined using three Cartesian coordinates. The orientation coordinates
can be introduced using three independent parameters that can represent angles or can be
parameters that do not have an obvious physical meaning. Therefore, in spatial analysis, ori-
entation parameters are not unique, and different sets of parameters have been used in the
literature and in developing computational multibody system (MBS) algorithms. Further-
more, in spatial analysis, angular velocities are not exact differentials; and therefore, they
are not the time derivatives of orientation parameters. That is, angular velocities cannot
be directly integrated to determine orientation parameters. Nonetheless, angular velocities
can always be written as linear functions of the derivatives of the orientation parameters
using a velocity transformation matrix. This velocity transformation plays a fundamental
role in determining the generalized forces associated with the orientation parameters since
these orientation parameters serve as generalized coordinates and are not directly associated
with the Cartesian moments applied to the bodies, as discussed in Chapter 6.
The kinematic description that will be used in this book to develop the equations of
motion of the components of railroad vehicles are introduced in this chapter. As discussed
in this chapter, using three parameters, such as Euler angles, to define a body orienta-
tion in space leads to kinematic singularities. This singularity can be avoided by using
the four Euler parameters at the expense of adding an algebraic constraint equation that
relates them. Euler parameters, which are becoming more popular in developing general
MBS algorithms, have many identities that can be used to simplify kinematic and dynamic
equations.

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
84 Mathematical Foundation of Railroad Vehicle Systems

In addition to defining the orientation of bodies in space, Euler angles have been also
used in railroad vehicle dynamics to define the geometry of the track based on given simple
industry inputs. For the most part, track is constructed using three main segments: tangent
(straight), curve, and spiral. The tangent segment has zero curvature; the curve segment has
constant curvature; and the spiral segment, used to connect two segments with different
curvature values, has a curvature that varies linearly in order to ensure a smooth transition
between the two segments. Track geometry is often described using three inputs at points
along the track at which the geometry changes. These three inputs – horizontal curvature,
superelevation, and grade – can be used to define uniquely three Euler angles that are used
to construct the track and rail space curves. To this end, Euler angles are converted to field
variables and used systematically to construct a curve with well-defined geometry based on
given simple track inputs. Because the three Euler angles are in general independent when
used as time-dependent motion coordinates for unconstrained bodies, while the geometry
of a curve is uniquely defined using one parameter that can be the arc length, converting
Euler angles to field variables expressed in terms of the curve arc length ensures unique
definition of the curve geometry.
Therefore, it is important to recognize that Euler angles are used in this chapter for two
fundamentally different purposes: (i) as motion-generalized coordinates to describe rigid
body kinematics in space; and (ii) as geometry field variables to uniquely define the geom-
etry of track and rail space curves. The analysis presented in this chapter is used as the
foundation for a computer procedure designed to develop the track geometry data required
for nonlinear dynamic simulations of railroad vehicle systems. The data can be generated
once before the dynamic simulation at a preprocessing stage in a track preprocessor com-
puter program, as discussed in Chapter 4. The track preprocessor output file normally has
data for three different space curves: the track centerline space curve, right rail space curve,
and left rail space curve. These three curves can have different geometries. The right and
left rail space curves are used in the formulation of wheel/rail contact conditions, while
the track space curve is used in the motion description of the coordinate systems of vehicle
components.

3.1 RIGID-BODY KINEMATICS


An arbitrary rigid body in a railroad vehicle/track system will be referred to as body i.
The unconstrained motion of this body in spatial analysis is described using six indepen-
dent coordinates, called body generalized coordinates. These coordinates are three indepen-
dent translation coordinates of a selected reference point on the body and three rotation
coordinates that define the body orientation in space. As shown in Figure 1, translational
coordinates can be defined in the global coordinate system XYZ using the position vector
[ ]T
Ri = Rix Riy Riz of the body reference point Oi , which is assumed to be rigidly attached
to the body in rigid body dynamics. The body orientation can be defined using orienta-
tion parameters that define the direction cosines of the axes of the body coordinate system
X i Y i Z i in the global XYZ coordinate system. It is clear from Figure 1 that one can write the
global position vector ri of an arbitrary point on the body as
ri = Ri + ui (3.1)
Motion and Geometry Descriptions 85

Zi
Yi

Oi
ui

Xi Pi
Z
Y Ri

ri

Figure 3.1 Rigid body coordinates.


[ ]T
where the vector ui = uix uiy uiz defines the location of the arbitrary point with respect
to the origin of the body coordinate system X i Y i Z i in the global system: that is,
[ ]T
ui = uix uiy uiz = uix i + uiy j + uiz k (3.2)
In this equation, i, j, and k are, respectively, unit vectors along the global axes X, Y , and Z.
Alternatively, vector ui can be written in terms of components defined in the body coordi-
nate system X i Y i Z i as
ui = uix ii + uiy ji + uiz ki

⎡uix ⎤
[ ] ⎢ ⎥
= ii ji ki ⎢uiy ⎥ = Ai ui (3.3)
⎢ i⎥
⎣ uz ⎦
where ii , ji , and ki are, respectively, unit vectors along the axes of the body coordinate system
X i , Y i , and Z i defined in the global system; uix , uiy , and uiz are the component of vector ui
[ ]T [ ]T
defined in the body coordinate system X i Y i Z i ; ui = uix uiy uiz = xi yi zi ; and Ai =
[i i ]
i j ki is the 3 × 3 transformation matrix whose columns define the axes (orientation)
of the body coordinate system X i Y i Z i . The components of vector ui are constant for rigid
body dynamics. Substituting Eq. 3 into Eq. 1, the global position vector of an arbitrary point
on the rigid body can be written as
ri = Ri + Ai ui (3.4)
While there is no restriction on the choice of body reference point Oi , selecting the center
of mass of the body as the reference point leads to significant simplifications in the form of
the equations of motion. Using the body center of mass as the reference point eliminates
inertia coupling between the body translation and rotation coordinates and leads to the
definition of Newton–Euler equations widely used in rigid body dynamics.
86 Mathematical Foundation of Railroad Vehicle Systems

3.2 DIRECTION COSINES AND SIMPLE ROTATIONS


This section introduces a general form of transformation matrix Ai that defines the
orientation of the body in space and is used to define transformation matrices as a result
of performing simple rotations about the axes of the body coordinate system X i Y i Z i .
These simple-rotation transformation matrices are used in a later section of this chapter
to introduce the three independent Euler angles widely used in the motion description of
three-dimensional bodies.

Direction
[
Cosines
]
It is clear that the columns of the transformation matrix
Ai = ii ji ki are unit vectors along the axes of the body coordinate system X i Y i Z i . The
elements of vectors ii , ji , and ki represent the projection of these unit vectors along the axes
of the global coordinate system defined by three unit vectors i, j, and k. For example, as
shown in Figure 2, the components of unit vector ii along the X, Y , and Z axes are defined,
respectively, by 𝛼11 i
= ii ⋅ i = cos 𝛽1 , 𝛼12
i
= ii ⋅ j = cos 𝛽2 , and 𝛼13
i
= ii ⋅ k = cos 𝛽3 , where
𝛽 1 , 𝛽 2 , and 𝛽 3 are, respectively, the angles between the body axis X i and the global axes
X, Y , and Z. Similar definitions for the components of unit vectors ji and ki can be made.
Therefore, one can write
[ ]T [ i i T
] ⎫
ii = ii ⋅ i ii ⋅ j ii ⋅ k = 𝛼11 𝛼12
i
𝛼13 , ⎪
[ ]T [ i i T
] ⎪
ji = ji ⋅ i ji ⋅ j ji ⋅ k = 𝛼21 𝛼22
i
𝛼23 , ⎬ (3.5)
[ ]T [ i ] ⎪
i T
ki = ki ⋅ i ki ⋅ j ki ⋅ k = 𝛼31 𝛼32
i
𝛼33 ⎪

Elements 𝛼jli , j, l = 1, 2, 3, which represent the components of orthogonal unit vectors ii ,
ji ,
and ki along the global X, Y , and Z axes are called direction cosines. Therefore, the trans-
formation matrix that defines the orientation of the body coordinate system with respect to
the global system can be written in terms of direction cosines as
⎡𝛼11
i
𝛼21
i
𝛼31
i

[ ]⎢ i ⎥
Ai = ii ji k = ⎢𝛼12 𝛼22 𝛼32 ⎥
i i i
(3.6)
⎢ i i ⎥
⎣𝛼13 𝛼23
i
𝛼33 ⎦

β3
Xi
ii
O β2

β1
i j
Y
X

Figure 3.2 Direction cosines.


Motion and Geometry Descriptions 87

Direction cosines are not independent because the columns of the transformation matrix
Ai are orthogonal unit vectors. Therefore, direction cosines are related by the following six
constraint equations:

𝛼k1
i
𝛼l1i + 𝛼k2
i
𝛼l2i + 𝛼k3
i
𝛼l3i = 𝛿kl , k, l = 1, 2, 3 (3.7)

where 𝛿 kl is the Kronecker delta, which is equal to one if k = l and equal to zero if k ≠ l. Using
the transformation matrix of Eq. 6 and the six constraint equations of Eq. 7, one can show
T T
that transformation matrix Ai is an orthogonal matrix: that is, Ai Ai = Ai Ai = I, where
I is the 3 × 3 identity matrix. It is also important to note that since there are nine direc-
tion cosines related by six algebraic constraint equations (Eq. 7), only three independent
parameters are required in order to describe the orientation of the body in space. These
three independent parameters will be introduced in a later section of this chapter.

Example 3.1
Axes X i and Z i of the coordinate system of a rigid body i are defined, respectively, in the
[ ]T [ ]T
global coordinate system by the vectors 0.0 2.0 2.0 and −2.0 −1.0 1.0 . Obtain
the transformation matrix that defines the orientation of body i with respect to the global
system. √ √
[ ]T
Solution The norm of the vector 0.0 2.0 2.0 is (0)2 + (2.0)2 + (2.0)2 = 2 2,
[ ]T √ √
and the norm of the vector −2.0 −1.0 1.0 is (−2.0)2 + (−1.0)2 + (1.0)2 = 6.
Therefore, unit vectors ii and ki along the body axes X i and Z i can be determined,
respectively, as
⎡0.0⎤ ⎡ 0.0 ⎤ ⎡−2.0⎤ ⎡−0.8165⎤
1 ⎢ ⎥ ⎢ ⎥ 1 ⎢ ⎥ ⎢ ⎥
i = √ ⎢2.0⎥ = ⎢0.7071⎥ ,
i
k = √ ⎢−1.0⎥ = ⎢−0.4082⎥
i
2 2⎢ ⎥ ⎢ ⎥ 6⎢ ⎥ ⎢ ⎥
⎣2.0⎦ ⎣0.7071⎦ ⎣ 1.0 ⎦ ⎣ 0.4082 ⎦
A unit vector ji along axis Y i can be determined using the cross product ji = ki × ii , which
leads to
⎡−0.5774⎤
i i ⎢ i ⎥
j = k × i = ⎢ 0.5774 ⎥
⎢ ⎥
⎣−0.5774⎦
The three vectors ii , ji , and ki represent the columns of the transformation matrix that
defines the orientation of the coordinate system X i Y i Z i with respect to the global coor-
dinate system XYZ as
⎡ 0.0 −0.5774 −0.8165⎤
i
[ i i i
⎢] ⎥
A = i j k = ⎢0.7071 0.5774 −0.4082⎥
⎢ ⎥
⎣0.7071 −0.5774 0.4082 ⎦
88 Mathematical Foundation of Railroad Vehicle Systems

Zi
Yi

ϕxi
Xi

Figure 3.3 Simple rotations.

Simple Rotations Using direction cosines, one can develop the form of the transforma-
tion matrices that result from simple rotations about the axes of the body coordinate system.
In the development presented in this section, it is assumed that the global XYZ and body
X i Y i Z i coordinate systems initially coincide before performing a simple rotation.
For a simple rotation 𝜙x about the X i axis, shown in Figure 3, one can show that
[ i i ]
i T
[ ]T [ ]T [ i i T
] [ ]T
𝛼11 𝛼12 𝛼13 = ii ⋅ i ii ⋅ j ii ⋅ k = 1 0 0 , 𝛼21 𝛼22
i
𝛼23 = ji ⋅ i ji ⋅ j ji ⋅ k =
[ ]T [ i i T
] [ ]T [ ]T
0 cos 𝜙x sin 𝜙x , and 𝛼31 𝛼32
i
𝛼33 = ki ⋅ i ki ⋅ j ki ⋅ k = 0 − sin 𝜙x cos 𝜙x .
It follows that the transformation matrix in Eq. 6 can be written as a result of this simple
rotation as
⎡1 0 0 ⎤
⎢ ⎥
A = ⎢0 cos 𝜙x − sin 𝜙x ⎥
i
(3.8)
⎢ ⎥
⎣ 0 sin 𝜙 x cos 𝜙x ⎦

Similarly, a simple rotation 𝜙y about the Y i axis is defined by the following rotation
matrix:
⎡ cos 𝜙y 0 sin 𝜙y ⎤
i ⎢ ⎥
A =⎢ 0 1 0 ⎥ (3.9)
⎢ ⎥
⎣− sin 𝜙y 0 cos 𝜙y ⎦
And a simple rotation 𝜙z about the Z i axis defines the following transformation matrix:
⎡cos 𝜙z − sin 𝜙z 0⎤
⎢ ⎥
A = ⎢ sin 𝜙z cos 𝜙z 0⎥
i
(3.10)
⎢ ⎥
⎣ 0 0 1⎦
The forms of the simple-rotation matrices in Eqs. 8–10 can be used to develop a more gen-
eral transformation matrix expressed in terms of three independent rotations. This general
transformation matrix can describe any orientation of the body in space. The three angles
used to form this general transformation are called Euler angles.

3.3 EULER ANGLES


In railroad vehicle dynamics, Euler angles are used for two fundamentally different pur-
poses. First, they are used as rotation parameters for the definition of a body orientation in
Motion and Geometry Descriptions 89

space. In this case, transformation matrix Ai in Eq. 6 is expressed in terms of three indepen-
dent angles that represent motion-generalized coordinates that vary with time as the body
moves and changes its orientation. Second, Euler angles are used as field variables to define
the geometry of rail and track space curves. In this case, Euler angles are no longer inde-
pendent, and they represent geometric variables that vary with the rail arc length. Using
Euler angles to define the curve geometry allows for developing a simple procedure based
on simple inputs used by the rail industry to create a data file in a track preprocessor com-
puter program that defines space curves in terms of nodes. This discretized form of the space
curve is used during dynamic simulations to define the tangent, normal, and curvature vec-
tors required to solve the wheel/rail contact problem, as discussed in later chapters.

Euler Angle Transformation Matrix As discussed in the preceding section, the trans-
formation matrix that defines the body orientation can be written in terms of three inde-
pendent parameters. These three independent parameters can be three angles, called Euler
angles, performed about three independent axes. When Euler angles are used, three simple
successive rotations are performed about three axes of the body coordinate system X i Y i Z i .
Different rotation sequences can be used to define Euler angles. For example, Euler used
the sequence Z i , X i , and Z i to study gyroscopic motion. In this sequence used by Euler, the
first axis Z i is different from the third axis Z i as a result of rotation about the second axis X i .
In railroad vehicle dynamics, a different sequence of Euler angles is used. First a rotation 𝜓 i
(yaw) about the Z i axis is performed, followed by a rotation 𝜙i (roll) about the new X i axis,
followed by a rotation 𝜃 i (pitch) about the new Y i body axis, as shown in Figure 4. These
three simple successive rotations define the following simple rotation matrices:

⎡cos 𝜓 i − sin 𝜓 i 0⎤ ⎡1 0 0 ⎤ ⎡ cos 𝜃 i 0 sin 𝜃 i ⎤


⎢ ⎥ ⎢ ⎥ ⎢ ⎥
Ai𝜓 = ⎢ sin 𝜓 i cos 𝜓 i 0⎥ , A𝜙 = ⎢0 cos 𝜙 − sin 𝜙i ⎥ ,
i i Ai𝜃 = ⎢ 0 1 0 ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 0 0 1⎦ ⎣0 sin 𝜙i cos 𝜙i ⎦ ⎣− sin 𝜃 i 0 cos 𝜃 i ⎦
(3.11)

ψi

Zi Zi
Zi Yi θi
Y i Zi
Yi
Yi
ϕ i

Xi Xi
i
X Xi
A
i AA
i i AiψAiϕAiθ
Z ψ ψ ϕ

O
X

Figure 3.4 Euler angles.


90 Mathematical Foundation of Railroad Vehicle Systems

This sequence, as explained later, is used to avoid kinematic singularities in most railroad
vehicle simulations in which the yaw and roll angles are always small and the pitch angle
can be very large due to wheel rotation. It is clear from Figure 4 that the orthogonal trans-
formation matrix Ai can be defined as the product of the three simple rotation matrices in
Eq. 11 as Ai = Ai𝜓 Ai𝜙 Ai𝜃 , that is,
Ai = Ai𝜓 Ai𝜙 Ai𝜃

⎡cos 𝜓 i cos 𝜃 i − sin 𝜓 i sin 𝜙i sin 𝜃 i − sin 𝜓 i cos 𝜙i cos 𝜓 i sin 𝜃 i + sin 𝜓 i sin 𝜙i cos 𝜃 i ⎤
⎢ ⎥
= ⎢sin 𝜓 i cos 𝜃 i + cos 𝜓 i sin 𝜙i sin 𝜃 i cos 𝜓 i cos 𝜙i sin 𝜓 i sin 𝜃 i − cos 𝜓 i sin 𝜙i cos 𝜃 i ⎥
⎢ ⎥
⎣ − cos 𝜙i sin 𝜃 i sin 𝜙i cos 𝜙i cos 𝜃 i ⎦
(3.12)
The columns of this transformation matrix are orthogonal unit vectors along the axes of
the body coordinate system; therefore, the magnitude of any of the elements in the matrix
of Eq. 12 should not exceed one. These elements are simply the direction cosines of the
axes of the body coordinate system X i , Y i , and Z i . It is also important to note that because
of the non-commutativity of finite rotations, Aiz Aix Aiy ≠ Aiy Aix Aiz , and the order in which
Euler angles are performed must be observed.

Kinematic Singularity All three-parameter representations of finite rotations suffer


from singularities. This is a major drawback that can lead to serious numerical problems
in computer simulations of railroad vehicle systems (Roberson and Schwertassek 1988;
Shabana 2010). When Euler angles are used, the singularity arises when the three axes
about which Euler angles are performed are not independent. The Euler-angle singular
configuration depends on the sequence of rotation used. For the sequence Z i , X i , and Y i
used in this section, the singularity occurs when the roll angle 𝜙i is equal to ±𝜋/2 because
the axes of rotations for angles 𝜓 i and 𝜃 i become parallel, and the yaw angle 𝜓 i and pitch
angle 𝜃 i are not independent. In this case, as demonstrated in a later section of this chapter,
one cannot write the derivatives of the angles in terms of the angular velocity vector when
the singular configuration is encountered. As previously mentioned, this type of kinematic
singularity is encountered when any known three-parameter method is used to define the
body orientation in space. This kinematic singularity can be avoided by using four Euler
parameters at the expense of adding an algebraic constraint equation.

Example 3.2 As previously mentioned, in his study of gyroscopic motion, Euler used
the sequence Z i − X i − Z i . This sequence of Euler angles can be defined by a rotation 𝜙i
about the body Z i axis, a rotation 𝜃 i about the body X i axis, and a rotation 𝜓 i about the
body Z i axis. These three simple rotations lead to the three transformation matrices
⎡cos 𝜙i − sin 𝜙i 0⎤ ⎡1 0 0 ⎤ ⎡cos 𝜓 i − sin 𝜓 i 0⎤
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
Ai𝜙 = ⎢ sin 𝜙i cos 𝜙i 0⎥ , A𝜃 = ⎢0 cos 𝜃 − sin 𝜃 i ⎥ ,
i i Ai𝜓 = ⎢ sin 𝜓 i cos 𝜓 i 0⎥
⎢ ⎥ ⎢ i ⎥ ⎢ ⎥
⎣0 sin 𝜃 cos 𝜃 ⎦
i
⎣ 0 0 1⎦ ⎣ 0 0 1⎦
Motion and Geometry Descriptions 91

Using these matrices, the orthogonal transformation matrix that defines the orientation
of the body X i Y i Z i coordinate system with respect to the XYZ global coordinate system
is
Ai = Ai𝜙 Ai𝜃 Ai𝜓

⎡cos 𝜓 cos 𝜙 − cos 𝜃 sin 𝜙 sin 𝜓 − sin 𝜓 cos 𝜙 − cos 𝜃 sin 𝜙 cos 𝜓 sin 𝜃 sin 𝜙 ⎤
i i i i i i i i i i i i

⎢ ⎥
= ⎢cos 𝜓 i sin 𝜙i + cos 𝜃 i cos 𝜙i sin 𝜓 i − sin 𝜓 i sin 𝜙i + cos 𝜃 i cos 𝜙i cos 𝜓 i − sin 𝜃 i cos 𝜙i ⎥
⎢ ⎥
⎣ sin 𝜃 i sin 𝜓 i sin 𝜃 i cos 𝜓 i cos 𝜃 i ⎦

3.4 EULER PARAMETERS


The four Euler parameters can be used to avoid singular configurations at the expense of
using a nonlinear algebraic constraint equation. Euler parameters, however, are becoming
more popular in MBS algorithms because of the singularity problem associated with
using three rotation parameters. This singularity problem can be the source of numerical
problems in the dynamic simulation of spinning bodies or bodies rotating at high speeds,
as is the case with rail wheelsets. Euler parameters can be systematically defined using
Rodrigues’ formula, which defines the transformation matrix in terms of the angle of rota-
tion and the components of a unit vector along the instantaneous axis of rotation. A general
three-dimensional rotation is equivalent to a single rotation about this instantaneous axis
of rotation.

Rodrigues’ Formula In order to derive Rodrigues’ formula, vector ri on body i is


assumed to rotate with an angle 𝛾 i about the axis of rotation defined by unit vector vi =
[ i i i ]T
v1 v2 v3 , as shown in Figure 5. Vector ri is assumed to make an angle 𝛼 with the axis of
rotation. It follows that radius a of the circle shown in the figure is defined as a = ||ri || sin 𝛼 =
|vi × ri |. As a result of this rotation, vector ri occupies a new position defined by vector
| |
ri , as shown in Figure 5. It is clear from the figure that ri = ri + Δri . Vector Δri has two
components (Δri )1 and (Δri )2 that are perpendicular to the axis of rotation defined by unit
vector vi . It is clear from Figure 5 that (Δri )1 has a magnitude a sin 𝛾 and a direction defined
( )
by unit vector vi × ri ∕ ||vi × ri ||, while (Δri )2 has a magnitude a(1 − cos 𝛾 i ) = 2a sin2 (𝛾 i /2)
( ) | ( )| ( )
and a direction vi × vi × ri ∕ |vi × vi × ri | = vi × vi × ri ∕a. Therefore, one has
| |
( i ) ⎫
( i) v × ri ( i )
Δr 1 = a sin 𝛾 i | i = v × r i
sin 𝛾 i ⎪
i| ⎪
|v × r | ( )
( i) i ( i )⎬ (3.13)
( i) 2 𝛾 v × v i × ri ( i
( i i
)) 2 𝛾 ⎪
Δr 2 = 2a sin = 2 v × v × r sin
2 a 2 ⎪ ⎭
92 Mathematical Foundation of Railroad Vehicle Systems

γi
(Δr i)
1
(Δri)2 Q
Q Δr
i

α ri
ri

vi

Oi

Figure 3.5 Rodrigues’ formula.

Using this equation, one can write ri = ri + Δri in a more explicit form as
( )
( i ) ( i ( i )) 2 𝛾 i
r = r + v × r sin 𝛾 + 2 v × v × r sin
i i i i i
(3.14)
2
The cross products in Eq. 14 can be written using matrix notation as vi × ri = ṽ i ri and vi ×
( i ) ( )2
v × ri = ṽ i ri , where ṽ i is the skew-symmetric matrix associated with vector vi and is
defined as
⎡ 0 −vi3 vi2 ⎤
⎢ ⎥
ṽ i = ⎢ vi3 0 −vi1 ⎥ (3.15)
⎢ i ⎥
⎣−v2 vi1 0 ⎦
Using skew-symmetric matrix notation, Eq. 14 can be written as
[ ( i )]
( i )2 𝛾
r = A r = I + ṽ sin 𝛾 + 2 ṽ sin
i i i i i 2
ri (3.16)
2
where I is the 3 × 3 identity matrix, and transformation matrix Ai is defined as
( i)
( )2 𝛾
Ai = I + ṽ i sin 𝛾 i + 2 ṽ i sin2 (3.17)
2
This equation, called Rodrigues’ formula, defines the transformation matrix in terms of
four parameters vi1 , vi2 , vi3 , and 𝛾 i . These parameters are not totally independent because
Motion and Geometry Descriptions 93

∑3 ( i )2
k=1 vk = 1, demonstrating again that general three-dimensional rotation can be
described using only three independent parameters. Rodrigues’ formula, in which matrix
( i )2
ṽ is a symmetric matrix, can be used to determine the transformation matrix as a
result of rotation about a vector that has arbitrary orientation. It can also be used to obtain
the simple rotation matrices previously presented in this chapter, as demonstrated by the
following example.

Example 3.3 Consider the simple rotation 𝜙i about the Z i axis, which is defined by
[ ]T
unit vector vi = 0 0 1 . In this case,

⎡0 −1 0⎤ ⎡−1 0 0⎤
⎢ ⎥ ( i )2 ⎢ ⎥
ṽ = ⎢1 0 0⎥ ,
i
ṽ = ⎢ 0 −1 0⎥
⎢ ⎥ ⎢ ⎥
⎣0 0 0⎦ ⎣0 0 0⎦
The transformation matrix can be determined using Rodrigues’ formula as
( )
( i )2 2 𝜙i
A = A = I + ṽ sin 𝜙 + 2 ṽ sin
i i i i
2

⎡1 0 0⎤ ⎡0 −1 0⎤ ⎡−1 0 0⎤
⎢ ⎥ ⎢ ⎥ ⎢ ⎥( )
= ⎢0 1 0⎥ + ⎢1 0 0⎥ sin 𝜙 + ⎢ 0 −1 0⎥ 1 − cos 𝜙i
i

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣0 0 1⎦ ⎣0 0 0⎦ ⎣0 0 0⎦

⎡cos 𝜙i − sin 𝜙i 0⎤
⎢ ⎥
= ⎢ sin 𝜙i cos 𝜙i 0⎥
⎢ ⎥
⎣ 0 0 1⎦
which is the same matrix previously obtained as a result of a simple rotation about the
Z i axis.

Euler Parameters Euler parameters are defined in terms of the angle of rotation 𝛾 i and
the three components vi1 , vi2 , and vi3 of unit vector vi along the axis of rotation as

𝛾i 𝛾i 𝛾i
𝛾i
𝜃0i = cos , 𝜃1i = vi1 sin , 𝜃2i = vi2 sin , 𝜃3i = vi3 sin
(3.18)
2 2 2
2
[ ]T
The four Euler parameters can be written in a vector form as 𝛉i = 𝜃0i 𝜃1i 𝜃2i 𝜃3i . These
parameters are not totally independent because they are related by the algebraic constraint
∑3 ( )2
equation k=0 𝜃ki = 1. Using Rodrigues’ formula of Eq. 17 and the definitions of Euler
parameters of Eq. 18, one can show that matrix Ai can be written in terms of Euler param-
eters as
( )
Ai = I + 2𝛉̃ is 𝜃0i I + 𝛉̃ is = Ei Ei
T
(3.19)
94 Mathematical Foundation of Railroad Vehicle Systems

[ ]T
where 𝛉is = 𝜃1i 𝜃2i 𝜃3i , 𝛉̃ is is the skew-symmetric matrix associated with vector 𝛉is , and the
two matrices Ei and Ei are the 3 × 4 matrices defined as (Nikravesh 1988; Shabana 2005)
⎡−𝜃1i 𝜃0i −𝜃3i 𝜃2i ⎤ ⎡−𝜃1i 𝜃0i 𝜃3i −𝜃2i ⎤
⎢ ⎥ ⎢ ⎥
Ei = ⎢−𝜃2i 𝜃3i 𝜃0i −𝜃1i ⎥ , Ei = ⎢−𝜃2i −𝜃3i 𝜃0i 𝜃1i ⎥ (3.20)
⎢ i ⎥ ⎢ i i ⎥
⎣−𝜃3 −𝜃2i 𝜃1i 𝜃0i ⎦ ⎣−𝜃3 𝜃2 −𝜃1 𝜃0 ⎦
i i

∑3 ( )2
Because of the Euler parameter constraint k=0 𝜃ki = 1, which allows writing one param-
eter in terms of the other three, the transformation matrix written in terms of Euler param-
eters can take different forms. One of these forms is given as
⎡2(𝜃 i )2 + 2(𝜃 i )2 − 1 (
2 𝜃1i 𝜃2i − 𝜃0i 𝜃3i
) (
2 𝜃1i 𝜃3i + 𝜃0i 𝜃2i
) ⎤
⎢ 0( 1
) ( )2 ( )2 ( ) ⎥⎥
Ai = ⎢ 2 𝜃1i 𝜃2i + 𝜃0i 𝜃3i 2 𝜃0i + 2 𝜃2i − 1 2 𝜃2i 𝜃3i − 𝜃0i 𝜃1i (3.21)
⎢ ⎥
(
⎢ 2 𝜃i 𝜃i − 𝜃i 𝜃i ) ( ) ( ) ( )
2 𝜃0i + 2 𝜃3i − 1⎥⎦
2 2
⎣ 1 3 0 2 2 𝜃2i 𝜃3i + 𝜃0i 𝜃1i
This transformation matrix is quadratic in Euler parameters, while the transformation
matrix expressed in terms of Euler angles has trigonometric functions that have infinite
order. Furthermore, using Euler parameters allows eliminating the singularity associated
with Euler angles.
Rodrigues’ formula and Euler parameters can be used to define the transformation matrix
resulting from body rotation about an axis that has any orientation in space. Therefore,
they are not restricted to simple rotations about the axes of coordinate systems. Addition-
ally, Euler parameters have many useful identities that can be used to simplify the for-
mulation of the equations of motion. In Table 1, which shows some of the identities, 𝛉i =
[ i i i i ]T
𝜃0 𝜃1 𝜃2 𝜃3 , I is the 3 × 3 identity matrix, and I4 is the 4 × 4 identity matrix.

Table 3.1 Euler parameter identities.

Euler parameters EET = EET = I, ET E = ET E = I4 − 𝛉𝛉T , E𝛉 = E𝛉 = 𝟎, 𝛉T 𝛉 = 1


Euler parameter derivatives Ė 𝛉̇ = Ė 𝛉̇ = 𝟎, EĖ T = EE
̇ T , 𝛉̇ T 𝛉 = 0, E𝛉̇ = −E𝛉, ̇
̇ E𝛉̇ = −E𝛉

Example 3.4
A body rotates an angle 𝛾 i = 𝜋/2 about an axis of rotation defined by vector
[ ]T
1.0 0 −1.0 . Use Rodrigues’ formula to determine the transformation matrix as
a result of this rotation, and determine the values of the four Euler parameters.
( √ )[
Solution A unit vector along the axis of rotation is defined as vi = 1∕ 2 1.0 0
]T
−1.0 . Using this unit vector, one has

⎡0 1 0⎤ ⎡1 0 1⎤
1 ⎢ ⎥ ( i )2 1⎢ ⎥
ṽ = √ ⎢−1 0 −1⎥ ,
i
ṽ = − ⎢0 2 0⎥
2⎢ ⎥ 2⎢ ⎥
⎣0 1 0⎦ ⎣1 0 1⎦
Motion and Geometry Descriptions 95

For the given angle 𝛾 i = 𝜋/2, one has sin𝛾 i = 1 and 2sin2 (𝛾 i /2) = 1 − cos 𝛾 i = 1. Therefore,
Rodrigues’ formula can be written as
( )2 𝛾i
Ai = I + ṽ i sin 𝛾 i + 2 ṽ i sin2
2
⎡1 0 0⎤ ⎡0 1 0⎤ ⎡1 0 1⎤
⎢ ⎥ 1 ⎢ ⎥ 1⎢ ⎥
= ⎢0 1 0⎥ + √ ⎢−1 0 −1⎥ (1) − ⎢0 2 0⎥ (1)
⎢ ⎥ 2⎢ ⎥ 2 ⎢ ⎥
⎣0 0 1⎦ ⎣0 1 0⎦ ⎣1 0 1⎦

⎡ 0.5 0.7072 −0.5 ⎤


⎢ ⎥
= ⎢−0.7072 0 −0.7072⎥
⎢ ⎥
⎣ −0.5 0.7072 0.5 ⎦
( √ )[ ]T
Using the angle 𝛾 i = 𝜋/2 and the unit vector vi = 1∕ 2 1.0 0 −1.0 along the axis
of rotation, the four Euler parameters can be defined as
( ) ( √ )
𝜃0i = cos 𝛾 i ∕2 = 1∕ 2
[ i ]T ( ) [ ]T
𝜃1 𝜃2i 𝜃3i = vi sin 𝛾 i ∕2 = 0.5 0 −0.5

3.5 VELOCITY AND ACCELERATION EQUATIONS

Equation 4 shows that the global position of an arbitrary point on a body i is defined as
ri = Ri + Ai ui . When using Euler angles, transformation matrix Ai is a function of the Euler
[ ]T
angles defined by vector 𝛉i = 𝜓 i 𝜙i 𝜃 i . Therefore, the time-dependent coordinates of the
body can be written as
[ T T ]T [ ]T
qi = R i 𝛉i = Rix Riy Riz 𝜓 i 𝜙i 𝜃 i (3.22)
[ T T ]T
For Euler parameters, this is the seven-dimensional vector defined as qi = Ri 𝛉i =
[ i ] T
Rx Riy Riz 𝜃0i 𝜃1i 𝜃2i 𝜃3i . In this case, the constraint on the Euler parameters
∑3 ( i )2
k=0 𝜃k = 1 ensures that only three rotational parameters are independent.

Angular Velocity Vectors In three-dimensional motion analysis, angular velocities are


not time derivatives of orientation parameters. Nonetheless, angular velocities can be writ-
ten as linear functions of time derivatives of orientation coordinates. In this section, a gen-
eral expression for the angular velocity vector is obtained using the orthogonality property
of the transformation matrix.
T
Because transformation matrix Ai is an orthogonal matrix, one has Ai Ai = I. It follows
( T )T
that Ȧ i Ai + Ai Ȧ i = 𝟎. This equation shows that Ȧ i Ai = −Ai Ȧ i = − Ȧ i Ai . A matrix
T T T T

that is equal to the negative of its transpose must be a skew-symmetric matrix. Therefore,
one can write Ȧ i Ai = 𝛚 ̃ i , where 𝛚
̃ i is a skew-symmetric matrix associated with vector 𝝎i ,
T
96 Mathematical Foundation of Railroad Vehicle Systems

which is the angular velocity vector defined in the global coordinate system. It is clear from
̃ i = Ȧ i Ai that if the transformation matrix is known, the angular velocity
T
the equation 𝛚
vector can be defined.
T
Alternatively, one can use the orthogonality of the transformation matrix to write Ai Ai =
( )
I. By differentiating this equation, one can show that Ai Ȧ i = − Ai Ȧ i , which demon-
T T T

̃ i , where 𝛚
strates that Ai Ȧ i is a skew-symmetric matrix that can be written as Ai Ȧ i = 𝛚
T T
̃i
is a skew-symmetric matrix associated with vector 𝛚 , which is the angular velocity vec-
i

tor defined in the body coordinate system. Therefore, one can write the derivative of the
transformation matrix using one of the following forms:
Ȧ i = 𝛚
̃ i Ai , ̃i
Ȧ i = Ai 𝛚 (3.23)
This equation shows that = ̃i
𝛚 ̃ i AiT .
Ai 𝛚
The two representations of Eq. 23 can be used to
obtain two different expressions for the absolute velocity vector of an arbitrary point on
the body. Equation 23 can be used to show that Ȧ i 𝛚i = 𝟎

Absolute Velocity Vector In order to obtain the absolute velocity vector of an arbitrary
point on rigid body i, Eq. 4, ri = Ri + Ai ui , can be differentiated with respect to time to
obtain ṙ i = Ṙ i + Ȧ i ui , where Ṙ i is the absolute velocity of the reference point. This equation,
upon using the first equation in Eq. 23, can be written as ṙ i = Ṙ i + 𝛚 ̃ i Ai ui . Because ui =
A u and 𝛚
i i ̃ u = 𝛚 × u , the absolute velocity vector of the arbitrary point can be written in
i i i i
[ ]T
terms of the angular velocity vector 𝛚i = 𝜔i1 𝜔i2 𝜔i3 as
ṙ i = Ṙ i + 𝛚i × ui (3.24)
Alternatively, one can substitute the second equation in Eq. 23 into velocity equation ṙ i =
Ṙ i + Ȧ i ui and follow a procedure similar to the one previously outlined in this section to
obtain an alternate expression for the absolute velocity vector in terms of the angular veloc-
[ ]T
ity vector 𝛚i = 𝜔i1 𝜔i2 𝜔i3 defined in the body coordinate system as
( )
ṙ i = Ṙ i + Ai 𝛚i × ui (3.25)
where 𝛚i = Ai 𝛚i .
[ ]T
Absolute Acceleration Vector The angular acceleration vector 𝛂i = 𝛼1i 𝛼2i 𝛼3i is
defined as the time derivative of angular velocity vector 𝛚i : that is, 𝛂i = 𝛚̇ i . Differentiat-
ing the velocity equation ṙ i = Ṙ i + 𝛚i × ui with respect to time, and using the definition of
the angular acceleration vector and the identity u̇ i = Ȧ i ui = 𝛚i × ui , one can show that the
absolute acceleration vector of the arbitrary point on the body can be written as
( )
̈ i + 𝛂i × ui + 𝛚i × 𝛚i × ui
r̈ i = R (3.26)
̈ i is the absolute acceleration vector of the body reference point. The term
In this equation, R
𝜶 × u on the right-hand side of Eq. 26 is the tangential component of the acceleration, while
i i

the term 𝛚i × (𝛚i × ui ) is the normal component, which is also referred to as the centripetal
acceleration. As in the case of the absolute velocity vector, the absolute acceleration vector
can be expressed in terms of the angular velocity and acceleration vectors defined in the
Motion and Geometry Descriptions 97

body coordinate system as


( ) ( ( ))
̈ i + Ai 𝛂i × ui + Ai 𝛚i × 𝛚i × ui
r̈ i = R (3.27)
where 𝛂i = Ai 𝛂i . The absolute acceleration vector of an arbitrary point on the body can be
used to define the virtual work of inertia forces, as discussed in Chapter 6.

3.6 GENERALIZED COORDINATES


Because Ȧ i = 𝛚 ̃ i (Eq. 23), regardless of the orientation parameters used,
̃ i Ai and Ȧ i = Ai 𝛚
vectors 𝛚 and 𝛚 are always linear functions in the time derivatives of the orientation
i i

parameters 𝛉i . Therefore, regardless of the orientation parameters used, one can always
write angular velocity vectors 𝛚i and 𝛚i in terms of the time derivatives of the orientation
parameters, respectively, as
𝛚i = Gi 𝛉̇ i , 𝛚i = Gi 𝛉̇ i (3.28)
where coefficient matrices Gi and Gi are expressed in terms of orientation parameters 𝛉i . For
[ ]T
example, in the case of using Euler angles 𝛉i = 𝜓 i 𝜙i 𝜃 i with the sequence Z i − X i − Y i ,
matrices Gi and Gi are
⎡0 cos 𝜓 i − sin 𝜓 i cos 𝜙i ⎤ ⎡− cos 𝜙i sin 𝜃 i cos 𝜃 i 0⎤
⎢ ⎥ ⎢ ⎥
Gi = ⎢0 sin 𝜓 i cos 𝜓 i cos 𝜙i ⎥ , Gi = ⎢ sin 𝜙i 0 1⎥ (3.29)
⎢ ⎥ ⎢ ⎥
⎣1 0 sin 𝜙i ⎦ ⎣ cos 𝜙i cos 𝜃 i sin 𝜃 i 0⎦

Using the equations 𝛚 ̃ i = Ȧ i Ai and 𝛚


T
̃ i = AiT Ȧ i to determine the angular velocity vectors
can be very cumbersome. Such an approach can be avoided by using a much simpler
approach based on a physical interpretation of the columns of the two matrices Gi and Gi .
While finite rotations are not commutative and cannot be added, the angular velocities
can be added and can be treated as vectors. Therefore, the angular velocity vectors can be
written as 𝛚i = gi1 𝜓̇ i + gi2 𝜙̇ i + gi3 𝜃̇ i and 𝛚i = gi1 𝜓̇ i + gi2 𝜙̇ i + gi3 𝜃̇ i . In these two equations,
gik and gik , k = 1, 2, 3, are unit vectors defined, respectively, in the global and body coordi-
nate systems. Vectors gik and gik , k = 1, 2, 3, are the axes about which the Euler rotations
[ ] [ ]
are performed. It follows that Gi = gi1 gi2 gi3 and Gi = gi1 gi2 gi3 . Therefore, instead
of using the equations 𝛚 ̃ i = Ȧ i Ai and 𝛚
T
̃ i = AiT Ȧ i to define vectors 𝝎i and 𝛚i , respectively,
which can be used to determine the two matrices Gi and Gi , these matrices can be easily
determined by recognizing that the columns of Gi are unit vectors, defined in the global
coordinate system, about which the three Euler rotations are performed; and the columns
of matrix Gi are the same vectors defined in the body coordinate system. Consequently,
Gi = Ai Gi .
[ ]T
For the four Euler parameters 𝛉i = 𝜃0i 𝜃1i 𝜃2i 𝜃3i , matrices Gi and Gi are 3 × 4 matri-
ces defined as
Gi = 2Ei , Gi = 2Ei (3.30)
98 Mathematical Foundation of Railroad Vehicle Systems

where matrices Ei and Ei are defined in Eq. 20. The columns of matrices Gi and Gi should
not be interpreted the same way as for Euler angles, since Euler parameters are quaternions
and are not interpreted as rotations about axes of rotation as is the case with Euler angles.

[ ]
Example 3.5 In this example, the columns of matrices Gi = gi1 gi2 gi3 and
[ ]
Gi = gi1 gi2 gi3 are determined based on physical considerations instead of using
the cumbersome approach that requires the time derivatives of the Euler-angle
[ ]T
transformation matrix. The case of the three Euler angles 𝛉i = 𝜓 i 𝜙i 𝜃 i with the
sequence Z i − X i − Y i is considered in this example. The first rotation 𝜓 i is performed
about the Z i axis, which is defined in the global coordinate system XYZ by the unit
[ ]T
vector gi1 = 0 0 1 . The second rotation 𝜙i is performed about the new body
X i axis, which is defined in the global coordinate system XYZ by the unit vector
[ ]T [ ]T
gi2 = Ai𝜓 1 0 0 = cos 𝜓 i sin 𝜓 i 0 , where the simple transformation matrix Ai𝜓 is
defined in Eq. 11. The last Euler rotation 𝜃 i is performed about the body Y i axis, which
can be defined in the global coordinate system XYZ after the first two rotations 𝜓 i and
[ ]T [ ]T
𝜙i by the unit vector gi3 = Ai𝜓 Ai𝜙 0 1 0 = − sin 𝜓 i cos 𝜙i cos 𝜓 i cos 𝜙i sin 𝜙i ,
where matrix Ai𝜙 is given in Eq. 11. The three vectors gik , k = 1, 2, 3, form the columns
of matrix Gi in Eq. 29.
The columns gik , k = 1, 2, 3, of matrix Gi can be obtained in a similar manner, with
the understanding that these vectors are unit vectors along the axes of the Euler rota-
tions defined in the body coordinate system. Vector gi1 , about which the first rotation
𝜓 i is performed, is defined in the body coordinate system by the unit vector gi1 =
T T[ ]T [ ]T
Ai𝜃 Ai𝜙 0 0 1 = − cos 𝜙i sin 𝜃 i sin 𝜙i cos 𝜙i cos 𝜃 i . Vector gi2 , about which the
second rotation 𝜙i is performed, is defined in the body coordinate system by the unit vec-
T[ ]T [ ]T
tor gi2 = Ai𝜃 1 0 0 = cos 𝜃 i 0 sin 𝜃 i . Vector gi3 , about which the third rotation
𝜃 i is performed, is simply defined in the body coordinate system by the unit vector gi3 =
[ ]T
0 1 0 . The three vectors gik , k = 1, 2, 3, form the columns of matrix Gi in Eq. 29.

[ ]
Example 3.6 In this example, the columns of the matrices Gi = gi1 gi2 gi3 and Gi =
[ i ]
g1 gi2 gi3 are determined again using a sequence different from that used in the pre-
ceding example. As in the preceding example, these columns are determined based on
physical considerations instead of using the cumbersome approach that requires the
time derivatives of the Euler-angle transformation matrix. In this example, the case of
[ ]T
the three Euler angles 𝛉i = 𝜙i 𝜃 i 𝜓 i with the sequence Z i − X i − Z i is considered.
This sequence, which was used by Euler to study gyroscopic motion, was considered in
Example 2, where the simple rotation matrices used to define the Euler angle transfor-
mation matrix are provided.
As discussed in Example 2, the first rotation 𝜙i is performed about the Z i axis, which is
[ ]T
defined in the global coordinate system XYZ by the unit vector gi1 = 0 0 1 . The sec-
ond rotation 𝜃 i is performed about the new body X i axis, which is defined in the global
Motion and Geometry Descriptions 99

[ ]T [ ]T
coordinate system XYZ by the unit vector gi2 = Ai𝜙 1 0 0 = cos 𝜙i sin 𝜙i 0 ,
where the simple transformation matrix Ai𝜙 is defined in Example 2. The last Euler
rotation 𝜓 i is performed about the body Z i axis, which can be defined in the global
coordinate system XYZ after the first two rotations 𝜙i and 𝜃 i by the unit vector
[ ]T [ ]T
gi3 = Ai𝜙 Ai𝜃 0 0 1 = sin 𝜃 i sin 𝜙i − sin 𝜃 i cos 𝜙i cos 𝜃 i , where matrix Ai𝜃 is given
in Example 2. The three vectors gik , k = 1, 2, 3, form the columns of matrix Gi :
that is,
⎡0 cos 𝜙i sin 𝜃 i sin 𝜙i ⎤
⎢ ⎥
Gi = ⎢0 sin 𝜙i − sin 𝜃 i cos 𝜙i ⎥
⎢ ⎥
⎣1 0 cos 𝜃 i ⎦
The columns gik , k = 1, 2, 3, of matrix Gi can be obtained in a similar manner,
with the understanding that these vectors are unit vectors along the axes of the
Euler rotations defined in the body coordinate system. Vector gi1 , about which the
first rotation 𝜙i is performed, is defined in the body coordinate system by the unit
T T[ ]T [ ]T
vector gi1 = Ai𝜃 Ai𝜓 0 0 1 = sin 𝜃 i sin 𝜓 i sin 𝜃 i cos 𝜓 i cos 𝜃 i , where the sim-
ple rotation matrix Ai𝜓 is defined in Example 2. Vector gi2 , about which the second
rotation 𝜃 i is performed, is defined in the body coordinate system by the unit vector
T[ ]T [ ]T
gi2 = Ai𝜓 1 0 0 = cos 𝜓 i − sin 𝜓 i 0 . Vector gi3 , about which the third rotation
𝜓 i is performed, is simply defined in the body coordinate system by the unit vector
[ ]T
gi3 = 0 0 1 . The three vectors gik , k = 1, 2, 3, form the columns of matrix Gi , which
can be written as
⎡ sin 𝜃 i sin 𝜓 i cos 𝜓 i 0⎤
G = ⎢sin 𝜃 i cos 𝜓 i − sin 𝜓 i 0⎥
i
⎢ ⎥
⎣ cos 𝜃 i 0 1⎦

Euler Parameter Identities Because 𝛚i = Gi 𝛉̇ i , the angular acceleration vector can be


written as 𝛂i = 𝛚̇ i = Gi 𝛉̈ i + Ġ i 𝛉̇ i . Therefore, the angular acceleration vector 𝜶 i has, in gen-
eral, two terms: one linear in the second derivative of the orientation coordinates, and the
other quadratic in the first derivative of these orientation coordinates. It is clear, however,
from the identities presented in Table 1, that when using Euler parameters Ġ i 𝛉̇ i = 2Ė i 𝛉̇ i = 𝟎,
and in this case, the angular acceleration vector reduces to 𝛂i = 𝛚̇ i = Gi 𝛉̈ i . However, this is
not, the case when Euler angles are used. Similarly, one has Ġ i 𝛉̇ i = 2Ė i 𝛉̇ i = 𝟎, and 𝛂i = Gi 𝛉̈ i
when Euler parameters are used.
100 Mathematical Foundation of Railroad Vehicle Systems

Derivatives with Respect to Orientation Parameters It was previously shown


in this chapter that the absolute velocity vector can be written as ṙ i = Ṙ i + 𝛚i × ui or ṙ i =
( )
Ṙ i + Ai 𝛚i × ui . These two equations can be ( written
̇i ̇i
) in the forms r = R − u × 𝛚 = R −
i i ̇i
( )
̃ i Gi 𝛉̇ i . Using these equations, one can show
̃ i Gi 𝛉̇ i and ṙ i = Ṙ i − Ai ui × 𝛚i = Ṙ i − Ai u
u
that the absolute velocity and acceleration vectors of an arbitrary point on the body can be
written in terms of generalized velocities q̇ i as
ṙ i = Li q̇ i , r̈ i = Li q̈ i + L̇ i q̇ i (3.31)
[ T T ]T
where qi = Ri 𝛉i , and
[ ] [ ]
Li = I −̃ ui Gi = I −Ai u ̃ i Gi (3.32)

One can show that


(( ) )
( i )2 i 2
L̇ i q̇ i = 𝛚
̃ u −u ̃ i Ġ i 𝛉̇ i = Ai 𝛚 ̃ i Ġ i 𝛉̇ i
̃ i ui − u (3.33)

This is with the understanding that when the four Euler parameters are used, Ġ i 𝛉̇ i = 𝟎 and
Ġ i 𝛉̇ i = 𝟎. It is also clear from the preceding equations that
𝜕ri 𝜕 ṙ i 𝜕 r̈ i ⎫
= = = I,
𝜕Ri 𝜕 Ṙ i 𝜕R ̈i ⎪
⎬ (3.34)
𝜕r i 𝜕 ṙ i 𝜕 r̈ i
i i i ̃i i ⎪
= = = −̃u G = −A u G ⎭
𝜕𝛉i 𝜕 𝛉̇ i 𝜕 𝛉̈ i
The equations presented in this section will be used in Chapter 6 to develop the dynamic
equations of a railroad vehicle system by developing the expressions for the generalized
inertia and applied forces. Equation 34 is also used to define the Jacobian matrix of the
kinematic constraint equations that describe mechanical joints and specified motion tra-
jectories.

3.7 KINEMATIC SINGULARITIES


Kinematic singularities are encountered whenever three parameters are used to describe
a body orientation in space. Kinematic singularities can take different forms depending
on the parameters used. When using the three Euler angles, singularities appear in the
two matrices Gi and Gi used to define the angular velocity vector as 𝛚i = Gi 𝛉̇ i and 𝛚i =
Gi 𝛉̇ i , respectively. At the singular configurations, one cannot solve for 𝛉̇ i in terms of the
components of the angular velocities. The configurations at which the singularities occur
depend on the sequence of Euler angles used. Therefore, these singularities must be viewed
as mathematical singularities due to the techniques used in the motion description; such
singularities have nothing to do with the physics of the problem.
For the sequence of Euler angles Z i − X i − Y i used in railroad vehicle system dynamics,
the singularities in the two matrices Gi and Gi of Eq. 29 appear when 𝜙i = ± 𝜋/2. One
can show that at this configuration, the Z i and Y i axes of rotation become parallel, and
Motion and Geometry Descriptions 101

these two axes of Euler rotations are not independent. At this configuration, one cannot
determine the derivatives of Euler angles from the angular velocities using the equations
𝛚i = Gi 𝛉̇ i and 𝛚i = Gi 𝛉̇ i . Determining 𝛉̇ i is necessary during dynamic simulations since the
angular velocity vectors 𝝎i and 𝛚i are not exact differentials and cannot be integrated to
determine rotational coordinates. In fact, there are no known rotational coordinates such
that the matrices Gi and Gi are the identity matrices: therefore, 𝛚i ≠ 𝛉̇ i and 𝛚i ≠ 𝛉̇ i .
In railroad vehicle dynamics, the Euler-angle sequence Z i − X i − Y i is used to avoid these
kinematic singularities in most simulations. Figure 6 shows the coordinate system of a
wheelset in which the Y i axis is selected along the wheel axle. In most railroad vehicle
system problems, the yaw angle 𝜓 i and the roll angle 𝜙i remain small. Therefore, in these
simulation scenarios, it is not expected that the roll angle 𝜙i will approach the value ±𝜋/2
at which the singularity occurs.

ϕ i = π/2 θi

ψi Yi ϕi Yi
Z Zi Zi
Yi Yi Xi Xi
Y
Xi Xi
O
X Aiψ Aiψ Aiϕ Aiψ Aiϕ Aiθ

Figure 3.6 Euler-angle singularities.

The singular configurations depend on the sequence of Euler angles used, as previously
mentioned. If the sequence Z i − X i − Z i of Examples 2 and 6 is used with the angles 𝜙i ,
𝜃 i , and 𝜓 i , respectively, the singular configuration occurs when 𝜃 i = 0 or 𝜃 i = 𝜋. At this
configuration, one cannot distinguish between the first and second Z i axes about which
the Euler rotations 𝜙i and 𝜓 i are performed, respectively. One can show that if 𝜃 i = 0 or
𝜃 i = 𝜋, the two matrices Gi and Gi obtained in Example 6, which have the determinants
|Gi | = ||Gi || = − sin 𝜃 i , are singular, and the derivatives of Euler angles cannot be determined
| | | |
from the angular velocity vectors.
No such kinematic singularities are associated with Euler parameters. When Euler
parameters are used, one can still write the angular velocity vectors as 𝛚i = Gi 𝛉̇ i and
𝛚i = Gi 𝛉̇ i , where Gi = 2Ei and Gi = 2Ei . In order to solve for 𝛉̇ i using, for example, the
equation 𝛚i = Gi 𝛉̇ i , one can pre-multiply by Gi to obtain Gi 𝛚i = Gi Gi 𝛉̇ i . Using the iden-
T T T

( T)
tities of Euler parameters presented in Table 1, one has G 𝛚 = 4 I4 − 𝛉i 𝛉i 𝛉̇ i . Because
iT i

𝛉i 𝛉̇ i = 0, one has 𝛉̇ i = (1∕4) Gi 𝛚i for any values of Euler parameters. Consequently, when
T T

Euler parameters are used, the kinematic singularities associated with Euler angles can
be avoided. This is one of the main reasons Euler parameters are often used in developing
general MBS algorithms.
102 Mathematical Foundation of Railroad Vehicle Systems

Example 3.7
[ ]T
The orientation of a body i is defined by the Euler parameters 𝛉i = 0 0 1 0 . At
this configuration, the absolute angular velocity vector in the global coordinate system
[ ]T
is given by 𝛚i = 20.0 0 5.0 . Determine the time derivatives of the Euler parameters.
Solution The absolute angular velocity vector defined in the global XYZ coor-
dinate system is written in terms of the derivatives of Euler parameters as
𝛚i = Gi 𝛉̇ i = 2Ei 𝛉̇ i . Pre-multiplying both sides of this equation by Gi , and using
T

T ( T )
the identity Gi Gi = 4 I4 − 𝛉i 𝛉i given in Table 1 (Nikravesh 1988; Shabana et al.
2001), one can write 𝛉̇ i = (1∕4) Gi 𝛚i = (1∕2) Ei 𝛚i , as previously explained. Therefore,
T T

one has
⎡−𝜃1i −𝜃2i −𝜃3i ⎤
⎢ ⎥ ⎡𝜔i1 ⎤
⎢ 𝜃0i 𝜃 i
−𝜃 i⎥⎢ ⎥
1 T 1
𝛉̇ i = Ei 𝛚i = ⎢ ⎥ ⎢𝜔i ⎥
3 2
2 2 ⎢−𝜃 i 𝜃 i 𝜃1i ⎥ ⎢ ⎥
2
⎢ 3 0 ⎥ ⎢𝜔i ⎥
⎢ 𝜃 i −𝜃 i 𝜃 i ⎥ ⎣ 3 ⎦
⎣ 2 1 0 ⎦

⎡0 1 0⎤ ⎡ 0 ⎤
⎢ ⎥ ⎡20.0⎤ ⎢ ⎥
⎢ 0 −1⎥ ⎢ ⎥ ⎢−2.5⎥
1 0
= ⎢ ⎥⎢ 0 ⎥ = ⎢ ⎥
2 ⎢0 0 0 ⎥⎢ ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ 5.0 ⎥ ⎢ ⎥
⎢1 ⎣ ⎦ ⎢
⎣ 0 0 ⎥⎦ ⎣ 10 ⎥

3.8 EULER ANGLES AND TRACK GEOMETRY


In railroad vehicle system dynamics, Euler angles can be used for two fundamentally dif-
ferent purposes, as previously mentioned. First, Euler angles can be used to define the
orientation of a body in space, as explained in the preceding sections of this chapter. In
this case, three discrete Euler angles are assigned to each body in the system, and these
Euler angles are not treated as field variables since the body orientation is a characteristic
of the entire body and does not depend on a particular point. Euler angles used in motion
description vary with time as the bodies change their orientations due to the applied forces
and moments.
For the second purpose, Euler angles are used to represent the geometry of the track and
rail space curves without consideration of the track and rail motion. While each rail can be
modeled as a separate body that can move independently, and can be assigned its discrete
Euler angles to define its orientation, another set of field-variable Euler angles based on
another sequence can be used to develop the geometry of the space curves associated with
each rail, as described in the remainder of this chapter. To understand the process of defin-
ing track geometry, it is important to understand the basic sections and definitions used in
practice when the track layout is designed.
Motion and Geometry Descriptions 103

Track Sections The rail track is designed using a few sections with specific geome-
tries. In the layout of the track, it necessary to ensure smoothness during the transition
between sections that have different geometries, in order to avoid undesirable impulsive
forces, vibration, noise, and possibly derailments. Non-smooth transitions can lead to sud-
den changes in the wheel/rail contact forces, and such impulsive forces not only produce
undesirable effects such as noise, severe vibrations, and ride discomfort, but can also cause
derailments and serious accidents that can be deadly and result in environmental and eco-
nomic damage.
In general, three different types of segments with different geometries are used in the
layout of the rail track. These segments are
1. Tangent (straight), which has zero curvature
2. Curve, which has constant curvature
3. Spiral, which has linearly varying curvatures
The geometries of these three segments are shown in Figure 7. The spiral segment is used
to connect two segments that have two different curvature values. The spiral is designed to
have linearly varying curvature, with the spiral ends having the curvatures of the two seg-
ments (tangent or curve) connected by the spiral. For example, to connect a tangent section
of track that has zero curvature to a curve track segment with non-zero curvature, the spiral
connecting these two segments has zero curvature at the first end and the curvature of the
curve section at the second end. Within the spiral, the curvature varies linearly; an example
is shown in Figure 8.

Sprial Tangent
curve-to-tangent

Curve

Sprial
curve-to-curve

Curve
Sprial
tangent-to-curve
Tangent

Figure 3.7 Track segments.

Track Geometry Parameters As also discussed in Chapter 1, important geometric


parameters are used in the construction of the track. These are the track gage G, which
measures the spacing between the two rails of the track and is defined as the lateral dis-
tance between two points on the right and left rails located at a distance 5/8 in. (14 mm)
104 Mathematical Foundation of Railroad Vehicle Systems

Curve

Spiral
tangent-to-curve
Tangent

Figure 3.8 Spiral section.

from the top of the railhead, as shown in Figure 9. The gage can have different values in
different countries. The standard gage value in North America varies from 56 to 57.25 in.
The superelevation h, also shown in Figure 9, is defined as the vertical distance between
the right and left rails. The horizontal curvature is defined as the angle 𝜓 required to obtain
a 100′ -length chord of constant radius of curvature RH in the horizontal plane, as shown
in Figure 10. The grade is the ratio, given as a percentage, between the vertical elevation
and the longitudinal distance on a rail. The cant angle measures the rotation of a rail about
its longitudinal axis, as shown in Figure 11. Rail irregularities are defined using two types
of deviations: profile and alignment. The profile and alignment deviations are, respectively,
the vertical and lateral deviations of the rail space curve, as shown in Figure 12.

Track plane
w
G
ϕ h

Horizontal plane

Figure 3.9 Gage and superelevation.

Track Geometry Definitions In practice, very few inputs are provided to obtain the
discretized data used in computer simulations and virtual prototyping of railroad vehicle
systems. These data can be provided at the points on the rail at which the geometry changes
(tangent to spiral, spiral to curve, spiral to tangent, etc.). As explained in Chapter 4, in
addition to the distance (arc length) at which these points are located, three pieces of infor-
mation are needed at these geometry-change points in order to define the track geometry in
a track preprocessor computer program. The geometry data produced by the track prepro-
cessor at every nodal point include the distance, the Cartesian coordinates of the nodes in
Motion and Geometry Descriptions 105

P1

10
0′
RH s
Y

ψ
P2

O X

Figure 3.10 Horizontal curvature.

Cant angle

Figure 3.11 Cant angle.

a selected track or rail coordinate system, and three Euler angles that define the geometry
of the rail. In order for the track preprocessor to compute the positions of and the angles
at the nodes defined by the distance specified, the following three pieces of given industry
inputs are provided at the points at which there is a change in the track geometry:

Alignment
Chord length

Profile (vertical deviation)


Z
X

X
Chord length

Figure 3.12 Track irregularities.


106 Mathematical Foundation of Railroad Vehicle Systems

1. Horizontal curvature measures the curvature of the rail or track space curve when pro-
jected on a horizontal plane, as shown in Figure 10. The rail space curve is assumed to
have arc length s, while the projected curve is parameterized by the arc length S. The hor-
izontal curvature is used to define a Euler angle 𝜓, used as a measure of the curvature
of a curve or spiral segment.
2. Superelevation measures the vertical elevation between the two rails. The superelevation
is used to define another Euler angle called the bank angle 𝜙.
3. Grade or development is the ratio between the vertical elevation and the longitudinal
distance along the space curve. The grade is used to define a third Euler angle, called the
vertical - development (elevation) angle 𝜃.
Using these three inputs, the three Euler angles 𝜓, 𝜙, and 𝜃 can be obtained as field variables
that depend on the arc length of the space curve. It is important to point out that these
Euler angles are not used to describe the orientation of bodies, but are used to define the
geometry of space curves. These angles have a more general interpretation as the elements
of position-gradient vectors, as will be discussed in Chapter 4. Therefore, the Euler angles
considered in this section and the remaining sections of this chapter are used to define
geometry and are not used in the motion description. Their use as geometry variables gives
physical meaning to the basic steps used in the design of the track layout.

Track Euler Angles To develop a general approach for modeling railroad vehicle
systems, the geometry of three space curves needs to be defined: the track centerline,
right rail, and left rail space curves. Each of these space curves can have its own geometry.
As discussed in Chapter 4, a track preprocessor computer program can be designed to
generate the data for these three space curves for use in nonlinear dynamic simulations of
railroad vehicle systems.
The sequence of rotation used for the three Euler angles that define the geometry of the
track or rail space curves is different from the sequence previously used in this chapter to
describe a body orientation. The reason for using a different sequence is to be consistent
with the positive sign definitions of the curvature, superelevation, and grade used in prac-
tice. In this book, the X axis is considered the longitudinal axis, the Y axis is the lateral axis,
and the Z axis is the axis perpendicular to the plane of the track. The sequence of rotations
used for Euler angles that define the track geometry is a rotation 𝜓 about the Z axis, a rota-
tion 𝜃 about the −Y axis, and a rotation 𝜙 about the −X axis. Using this sequence of Euler
angles, one can write the following simple rotation matrices:

⎡cos 𝜓 − sin 𝜓 0⎤ ⎡cos 𝜃 0 − sin 𝜃 ⎤ ⎡1 0 0 ⎤


⎢ ⎥ ⎢ ⎥ ⎢ ⎥
A𝜓 = ⎢ sin 𝜓 cos 𝜓 0⎥ , A𝜃 = ⎢ 0 1 0 ⎥, A𝜙 = ⎢0 cos 𝜙 sin 𝜙 ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 0 0 1⎦ ⎣ sin 𝜃 0 cos 𝜃 ⎦ ⎣0 − sin 𝜙 cos 𝜙⎦
(3.35)
In the railroad literature, the angle 𝜓 is called the horizontal curvature angle, the angle
𝜃 is called the vertical - development (elevation) angle, and the angle 𝜙 is called the bank
angle. When the track geometry is considered, these angles are not called the yaw, pitch,
and roll angles. The horizontal curvature angle 𝜓, elevation angle 𝜃, and bank angle 𝜙 can
be obtained from a simple track geometry description, as discussed in later sections of this
Motion and Geometry Descriptions 107

chapter and in Chapter 4. Using the simple rotation matrices in the preceding equation,
one can define the following transformation matrix:
A = A𝜓 A𝜃 A𝜙

⎡cos 𝜓 cos 𝜃 − sin 𝜓 cos 𝜙 + cos 𝜓 sin 𝜃 sin 𝜙 − sin 𝜓 sin 𝜙 − cos 𝜓 sin 𝜃 cos 𝜙⎤
⎢ ⎥
= ⎢ sin 𝜓 cos 𝜃 cos 𝜓 cos 𝜙 + sin 𝜓 sin 𝜃 sin 𝜙 cos 𝜓 sin 𝜙 − sin 𝜓 sin 𝜃 cos 𝜙 ⎥
⎢ ⎥
⎣ sin 𝜃 − cos 𝜃 sin 𝜙 cos 𝜃 cos 𝜙 ⎦
(3.36)
This transformation can be used to completely define the geometry of the track and rail
space curves. To this end, the Euler angles 𝜓, 𝜃, and 𝜙 must be expressed as field variables,
as explained in a later section of this chapter. The first column of this matrix is assumed to
be the unit tangent to the space curve parameterized by the arc length s. In this case, one
can write 𝜓 = 𝜓(s), 𝜃 = 𝜃(s), and 𝜙 = 𝜙(s).

3.9 ANGLE REPRESENTATION OF THE CURVE


GEOMETRY
When Euler angles are used as time-dependent motion coordinates to describe a body orien-
tation, these angles are assumed to be independent coordinates in the case of unconstrained
motion. In curve geometry, on the other hand, the Frenet frame that completely defines the
curve geometry is a function of only one parameter, which can be the curve arc length. That
is, the Frenet frame, which is defined by three orthogonal unit vectors, can be completely
defined using the arc length parameter. Therefore, if Euler angles employed in develop-
ing the transformation of Eq. 36 are used to define the space curve geometry, these Euler
angles must be related, and one should be able to write them as functions of the arc length.
By so doing, the columns of the transformation matrix in Eq. 36 become the same as the
columns of the transformation matrix that defines the Frenet frame, with the exception that
the matrix of Eq. 36 is expressed in angles that have the physical meaning used in practice
by the rail industry in the construction of a track layout.
A track or rail space curve r can be defined by considering the first column in the trans-
formation matrix of Eq. 36 as the unit tangent vector. Assuming that curve r is given in its
parametric form in terms of the curve arc length parameter s, one can use the first column
of the transformation matrix of Eq. 36 to write a segment of the space curve as
( ) ⎡cos 𝜓 cos 𝜃 ⎤
dr ⎢ ⎥
dr = ds = ⎢ sin 𝜓 cos 𝜃 ⎥ ds (3.37)
ds ⎢ ⎥
⎣ sin 𝜃 ⎦
The curvature vector to the curve r is obtained by differentiating the unit tangent vector
[ ]T
t (s) = dr∕ds = cos 𝜓cos 𝜃 sin 𝜓cos 𝜃 sin 𝜃 with respect to the arc length s as
⎡−𝜓 ′ sin 𝜓 cos 𝜃 − 𝜃 ′ cos 𝜓 sin 𝜃 ⎤
dt ⎢ ⎥
r′′ (s) = = 𝜓 ′ cos 𝜓 cos 𝜃 − 𝜃 ′ sin 𝜓 sin 𝜃 ⎥
ds ⎢⎢
(3.38)

⎣ 𝜃 ′ cos 𝜃 ⎦
108 Mathematical Foundation of Railroad Vehicle Systems

In this equation, 𝜓 ′ = 𝜕𝜓/𝜕s and 𝜃 ′ = 𝜕𝜃/𝜕s. The curvature, which is the magnitude of this
curvature vector, is defined as

| dt |
𝜅 = ||r′′ (s)|| = || || = (𝜓 ′ cos 𝜃)2 + (𝜃 ′ )2 (3.39)
| ds |
Therefore, the unit normal vector to the curve r defined by Eq. 37 is along the curvature
vector and is defined as
⎡−𝜓 ′ sin 𝜓 cos 𝜃 − 𝜃 ′ cos 𝜓 sin 𝜃 ⎤
1⎢ ⎥
n = ⎢ 𝜓 ′ cos 𝜓 cos 𝜃 − 𝜃 ′ sin 𝜓 sin 𝜃 ⎥ (3.40)
𝜅⎢ ⎥
⎣ 𝜃 ′ cos 𝜃 ⎦
As explained in Chapter 2, the binormal vector b is the cross product of the tangent vector
t and the normal vector n. This cross product leads to
⎡−𝜓 ′ cos 𝜓 sin 𝜃 cos 𝜃 + 𝜃 ′ sin 𝜓 ⎤
1⎢ ⎥
b = t × n = ⎢−𝜓 ′ sin 𝜓 sin 𝜃 cos 𝜃 − 𝜃 ′ cos 𝜓 ⎥ (3.41)
𝜅⎢ ⎥
⎣ 𝜓 ′ cos2 𝜃 ⎦
The Frenet frame Af is formed by the three orthogonal unit vectors t, n, and b: that is,
[ ]
Af = t n b (3.42)
As discussed in Chapter 2, performing additional differentiations leads to the Serret–
Frenet formulas defined as
t′ = 𝜅n ⎫

n = −𝜅t + 𝜏b⎬
′ (3.43)
b′ = −𝜏n ⎪

where 𝜏 is the curve torsion, which is defined using the equation b′ (s) = − 𝜏(s)n(s).
In general, the two orthogonal matrices of Eqs. 36 and 42 share the first column. The
other two columns are not, in general, the same, and they differ by a simple planar rotation
about the tangent vector. In order for the two orthogonal matrices A and Af of Eqs. 36 and
42, respectively, to be the same, the Euler angles must be related because a curve geometry
is defined by two geometric invariants only. This allows using matrix A of Eq. 36 to define
the curve geometry using angles that can be given a physical meaning.

3.10 EULER ANGLES AS FIELD VARIABLES


By imposing conditions on the Euler angles, these angles can have a physical meaning and
can be directly related to definitions used by the rail industry, such as the horizontal curva-
ture, superelevation, and grade. Without these conditions, the three Euler angles 𝜓, 𝜃, and
𝜙 cannot be effectively used to define the geometry of the track.
In order for the two orthogonal matrices A and Af of Eqs. 36 and 42, respectively, to be the
same, the Euler angles must be related. As previously explained, because the matrices A and
Af share the first column, the other two columns in the two matrices lie in the same plane
but differ by a simple rotation about the longitudinal tangent t. By imposing conditions on
Motion and Geometry Descriptions 109

the angle 𝜙, one can show that the second column in matrix A becomes the unit curvature
vector n, which is the second column of matrix Af . To this end, the angle 𝜙 is defined using
the equations (Rathod and Shabana 2006a,b)

−𝜃 ′ ⎫
−𝜃 ′
sin 𝜙 = √ = ⎪
(𝜓 ′ cos 𝜃) + 2
(𝜃 ′ )2 𝜅

𝜓 cos 𝜃 ⎬
(3.44)
𝜓 cos 𝜃
′ ′

cos 𝜙 = √ =
(𝜓 ′ cos 𝜃)2 + (𝜃 ′ )2 𝜅 ⎪

By using this equation in Eq. 40, the normal vector n can be written as
⎡−𝜓 ′ sin 𝜓 cos 𝜃 − 𝜃 ′ cos 𝜓 sin 𝜃 ⎤ ⎡− sin 𝜓 cos 𝜙 + cos 𝜓 sin 𝜃 sin 𝜙⎤
1⎢ ⎥ ⎢ ⎥
n = ⎢ 𝜓 ′ cos 𝜓 cos 𝜃 − 𝜃 ′ sin 𝜓 sin 𝜃 ⎥ = ⎢ cos 𝜓 cos 𝜙 + sin 𝜓 sin 𝜃 sin 𝜙 ⎥ (3.45)
𝜅⎢ ⎥ ⎢ ⎥
⎣ 𝜃 ′ cos 𝜃 ⎦ ⎣ − cos 𝜃 sin 𝜙 ⎦
which is the same as the second column of matrix A. In Eq. 45, the normal vector has a
form that is not a function of the derivatives of the angles. The conditions of Eq. 44, which
allowed eliminating the angle derivatives, implies that
( )
−𝜃 ′
𝜙 = tan−1 (3.46)
𝜓 ′ cos 𝜃
Similarly, by using Eq. 44 or, equivalently, Eq. 46, the binormal vector of Eq. 41 can be
written as
⎡−𝜓 ′ cos 𝜓 sin 𝜃 cos 𝜃 + 𝜃 ′ sin 𝜓 ⎤ ⎡− cos 𝜓 sin 𝜃 cos 𝜙 − sin 𝜓 sin 𝜙⎤
1⎢ ⎥ ⎢ ⎥
b = ⎢−𝜓 ′ sin 𝜓 sin 𝜃 cos 𝜃 − 𝜃 ′ cos 𝜓 ⎥ = ⎢− sin 𝜓 sin 𝜃 cos 𝜙 + cos 𝜓 sin 𝜙⎥ (3.47)
𝜅⎢ ⎥ ⎢ ⎥
⎣ 𝜓 ′ cos2 𝜃 ⎦ ⎣ cos 𝜃 cos 𝜙 ⎦
which is the same as the third column of matrix A of Eq. 36. Again, the angle derivatives
have been eliminated from one of the vectors that define the Frenet frame.
By eliminating the angle derivatives in the normal and binormal vectors, the second
equation in the Serret–Frenet equations (Eq. 43), n′ = − 𝜅t + 𝜏b, leads, upon using
the third element in vectors n′ , t and b, to the condition 𝜃 ′ sin 𝜃 sin 𝜙 − 𝜙′ cos 𝜃 cos 𝜙
= − 𝜅 sin 𝜃 + 𝜏 cos 𝜃 cos 𝜙. This equation can be simplified using Eq. 44 as 𝜙′ cos 𝜃 cos 𝜙 =
𝜅 sin 𝜃 − 𝜅 sin 𝜃sin2 𝜙 − 𝜏 cos 𝜃 cos 𝜙, or
𝜙′ cos 𝜃 = 𝜅 sin 𝜃 cos 𝜙 − 𝜏 cos 𝜃 (3.48)
Therefore, using Eqs. 44 and 48, one has the following three differential equations that can
be used to define Euler angles as field variables:
𝜓 ′ cos 𝜃 = 𝜅 cos 𝜙 ⎫

𝜃′ = −𝜅 sin 𝜙 ⎬ (3.49)

𝜙 cos 𝜃 = 𝜅 sin 𝜃 cos 𝜙 − 𝜏 cos 𝜃 ⎭

The derivatives of the angles can then be written as


𝜓 ′ = 𝜅 (cos 𝜙∕ cos 𝜃) ⎫

𝜃 ′ = −𝜅 sin 𝜙 ⎬ (3.50)

𝜙 = 𝜅 tan 𝜃 cos 𝜙 − 𝜏 ⎭

110 Mathematical Foundation of Railroad Vehicle Systems

These equations can be used to define the angles in terms of the curve curvature 𝜅 and the
curve torsion 𝜏, which uniquely define the curve geometry. It is clear from this equation
that a singularity exists when the vertical-development angle 𝜃 approaches ±(𝜋/2). Such
a high value of the grade, which corresponds to the vertical rail, is not used in practice.
The elevation and superelevation angles 𝜃 and 𝜙, respectively, do not assume very high
values in practical applications. If both of these angles remain very small, one has 𝜓 ′ ≈ 𝜅
and 𝜙′ ≈ − 𝜏: that is, the angle 𝜓 can be directly related to the curvature of the curve 𝜅, and
the angle 𝜙 can be directly related to the torsion of the curve 𝜏. For a more general curve
defined by the curvature 𝜅(s) and torsion 𝜏(s), the derivative of the angles in Eq. 50 can be
integrated to determine the angles as functions of the arc length parameter. To this end, one
can use Eq. 50 to write
s ⎫
𝜓 (s) = 𝜓o + ∫s 𝜅 (cos 𝜙∕ cos 𝜃) ds ⎪
o
s ⎪
𝜃 (s) = 𝜃0 − ∫s 𝜅 sin 𝜙ds ⎬ (3.51)
o
s ⎪

𝜙 (s) = 𝜙o + s (𝜅 tan 𝜃 cos 𝜙 − 𝜏) ds⎪
o

where 𝜓 o , 𝜃 o , and 𝜙o are specified values of the angles at so . In general, obtaining analytical
closed-form solutions for the integrals in Eq. 51 can be difficult, and therefore, numerical
integration methods can be employed to evaluate these integrals. Equation 51 shows that
Euler angles can all be written in terms of the curve arc length s since the curvature 𝜅
and torsion 𝜏 depend on the arc length parameter. Therefore, Euler angles are no longer
considered independent parameters and can be used to uniquely define the curve geometry.
Previously, it was shown that the √ curve curvature 𝜅 can be written in terms of Euler angles
and their derivatives as 𝜅 (s) = (𝜓 ′ cos 𝜃)2 + (𝜃 ′ )2 (Eq. 39). Using Eq. 50, the curve torsion
𝜏 can also be written in terms of Euler angles and their derivatives as
𝜏 (s) = 𝜅 tan 𝜃 cos 𝜙 − 𝜙′

= (𝜓 ′ cos 𝜃)2 + (𝜃 ′ )2 (tan 𝜃 cos 𝜙) − 𝜙′ (3.52)
That is, Euler angles uniquely define the geometric curve properties. It is also clear that once
the angles are determined, the curve can be defined in its parametric form by integrating
Eq. 37 as

( ) ⎡cos 𝜓 (s) cos 𝜃 (s)⎤


s
dr
s
⎢ ⎥
r=
∫so ds
ds =
∫so ⎢ sin 𝜓 (s) cos 𝜃 (s) ⎥ ds (3.53)
⎢ ⎥
⎣ sin 𝜃 (s) ⎦
In the layout of the track, Euler angles are expressed in terms of given specific industry
inputs, as discussed in the following section. It is to be noted that the bank angle 𝜙 that
enters into the definition of the curve geometry is different from the bank angle that defines
the track super-elevation and the orientation of the track coordinate systems that follow the
motion of the vehicle components in railroad vehicle algorithms (Ling and Shabana, 2020).
Motion and Geometry Descriptions 111

3.11 EULER-ANGLE DESCRIPTION OF THE TRACK


GEOMETRY
To understand the mathematical foundations of some of the terminology used in track
descriptions, some basic geometry concepts are discussed in this section, starting with the
curvature vector in Eq. 38. This vector can be written as the sum of two vectors as

⎡−𝜓 ′ sin 𝜓 cos 𝜃 − 𝜃 ′ cos 𝜓 sin 𝜃 ⎤ ⎡− sin 𝜓 ⎤ ⎡− cos 𝜓 sin 𝜃 ⎤


dt ⎢ ′ ⎥ ⎢ ⎥ ′⎢ ⎥
′′
r (s) = = ⎢ 𝜓 cos 𝜓 cos 𝜃 − 𝜃 sin 𝜓 sin 𝜃 ⎥ = 𝜓 cos 𝜃 ⎢ cos 𝜓 ⎥ + 𝜃 ⎢ − sin 𝜓 sin 𝜃 ⎥
′ ′
ds ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 𝜃 ′ cos 𝜃 ⎦ ⎣ 0 ⎦ ⎣ cos 𝜃 ⎦
(3.54)
This equation shows that the curvature vector has two components
[ ]T
𝜓 ′ cos 𝜃 and 𝜃 ′ along the two orthogonal unit vectors a1 = − sin 𝜓 cos 𝜓 0 and a2 =
[− cos 𝜓 sin 𝜃 − sin 𝜓 sin 𝜃 cos 𝜃]T , respectively. These two vectors are also orthogonal to
[ ]T
the longitudinal unit tangent to the curve t = cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 sin 𝜃 . Therefore,
the three vectors t, a1 , and a2 represent an orthogonal triad, with vector a1 lying in a plane
parallel to the horizontal plane and vector a2 perpendicular to the plane formed by the two
vectors t and a1 . In railroad literature, the curvature component 𝜃 ′ is called the vertical
curvature CV . Therefore, the vertical curvature is defined as
d𝜃
CV = 𝜃 ′ = (3.55)
ds
That is, the vertical curvature is defined to be the derivative of the elevation angle 𝜃 with
respect to the curve arc length parameter s.

Horizontal Projection and Curvature In the mathematical description of a railroad


track, the track and rail space curves are first projected on the horizontal plane. The hori-
zontal projection of the curve segment defined by Eq. 37 leads to
[ ]T
drH = cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 0 ds (3.56)
[ ]T
By factoring out cos𝜃, one can write drH = cos 𝜓 sin 𝜓 0 cos 𝜃ds. This equation
defines the unit tangent vector tH to and the arc length dS of the projected planar curve
shown in Figure 10, respectively, as
[ ]T }
tH = cos 𝜓 sin 𝜓 0
(3.57)
dS = cos 𝜃ds
Using the relationship between the space curve arc length s and the arc length S of the pro-
jected planar curve, one has 𝜕𝜓/𝜕S = (𝜕𝜓/𝜕s)(𝜕s/𝜕S), where 𝜕s/𝜕S = 1/ cos 𝜃. The curvature
vector of the projected planar curve that is parameterized by the arc length S is defined as
⎡− sin 𝜓 ⎤
𝜕tH ⎢ ⎥ 𝜕𝜓
cH = = ⎢ cos 𝜓 ⎥ (3.58)
𝜕S ⎢ ⎥ 𝜕S
⎣ 0 ⎦
112 Mathematical Foundation of Railroad Vehicle Systems

This equation defines the horizontal curvature CH of the projected planar curve, which is
the magnitude of the curvature vector cH , as CH = |cH | = 𝜕𝜓/𝜕S, or
dS
d𝜓 = CH dS = (3.59)
RH
In this equation, RH is the radius of curvature of the projected planar curve shown in
Figure 10. It is clear from the preceding equation that the angle 𝜓 can be determined by
integration of the horizontal curvature CH . In this case, one can write
S ( )
𝜓 = 𝜓 (S) = 𝜓o + CH (S) dS (3.60)
∫So
where 𝜓 o is the value of 𝜓 at So . The horizontal curvature CH , as discussed in Chapter 4, is
one of the standard data used in practice for the track description. In the preceding equation,
for convenience, the horizontal curvature angle 𝜓 is written in terms of the projected-curve
arc length S instead of the actual curve arc length s. Recall from Eq. 57 that S and s are
related by dS = cos 𝜃ds.
In the design of railroad tracks, the chord definition of curvature is used. The chord defi-
nition of curvature is based on the number of degrees encompassed by a 100-ft line segment
that has endpoints located on the arc of the track curve. An approximate method for mea-
suring the curvature is to stretch a string that has length 62 ft between two points on the
inside face of the outer railhead and determine the number of inches between the center
point of the string and the rail. This number corresponds to the degree of curvature: 1 in.
equals 1∘ and 2 in. equals 2∘ . The sharpest curve that can be handled by a single four-axle
diesel locomotive is 40∘ . When the locomotive is in a train consist, the limit decreases to
20∘ . Uneven terrain, such as mountains, further limits the curvature to 5–10∘ . Curves of 1∘
or 2∘ are the most commonly used (McGonigal, 2006).

Elevation and Grade The vertical-development angle 𝜃 can be determined from the
vertical curvature CV defined using the grade, which is a given industry input that mea-
sures the ratio between the elevation and the longitudinal distance between two points. It
is always preferable to use a tangent track when possible, to avoid more energy consump-
tion, wheel and rail wear, and reduced speeds. This desired tangent-track option, however,
is not always possible because of uneven ground terrain and the need to avoid obstacles.
Consequently, track curvatures and elevations are difficult to avoid.
The grade is defined in North America in terms of the number of feet the track rises per
100 ft of horizontal distance. For example, if a track rises 2 ft over a distance of 200 ft, the
grade is considered to be 1%. In other parts of the world, the grade is measured in terms of
the horizontal distance required to achieve a one-foot rise. In general, grades do not exceed
1%, and grades higher than 2.2% are not common. For every 1% increase in the grade, the
resistance force is approximately four times the force required when a locomotive travels
on a tangent track (McGonigal 2006).
The vertical curvature was defined by Eq. 55 in terms of the derivative of the
vertical-development angle 𝜃 as CV = 𝜃 ′ = d𝜃/ds. Therefore, the data at two points defined
by the arc length values s1 and s2 can be used to define the vertical curvature, which is
assumed constant for a given track segment, as
𝜃 − 𝜃1
CV = 2 (3.61)
s 2 − s1
Motion and Geometry Descriptions 113

where 𝜃 1 and 𝜃 2 are the vertical-development angles at the two points defined by the arc
length values s1 and s2 , respectively. Using the definition of CV in the preceding equation,
the vertical-development angle can be written in a differential form as d𝜃 = CV ds. This
equation, upon integration between the two points, defines the vertical-development angle
as
( )
𝜃 = 𝜃1 + CV s − s1 (3.62)

This definition of the elevation angle 𝜃 can be used with the relationship dS = cos 𝜃ds, if
desired, to write d𝜓 in Eq. 59 in terms of the space curve arc length s.

Bank Angle and Superelevation The bank angle 𝜙 can be obtained using the superel-
evation, which is a given industry input that measures the elevation of one rail with respect
to the other as a result of a bank-angle rotation 𝜙 about the longitudinal axis. The bank
angle is assumed to vary linearly as a function of the projected arc length S as
( )
𝜙 = 𝜙 (S) = 1 − 𝜉S 𝜙1 + 𝜉S 𝜙2
( ) ( )
𝜙2 S − S1 − 𝜙1 S − S2
= ( ) (3.63)
S2 − S1
where 𝜉 S = S/LS , LS = S2 − S1 , and 𝜙1 and 𝜙2 are the bank angles at the two end
nodes of the segments defined, respectively, by the projected arc length coordinates S1
and S2 .
Equations 60, 62, and 63, which define the three Euler angles 𝜓 = 𝜓(S), 𝜃 = 𝜃(s), and
𝜙 = 𝜙(S) as field variables, also define the geometry of the space curve. As previously
mentioned, for convenience, the angles 𝜓 and 𝜙 are written in terms of the projected arc
length S, while the angle 𝜃 is written in terms of the actual arc length s of the space curve.
Recall that the sequence Z, − Y , − X is selected when Euler angles are used to describe the
track geometry in order to have positive definitions for the horizontal curvature, grade, and
superelevation. The curvature is considered positive if it is the result of a positive rotation 𝜓
about the Z axis; the vertical development is considered positive if it is the result of a rota-
tion 𝜃 about the −Y axis; and the superelevation is considered positive if it is the result of a
rotation 𝜙 about the −X axis. This convention is consistent with the measurements made
in practice. These three Euler angles, obtained as field variables, can be used to define the
space curve in a parametric form as described in the remainder of this section. It is impor-
tant, however, to point out that the linearly interpolated bank angle 𝜙 of Eq. 63 is used to
define the orientation of the track (body trajectory) frames that follow the vehicle compo-
nents in railroad vehicle algorithms. This angle does not enter into the definition of the
curve geometry which can be completely defined by two geometric parameters only (Ling
and Shabana, 2020).

Curve Position Coordinates In addition to determining the angles at arbitrary points


on the space curve from the given industry input, the position coordinates of these arbi-
trary points are also required. As previously discussed, the space curve segment is defined
by considering the first column of the transformation matrix of Eq. 36 as the unit vec-
tor tangent to the curve. This allows writing the curve equation in the differential form
[ ]T
dr = cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 sin 𝜃 ds. Using the second equation in Eq. 57, which defines
114 Mathematical Foundation of Railroad Vehicle Systems

the relationship between s and S, dS = cos 𝜃ds, the equation for dr can be written as
⎡cos 𝜓dS⎤
⎢ ⎥
dr = ⎢ sin 𝜓dS ⎥ (3.64)
⎢ ⎥
⎣ sin 𝜃ds ⎦
This equation can be used, upon integration, to define the position coordinates of an arbi-
[ ]T
trary point on the curve r = r1 r2 r3 . This also defines the curve in its parametric form as

⎡ro1 ⎤ ⎡∫So cos 𝜓dS⎤ ⎡r1 (S)⎤


S1

⎢ ⎥ ⎢ S ⎥ ⎢ ⎥
r = ⎢ro2 ⎥ + ⎢ ∫S 1 sin 𝜓dS ⎥ = ⎢r2 (S)⎥ (3.65)
⎢ ⎥ ⎢ s1 ⎢ o ⎥ ⎢ ⎥
⎣ro3 ⎦ ⎣ ∫s sin 𝜃ds ⎥⎦ ⎣ r3 (s) ⎦
o
[ ]T
where ro = ro1 ro2 ro3 is vector r at the start of the segment. This equation shows that,
by using Euler angles as field variables, the coordinates of the points on the space curve
can be determined. Euler angles, as field variables, are assumed to be known from the
horizontal curvature CH , grade 𝜃, and superelevation h, which are given industry inputs
as discussed in Chapter 4. The locations of the points on the space curve and the Euler
angles that define the curve geometry can be determined for the tangent, curve, and spiral
rail sections. These segment types have different geometries because the tangent section
has zero curvature, the curve section has constant curvature, and the spiral section is
designed to have linearly varying horizontal curvature, as discussed in Chapter 4. It is also
important to note that Eq. 65 requires using only two angles because the curve geometry
is completely defined in terms of two geometric invariants only (Ling and Shabana, 2020).

3.12 GEOMETRIC MOTION CONSTRAINTS


Railroad vehicle systems consist of large number of components that have independent
motion relative to each other. Example of these railroad components are the rails, car bod-
ies, frames, and wheelsets. For example, the wheelsets, which are mounted on the frame
using a suspension system, experience large relative rotations with respect to all other vehi-
cle components. Mechanical joints and force elements are used to connect the rail vehicle
components in order to restrict the motion amplitudes and ensure the safe operation of the
rail vehicles. The joints and force elements are also selected in a manner that allows obtain-
ing the desired design and performance characteristics of the system. Using the motion
description introduced at the beginning of this chapter, each vehicle component i with dis-
[ T T ]T
tributed inertia can have the set of generalized coordinates qi = Ri 𝛉i , i = 1, 2, … , nb ,
where Ri is the vector that defines the global position of the body reference point Oi in
the global coordinate system; 𝜽i is the set of orientation parameters that can be selected to
be Euler angles or Euler parameters, as previously discussed in this chapter; and nb is the
total number of bodies in the system. Therefore, the system generalized coordinates can be
written as
[ T T T ]T
q = q1 q2 … qnb
[ T T T T T T ]T
= R1 𝛉1 R2 𝛉2 … Rnb 𝛉nb (3.66)
Motion and Geometry Descriptions 115

If Euler angles are used, the number of system coordinates is 6 × nb , while if Euler param-
eters are used, the number of the system coordinates is 7 × nb .

Degrees of Freedom Due to mechanical joints and specified motion trajectories, the
system coordinates in Eq. 66 are not totally independent. The joints and specified motion
trajectories can be formulated mathematically using a set of nonlinear algebraic constraint
equations that represent relationships between the system coordinates. Because of these
kinematic relationships, dependent coordinates can be determined using other coordinates,
called independent coordinates. If the constraint equations are independent, the number
of dependent coordinates is equal to the number of the algebraic constraint equations nc .
Therefore, if the total number of system coordinates in Eq. 66 is n, the total number of
independent coordinates – also called the system degrees of freedom – is nd = n − nc . Knowl-
edge of the system degrees of freedom is important in understanding a rail vehicle dynamic
behavior and stability. This knowledge is also important in the control of any mechanical
system if under- or over-actuation is to be avoided. For example, to fully control a mechan-
ical system such as a robotic manipulator, the number of actuators and motors must be
equal to the number of system degrees of freedom, to avoid under- or over-actuation. In
railroad vehicle systems, not all of the degrees of freedom are controlled, due to the com-
plexity of the system and also because of the need to provide the vehicle with the freedom
to negotiate unexpected disturbances at a reasonable operating cost. The simple formula
nd = n − nc , which can be used to give an estimate of the number of degrees of freedom nd ,
is called the mobility criterion. This criterion defines the number of degrees of freedom as
the total number of coordinates minus the number of algebraic constraint equations that
relate these coordinates.

Nonlinear Algebraic Constraint Equations Most mechanical joints such as spheri-


cal, revolute, prismatic, cylindrical, and universal joints can be formulated mathematically
using nonlinear algebraic constraint equations in terms of the Cartesian coordinates of
Eq. 66. These algebraic constraint equations can be written in a vector form as
[ ]T
C (q, t) = C1 (q, t) C2 (q, t) … Cnc (q, t) = 𝟎 (3.67)
where nc is the total number of constraint functions. Geometric constraints that can be writ-
ten at the position level, like the ones given in the preceding equation, are called holonomic
constraint equations. An example of these holonomic constraints is the spherical (ball) joint
constraint equations, which eliminate the relative translations and allow only three rela-
tive rotations between the two bodies connected by this joint. Figure 13 shows two bodies i
and j connected by a spherical (ball) joint. The joint attachment points on bodies i and j are,
respectively, denoted as Pi and Pj . The three nonlinear algebraic scalar constraint equations
that describe the spherical joint can be written using the kinematic equations previously
presented in this chapter as
( ) [ ]T j
C qi , qj = C1 C2 C3 = riP − rP
( ) ( )
j
= Ri + Ai uiP − Rj + Aj uP = 𝟎 (3.68)

These three scalar equations eliminate all the relative translational degrees of freedom
between the two bodies.
116 Mathematical Foundation of Railroad Vehicle Systems

Zi Yi

uiP
Xi j
P

R
i Pi
j j
Z uP Z
j
Y X
j j
R Y

Figure 3.13 Spherical joint.

Other joints such as the cylindrical joint can also be systematically formulated in terms of
the Cartesian coordinates in Eq. 66. As shown in Figure 14, the cylindrical joint allows
one relative translation and one relative rotation between the two bodies connected by
the joint along the joint axis. Therefore, this joint eliminates four degrees of freedom: two
translations in the plane perpendicular to the joint axis and two rotations about two axes
perpendicular to the joint axis. Therefore, the cylindrical joint has two degrees of freedom.
If vi = Ai vi and vj = Aj vj are two vectors on bodies i and j, respectively, defined along
the joint axis; and Pi and Pj are two points on the two bodies located on the joint axis,
as shown in Figure 14, the relative rotations between the two bodies about two axes per-
pendicular to the joint axis can be prevented by imposing the mathematical conditions
( )T ( j j ) ( )T ( j j )
v1i ⋅ vj = Ai vi1 A v = 0 and v2i ⋅ vj = Ai vi2 A v = 0, where v1i and v2i are two vec-
tors orthogonal to vi ; vi , v1 , and v2 are constant vectors defined in the coordinate system of
i i

body i; and vj is a constant vector defined with respect to the coordinate system of body j. The
relative translation between the two bodies in a plane perpendicular to the joint axis can be
ij ij ij j
prevented using ( the two scalar
) equations v1i ⋅ rP = 0 and v2i ⋅ rP = 0, where rP = riP − rP =
( i ) j
R + Ai uiP − Rj + Aj uP . One can, therefore, write the four constraint equations of the
cylindrical joint as
( ) [ ]T
C qi , q j = C1 C2 C3 C4
[( T ) ( T ) ( T ) ( T )]T
ij ij
= v1i vj v2i vj v1i rP v2i rP =𝟎 (3.69)

These four constraint equations define the two-degree-of-freedom cylindrical joint.


The revolute joint, also called the pin joint, which allows one relative rotation between two
bodies about the joint axis, can be obtained from the cylindrical joint constraint equations
by adding an additional constraint equation that eliminates the relative translation between
the two bodies. This condition can be defined by requiring that the distance between the
ijT ij
two points Pi and Pj to remain constant: that is, rP rP = c, where c is a constant that defines
the distance between the two points Pi and Pj in the initial configuration. Therefore, the five
Motion and Geometry Descriptions 117

Zi vi
Yi Pi
v2i
v1i

Xi

j
riP v
Z j
P
j
Y j Z
j
X
rP
j
Y
X

Figure 3.14 Cylindrical joint.

constraint equations that define the single-degree-of-freedom revolute joint can be written
as
( ) [ ]T
C qi , qj = C1 C2 C3 C4 C5
[( T ) ( T ) ( T ) ( T ) ( T )]T
ij ij ij ij
= v1i vj v2i vj v1i rP v2i rP rP rP − c =𝟎 (3.70)

An alternate formulation of the revolute joint uses the spherical joint constraints, which
eliminate the relative translations between the two bodies with two additional constraints
that eliminate two relative rotations between the two bodies about two axes perpendicular
to the joint axis. This alternate formulation of the revolute joint can be written as
( ) [ ]T
C qi , qj = CT1−3 C4 C5
[( )T ]T
= riP − rjP v1i vj v2i vj = 𝟎
T T
(3.71)

Equations 70 and 71 lead to the same physical constraints on the motion of the two bodies
connected by the revolute joint.
The constraint equations for the prismatic joint, also called the translational joint, allow
one translational degree of freedom along the joint axis. These constraint equations can also
be obtained by adding to the cylindrical joint constraint equations one algebraic equation
that eliminates the relative rotation between the two bodies connected by the joint. As
shown in Figure 15, one can define the two orthogonal vectors hi = Ai Hi and hj = Aj hj
on bodies i and j, respectively, perpendicular to the joint axis, where Hi and hj are constant
118 Mathematical Foundation of Railroad Vehicle Systems

vi
Yi Pi
Zi v2i
v1i
Xi

j
hi j h
v j
P
riP j
X j
Z
j
rP
Z
j
Y
Y

Figure 3.15 Prismatic (translational joint).

vectors defined, respectively, in the coordinate systems of bodies i and j. In order to pre-
T
vent the relative rotations between the two bodies, one can introduce the condition hi hj =
( i i )T ( j j )
AH A h = 0. Therefore, expanding on the cylindrical joint constraint equations of
Eq. 69, the five constraint equations that define the single-degree-of-freedom prismatic joint
can be written as
( ) [ ]T
C q i , q j = C1 C2 C3 C4 C5
[( T ) ( T ) ( T ) ( T ) ( )]T
ij ij T
= v1i vj v2i vj v1i rP v2i rP hi hj =𝟎 (3.72)

These five equations allow only one translational degree of freedom along the joint axis.
The constraint equations of other joints can be formulated in a similar manner. In addi-
tion to the joint constraints, specified motion trajectory constraints are often used in the
dynamic simulation of railroad vehicle systems. For example, the specified motion trajec-
tory constraints can be used to specify the forward velocity of the vehicle. Other types of
geometric constraints are the constraints imposed on the coordinates in Eq. 66 when Euler
parameters are used. As previously described in this chapter, if the four Euler parameters
∑3 ( )2
are used to define the orientation of a body, one must impose the condition k=0 𝜃ki = 1.

Treatment of the Algebraic Constraint Equations It is clear from the examples


presented in this section for the formulation of some widely used mechanical joints that
the mathematical description of different constraints imposed on the motion of the system
Motion and Geometry Descriptions 119

can be systematically developed in terms of the generalized coordinates used. Kinematic


constraints produce constraint forces that must be distinguished from applied forces
such as gravity, spring, damper, actuator, and aerodynamic forces. Applied forces do not
eliminate degrees of freedom regardless of their magnitude. Constraint forces, on the other
hand, produce reaction or constraint forces that do not do work, as explained in Chapter 6.
As demonstrated by the joint examples presented in this section, the geometric constraint
conditions can always be formulated using a set of nonlinear algebraic equations. When
these algebraic constraint equations are combined with the second-order ordinary differ-
ential equations of the motion of the system, one obtains a system of differential/algebraic
equations that need to be solved simultaneously using numerical methods.
As discussed in Chapter 6, two different MBS approaches can be used for the treatment
of the resulting differential/algebraic equations. In the first approach, referred to in this
book as the augmented formulation, the equations of motion are formulated in terms of
the vector of redundant coordinates q of Eq. 66. Because these coordinates are not inde-
pendent in the presence of the algebraic constraint equations, constraint forces appear in
the equations of motion. In the augmented formulation, the algebraic constraint equations
are combined with the system differential equations of motion to form a large system that
has a sparse matrix structure. This system can be efficiently solved for accelerations and
constraint forces using sparse matrix techniques. In the second approach, referred to in this
book as the embedding technique, the nonlinear algebraic constraint equations are used to
eliminate the dependent variables and write the final form of the differential equations of
motion in terms of the independent accelerations. Because the dependent accelerations are
eliminated and independent accelerations are not related by any algebraic equations, using
this approach allows for systematically eliminating the constraint forces and obtaining a
minimum number of equations equal to the number of the system degrees of freedom. The
obtained system of equations, which often has a dense and highly nonlinear inertia matrix
associated with the independent coordinates, can be solved to determine the independent
accelerations, which can be integrated forward in time to determine the independent coor-
dinates and velocities. The dependent variables (coordinates, velocities, and accelerations)
can be determined from the independent variables using the algebraic constraint equations.
This numerical procedure is consistent with the Lagrange–D’Alembert principle in which a
velocity transformation matrix is formulated to write dependent variables in terms of inde-
pendent variables.

3.13 TRAJECTORY COORDINATES


Using the absolute Cartesian coordinates of Eq. 66 allows for developing general MBS algo-
rithms for railroad vehicle systems. Algorithms based on absolute Cartesian coordinates are
easier to generalize for flexible body dynamics as compared to other algorithms specialized
for specific applications. Nonetheless, specialized railroad vehicle system algorithms have
been developed in the literature based on a set of coordinates called the trajectory coordi-
nates. The trajectory coordinates are also used in specialized railroad software designed to
study longitudinal train dynamics (LTD). LTD software is designed to efficiently solve the
equations of motion of trains consisting of a large number of cars. In some LTD software,
120 Mathematical Foundation of Railroad Vehicle Systems

each car is assumed to have only one degree of freedom that defines the distance traveled
by the car along the track.

X
Y
O

Figure 3.16 Trajectory coordinates.

As discussed in Chapter 6, trajectory coordinates can also be used to formulate the


equations of motion of railroad vehicle systems, and they are related to the absolute
Cartesian coordinates of Eq. 66 by nonlinear relationships. When trajectory coordinates
are used, the configuration of each body like the one shown in Figure 16 can be described
using six coordinates: three translational and three rotational. For each body in the system,
as shown in Figure 16, three coordinate systems are used to define the body configuration.
These three coordinate systems are the following:
● The global coordinate system XYZ is used for all bodies in the system to define global vec-
tors and absolute velocities and accelerations. In this book, the global coordinate system
is assumed to be fixed.
● The trajectory body coordinate system X ti Y ti Z ti follows the motion of body i. This tra-
jectory body coordinate system is uniquely defined using the arc length parameter si ,
as previously described in this chapter. Three Euler rotations 𝜓 ti = 𝜓 ti (si ), 𝜃 ti = 𝜃 ti (si ),
and 𝜙ti = 𝜙ti (si ), respectively, about the Z ti , − Y ti , and −X ti axes, respectively, are used to
define this coordinate system. This sequence leads to the transformation matrix defined
by Eq. 36 and denoted here as Ati . Note that the Euler angles 𝜓 ti = 𝜓 ti (si ), 𝜃 ti = 𝜃 ti (si ),
and 𝜙ti = 𝜙ti (si ) are not generalized coordinates and can be obtained once the arc length
coordinate si is known and the track geometry is given. If the track is fixed, the global posi-
tion of the origin of the trajectory body coordinate system X ti Y ti Z ti is defined by vector
Rti = Rti (si (t)).
● The body coordinate system X i Y i Z i has an origin rigidly attached to a point on the body,
as shown in Figure 16. The orientation of this coordinate system X i Y i Z i is defined with
respect to the trajectory body coordinate system X ti Y ti Z ti using three Euler angles 𝛉ir =
[ ir ]T
𝜓 𝜙ir 𝜃 ir , where 𝜓 ir = 𝜓 ir (t) is the yaw angle resulting from rotation about the Z i
Motion and Geometry Descriptions 121

axis, 𝜙ir = 𝜙ir (t) is the roll angle resulting from rotation about the X i axis, and 𝜃 ir = 𝜃 ir (t)
is the pitch angle resulting from rotation about the Y i axis. Therefore, the transformation
matrix that defines the orientation of the body X i Y i Z i coordinate system with respect to
the trajectory body coordinate system is given by Eq. 12.
[ ( ) ( )
It is important to understand the difference between the angles 𝛉ti = 𝜓 ti si 𝜃 ti si
( )]T
𝜙ti si used to formulate the transformation matrix Ati that defines the orientation of
[ ]T
the trajectory body coordinate system X ti Y ti Z ti and the angles 𝛉ir = 𝜓 ir 𝜙ir 𝜃 ir used to
formulate the transformation matrix Air that defines the orientation of the body coordinate
system X i Y i Z i , as previously discussed in this chapter. The angles used for the coordinate
system X ti Y ti Z ti describe geometry, while the angles used for the coordinate system X i Y i Z i
describe motion. The two transformation matrices Ati and Air are reproduced here for con-
venience (Sanborn et al. 2007; Sinokrot et al. 2008; Shabana et al. 2008):

⎡cos 𝜓 ti cos 𝜃 ti − sin 𝜓 ti cos 𝜙ti + cos 𝜓 ti sin 𝜃 ti sin 𝜙ti − sin 𝜓 ti sin 𝜙ti − cos 𝜓 ti sin 𝜃 ti cos 𝜙ti ⎤
ti⎢ ⎥
A = ⎢ sin 𝜓 ti cos 𝜃 ti cos 𝜓 ti cos 𝜙ti + sin 𝜓 ti sin 𝜃 ti sin 𝜙ti cos 𝜓 ti sin 𝜙ti − sin 𝜓 ti sin 𝜃 ti cos 𝜙ti ⎥
⎢ sin 𝜃 ti − cos 𝜃 ti sin 𝜙ti cos 𝜃 ti cos 𝜙ti ⎥
⎣ ⎦
(3.73)

and
⎡cos 𝜓 ir cos 𝜃 ir − sin 𝜓 ir sin 𝜙ir sin 𝜃 ir − sin 𝜓 ir cos 𝜙ir cos 𝜓 ir sin 𝜃 ir + sin 𝜓 ir sin 𝜙ir cos 𝜃 ir ⎤
⎢ ⎥
Air = ⎢sin 𝜓 ir cos 𝜃 ir + cos 𝜓 ir sin 𝜙ir sin 𝜃 ir cos 𝜓 ir cos 𝜙ir sin 𝜓 ir sin 𝜃 ir − cos 𝜓 ir sin 𝜙ir cos 𝜃 ir ⎥
⎢ − cos 𝜙ir sin 𝜃 ir sin 𝜙ir cos 𝜙ir cos 𝜃 ir ⎥
⎣ ⎦
(3.74)

Two sets of generalized trajectory coordinates, which are different from the absolute
Cartesian coordinates previously used in this chapter, can then defined as follows:

1. The first coordinate is an arc length coordinate si = si (t) that defines the distance traveled
by the body along the track, and the second and third coordinates are the lateral and
vertical displacements yir = yir (t) and zir = zir (t) relative to a trajectory body coordinate
system X ti Y ti Z ti that follows the body, as shown in Figure 16. As previously described in
this chapter, the trajectory body coordinate system can be uniquely defined in terms of
the arc length si , using the definition of the track geometry. The coordinates yir and zir
are relative coordinates and are not absolute even if the track is fixed. In most railroad
vehicle system applications, and under normal operating conditions, the two relative
coordinates yir and zir are normally small.
2. Three relative (not absolute) rotation angles define the body orientation with respect
to the trajectory body coordinate system X ti Y ti Z ti . The first is a yaw angle 𝜓 ir = 𝜓 ir (t)
obtained by a rotation about the Z i axis of the body coordinate system, the second is a
roll angle 𝜙ir = 𝜙ir (t) obtained by a rotation about the X i axis of the body coordinate
system, and the third is a pitch angle 𝜃 ir = 𝜃 ir (t) obtained by a rotation about the Y i axis
of the body coordinate system. In most railroad vehicle system applications, and under
normal operating conditions, the two relative rotations 𝜓 ir and 𝜙ir are normally small,
while the pitch angle 𝜃 ir can be very large.
122 Mathematical Foundation of Railroad Vehicle Systems

The equations of motion of a body can be defined using the translational coordinates si ,
[ ]T
yir , and zirand the relative orientation angles 𝛉ir = 𝜓 ir 𝜙ir 𝜃 ir . Knowing the arc length
parameter si = si (t), the location of the origin and the orientation of the trajectory body
coordinate system X ti Y ti Z ti that follows the motion of the body can be uniquely defined,
as previously discussed in this chapter. Therefore, the vector of the six time-dependent tra-
jectory coordinates of body i is given by (Sanborn et al. 2007; Sinokrot et al. 2008; Shabana
et al. 2008)
[ ]T
pi (t) = si yir zir 𝜓 ir 𝜙ir 𝜃 ir (3.75)
If the rails are assumed fixed, the global position vector of the body reference point (origin
of the coordinate system X i Y i Z i ) can be written as
( ) ( )
Ri = Rti si + Ati si uir (3.76)
[ ] T
where vector uir = 0 yir zir is the position vector of the body reference point with respect
to the origin of the trajectory body coordinate system. As previously mentioned, yir and
zir are, respectively, the lateral and normal components of the position vector of the body
reference point with respect to the origin of the trajectory body coordinate system. In writing
Eq. 76, it is assumed that the longitudinal component in vector uir is zero. Equation 76
can be used to define the global position of an arbitrary point on the body, as previously
discussed in this chapter, as ri = Ri + Ai ui , where Ai = Ati Air .
It is clear that using the equation ri = Ri + Ai ui with Eq. 76 can make the formulation
of the velocity and acceleration equations as well as the equations of motion cumbersome.
It is easier, for a given arbitrary body i, to develop the velocity transformation that relates
[ T T ]T
the derivatives of the absolute Cartesian coordinates qi = Ri 𝛉i to the derivatives of
[ i ir ir ]T
the trajectory coordinates p = s y z 𝜓 𝜙 𝜃
i ir ir ir and use this velocity transfor-
mation with the relatively simple Newton–Euler equations obtained in Chapter 6 to derive
the equations of motion in terms of trajectory coordinates. To develop this velocity trans-
formation, Eq. 76 is differentiated with respect to time to yield
Ṙ i = Ṙ ti + Ȧ ti uir + Ati u̇ ir = Li p̈ i (3.77)
[(( ) ( ) ) ]
where Li = 𝜕Rti ∕𝜕si + 𝜕Ati ∕𝜕si uir a2ti a3ti 𝟎 is a 3 × 6 matrix in which a2ti and a3ti
are the second and third columns of the transformation matrix Ati of Eq. 73, and 0 is a 3 × 3
null matrix resulting from the fact that the position vector of Eq. 76 does not depend on the
[ ]T
angles 𝛉ir = 𝜓 ir 𝜙ir 𝜃 ir . By differentiating Eq. 77 with respect to time, one obtains
̈ i = Li p̈ i + 𝛄iR
R (3.78)
where 𝛄iR is a vector that includes terms that are quadratic in the velocities.
Because angular velocities can be added and treated as vectors, one can write the absolute
angular velocity vector of body i as 𝛚i = 𝛚ti + 𝛚ir , which is the sum of the absolute angu-
lar velocity 𝛚ti = Gti 𝛉̇ ti of the trajectory body coordinate system and 𝛚ir = Ati Gir 𝛉̇ ir , which
defines the angular velocity of the body with respect to the trajectory body coordinate sys-
tem. Therefore, the absolute angular velocity 𝛚i of body i can be written in terms of the time
derivatives of the trajectory coordinates as
𝛚i = Gti 𝛉̇ ti + Ati Gir 𝛉̇ ir = Hi p̈ i (3.79)
Motion and Geometry Descriptions 123

where matrices Gti and Gir are, respectively, defined as

⎡0 sin 𝜓 ti − cos 𝜓 ti cos 𝜃 ti ⎤ ⎡0 cos 𝜓 ir − sin 𝜓 ir cos 𝜙ir ⎤


⎢ ⎥ ⎢ ⎥
Gti = ⎢0 − cos 𝜓 ti − sin 𝜓 ti cos 𝜃 ti ⎥ , Gir = ⎢0 sin 𝜓 ir cos 𝜓 ir cos 𝜙ir ⎥
⎢ ⎥ ⎢ ⎥
⎣1 0 − sin 𝜃 ti ⎦ ⎣1 0 sin 𝜙ir ⎦
(3.80)
[( ( )) ( )]
and Hi = Gti 𝜕𝛉ti ∕𝜕si 𝟎 𝟎 Ati Gir is a 3 × 6 matrix in which 0 is a three-
dimensional zero vector. Differentiating Eq. 79 with respect to time, the absolute angular
acceleration vector 𝜶 i of body i can be written in terms of the second time derivatives of
the trajectory coordinates as
𝛂i = Hi p̈ i + 𝛄i𝛼 (3.81)
where 𝛄i𝛼 is a vector that absorbs terms that are quadratic in the velocities (Sanborn et al.
2007; Sinokrot et al. 2008; Shabana et al. 2008). Equation 81 can also be used to write the
absolute angular acceleration vector defined in the body coordinate system X i Y i Z i as
T
𝛂i = Ai 𝛂i = Hi p̈ i + 𝛄i𝛼 (3.82)
T T
iT
where Hi = A Hi = Air Ati Hi and 𝛄i𝛼 iT
= A 𝛄i𝛼 . Combining Eqs. 78 and 82, one obtains
(Sanborn et al. 2007; Sinokrot et al. 2008)
[ ] [ ] [ ]
R̈i Li 𝛄iR
= p̈ + i
i
(3.83)
𝛂i Hi 𝛄𝛼
As demonstrated in Chapter 6, the mass matrix associated with the absolute Cartesian
[ iT iT ]T
accelerations R ̈ 𝛂 is constant in the Newton–Euler formulation of the equations of
motion; therefore, it is preferred to use the angular acceleration vector 𝛂i instead of 𝜶 i .
Equation 83 can be used with Newton–Euler equations to obtain the equations of motion in
terms of the trajectory coordinates, as discussed in Chapter 6. These equations are obtained
without using any small angle assumptions; therefore, they include all the nonlinear terms
that can be significant in some railroad-vehicle motion scenarios.
125

Chapter 4

RAILROAD GEOMETRY

Wheel and rail geometry has a significant influence on the dynamic behavior of railroad
vehicle systems. Consequently, an accurate description of the geometry is necessary in order
to correctly predict wheel/rail contact forces. These forces have a significant effect not only
on vehicle dynamics and stability but also on the integrity of the track structure. The geom-
etry, which enters into formulating the normal and tangential wheel/rail contact forces,
can be described using the theories of curves and surfaces introduced in Chapter 2. For
example, to determine the dimensions of the wheel/rail contact area, the principal curva-
tures and principal directions of the wheel and rail surfaces in the contact region must be
evaluated. While the wheel geometry can be described using a surface of revolution, the
rail geometry can be defined by extruding the rail profile in the rail longitudinal direction.
More complex surface geometries can be defined using numerical approximation methods
to capture details that cannot be captured using analytical techniques that are more suited
for simple or idealized geometries.
The geometry of an unworn wheel can be described as a surface of revolution by rotating
the profile curve about the wheel axis. While the profile can assume any shape, conical
wheels are often used to improve vehicle stability and avoid derailments. Different profiles
with different conicity values are used depending on the type of vehicle, speed of operation,
and loading conditions. The functions that define the profiles are not simple straight-line
functions, and in most practical applications, a numerical description of the profile function
is required for accurate computer modeling and virtual prototyping.
Developing an accurate description of track geometry is one of the basic steps in formulat-
ing railroad-vehicle nonlinear dynamic equations of motion and the numerical solution of
these equations. For the most part, tracks are constructed using three segment types with
different geometries: tangent, curve, and spiral segments. A tangent segment is a straight
section of track with zero curvature, which corresponds to a radius of curvature equal to
infinity. A curve segment is a circular section of track that has a constant radius of curvature
and constant curvature. To connect tangent and curve segments or two curves with differ-
ent curvature values, a spiral segment is used. When a spiral segment is used to connect
a tangent section and a curve section, for example, the geometry of the spiral segment is
designed such that the spiral has zero curvature at the end connected to the tangent seg-
ment and the value of the curve curvature at the end connected to the curve segment. This
spiral design allows for smoothly varying the curvature and ensures smooth operation of
rail vehicles during the transition between tangent and curve sections and vice versa.
Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,
First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
126 Mathematical Foundation of Railroad Vehicle Systems

When a railroad vehicle negotiates a curve or spiral section, the forces exerted on the
vehicle can be significantly different from the forces that arise when the vehicle negotiates
a straight segment. During curve and spiral negotiations, centrifugal forces must be taken
into account in order to ensure safe operation of the rail vehicle. To avoid derailments, the
magnitude of the centrifugal forces is used to put a limit on the vehicle speed; this limit
is referred to as the balance speed. For a curve that has a constant radius of curvature, the
variations in the centrifugal forces acting on the vehicle are not as significant as for a spiral,
which has variable curvature. When constructing a track, the geometry of spiral sections is
often designed to have a linearly varying curvature.
This chapter discusses the geometries of wheel and rail surfaces, which are fundamental
in defining the kinematic and dynamic equations of railroad vehicles. The wheel and rail
surface equations are defined in terms of two independent surface parameters, as explained
in Chapter 2. These surface parameters, which define the surface equations in their para-
metric form, can be used to determine the locations of arbitrary points on wheel and rail
surfaces. The parametric surface equations are necessary for developing wheel/rail contact
conditions, which are used to determine online wheel/rail contact points. This chapter dis-
cusses two approaches for describing rail geometry: the semi-analytical approach and the
absolute nodal coordinate formulation (ANCF) interpolation approach. The semi-analytical
approach has two main disadvantages. The first disadvantage is the need to evaluate the
derivatives of angles with respect to the rail longitudinal surface parameter in order to
determine the tangent, normal, and curvature vectors. The second disadvantage is the low
order of interpolation used to determine the position coordinates of arbitrary points on
the rail with specific output from the track preprocessor, which is designed to generate
the data required at discrete nodal points for use in nonlinear dynamic simulations of rail-
road vehicles. The ANCF interpolation approach is preferred because it does not have these
disadvantages: it does not require differentiation of angles, and it allows for higher-order
interpolation for the position coordinates of arbitrary points on the rail space curve.

4.1 WHEEL SURFACE GEOMETRY


In the case of unworn wheels, the wheel surface can be considered a surface of revolu-
tion generated by rotating the profile curve about the wheel axis, as shown in Figure 1.

w
Profile curve Z
yw
s w
X

θw
s w
Y

Figure 4.1 Wheel profile curve.


Railroad Geometry 127

wp
Z

θw
s
wp
X

Figure 4.2 Wheel profile frame.

As explained in Chapter 2, the locations of points on a surface can be expressed in terms of


two surface parameters. In this chapter, the wheel surface is assumed to have the parame-
terization sw1 = yws and sw2 = 𝜃sw , as shown in Figure 1. To develop a mathematical definition
of the wheel surface, a profile frame X wp Y wp Z wp is first introduced, as shown in Figure 2, for
the convenience of defining the profile curve. As shown in the figure, the angular surface
parameter 𝜃sw is measured from the Z wp axis. The position of the origin of the profile frame
with respect to the wheel coordinate system X w Y w Z w is defined by the Cartesian coordi-
wp wp wp
nates xo , yo , and zo . The Y w axis is assumed to coincide with the wheel axis of rotation.
In the case of a rigid wheelset, the wheelset coordinate system X w Y w Z w is located at the
center of mass of the wheelset. The negative sign is used in the third element of the vec-
tor of Eq. 1 to associate the contact point in the initial configuration with the zero value of
the radial surface parameter sw2 . A profile function gwp is used to define the wheel profile
curve in the wheel profile frame. In the case of unworn wheels, profile function gwp does
not depend on the wheel angular surface parameter sw2 = 𝜃sw : in this special case, one has
gwp = gwp (sw1 ). In this case of unworn wheels, the coordinates of a point on the wheel surface
that comes into contact with the rail can be defined mathematically in the selected wheel
coordinate system X w Y w Z w in terms of surface parameters sw1 and sw2 as
⎡xowp + gwp (sw1 ) sin sw2 ⎤
⎢ wp ⎥
uw (sw1 , sw2 ) = Rwp + uwp =⎢ yo + sw1 ⎥ (4.1)
⎢ wp wp w w⎥
⎣zo − g (s1 ) cos s2 ⎦
wp wp wp
where Rwp = [xo yo zo ]T is the vector that defines the origin of the wheel profile frame
X wp Y wp Z wp with respect to the coordinate system X w Y w Z w of the wheel or wheelset, and
[ ]T
uwp = gwp (sw1 ) sin sw2 sw1 −gwp (sw1 ) cos sw2 is the vector that defines the location of the point
wp wp wp
in the profile frame. For a single wheel, the coordinates xo , yo , and zo define the position
of the origin of the profile frame X Y Z in the wheel coordinate system X w Y w Z w . For
wp wp wp

a rigid wheelset that has two wheels rigidly connected by an axle, the coordinate systems
of the rigidly connected right and left wheels can be assumed the same and have origins
wp wp wp
located at the axle center point, as shown in Figure 3; in this case, xo = zo = 0, and yo
can be selected to be half the distance between the centers of the two wheels. Using the
wp wp wp
coordinates xo , yo , and zo of the origin of the wheel profile frame allows the system-
atic generalization of the geometric description presented in this section to the case of a
deformable wheel axle or independent non-rigidly connected wheels.
128 Mathematical Foundation of Railroad Vehicle Systems

w
Z
w
Y

w
X

Figure 4.3 Wheelset coordinate system.

Profile Geometry and Wheel Flanges Wheel profiles are designed to have conical
shapes to improve stability and steering. In practice, the conical profile is not an exact
straight line, but a curve generated by combinations of arcs. Flanges are added to wheels
to restrict lateral displacements and prevent derailments. Table 1 shows details of wheel
sections, profile terminologies, and an example of wheel profiles that have been used in
practice. The profile can be measured using a device known as a MiniProf, which generates
y − z numerical data that can be used with a cubic spline function representation to define
the wheel profile function numerically: gwp = gwp (sw1 ).
Different wheel designs with different profile conicity are used in practice (Cummings
2018). An example is the AAR 1:20 profile, which was widely used before 1990 and which
represented a significant improvement over the cylindrical profile. This wheel design has
a flange angle of 70∘ , flange fillet radius of 0.75 in., and tread taper of 1:20. The AAR 1:20
wheel design was altered to obtain the AAR-1 design that has been used since 1990. The
AAR-1 wheel was initially introduced in the 1980s based on measurements of worn wheel
profiles. It demonstrated better performance during curve negotiations and has improved
rolling resistance at the expense of lower critical hunting speed: 49 mph compared to
70 mph for the AAR 1:20 wheel. Revised versions, AAR-1A and AAR-1B, were introduced
later to improve stability characteristics and wear resistance. The AAR-1 wheel has a 75∘
flange angle; and because its design is based on worn-wheel measurements, it has multiple
flange fillet radii. Like the AAR 1:20, AAR-1 designs have a 1:20 tread taper (Cummings
2018).
Another design based on worn-wheel geometry, which was introduced later, is the AAR-
2A design, which has a 75∘ flange angle, a 1:20 tread taper, and multiple flange fillet radii.
The AAR-2A design, which was originally called TTCI-1A and later SRI-1A, offers improved
curving performance. As reported by Cummings (2018), the final version of this design
reduces the flange thickness by 1/8 in. and is available with a wide flange (1.25 in.) or nar-
row flange (1.15). Several other profile designs were also presented in a report produced by
the American Public Transportation Association (APTA Press Task Force 2007).
In general, profile dimensions are measured relative to a point on the profile called
the tread datum position, shown in Table 1. The tread datum position defines a circle
on the tread or a tread cross section, which is 70 mm from the flange back face. For
example, the flange height, shown in Table 1, is defined as the difference between the
wheel radius at the flange tip and the wheel radius at the tread datum position. In addition
to using the tread datum position to define the dimensions of the profile, profile gauges
Railroad Geometry 129

Table 4.1 Wheel geometry.

Wheel structure ● Hub, the center portion of the wheel, is


Hub used to mount the wheel on the axle.
● Disc is the section of the wheel that
Tire Disc connects the hub to the tire.
● Tire is the section of the wheel that has
the wheel profile, which contacts the rail.

Tire details Sections


St is the tread.
Pft Sfr is the flange root.
Sft
Sft is the flange toe.
Sfbb Sfbb is the flange back blend.
dfh Sfr
Stc αf Sfb is the flange back.
St
Stc is the tread chamfer.
Ptd Sww is the wheel web.
dtt Sfb
dtw Points
Pft is the flange tip.
Ptd is the tread datum.
dft Dimensions
dfh is the flange height.
dtw is the tire/rim width.
Sww dtt is the throat thickness.
dft is the flange-back/tread-datum distance
(normally 70 mm).
Angles
𝛼 f is the flange angle.

Example of basic dimensions (Riftek Sensors and Instruments 2008)

β5 β1
β4
β3

β2
β6
β7
β8
(Continued)
130 Mathematical Foundation of Railroad Vehicle Systems

Table 4.1 (Continued)

Rolling stock
Parameter Locomotive coach Description

𝛽1 2 mm 5 mm Used to calculate flange slope


𝛽2 70 mm 70 mm Defines wheel rolling circle position
𝛽3 13 mm 18 mm Used to calculate flange thickness
𝛽4 30 mm 28 mm Used to calculate tire roll wear = height
of reference profile flange
𝛽5 — 60∘ Reference profile slope
𝛽6 70 70 Used to calculate rolling surface section
slope
𝛽7 105 105 Used to calculate rolling surface section
slope
𝛽8 140 140 Used to calculate tire/rim width

Profile example (APTA Press Task Force 2007)

APTA 120 wheel profile (based on former AAR S-621-79, 1:20 taper; dimensions in inches)

5 12 ±1/8
5 +1/16
1 32 1 11
–0 16
5 1 14 31
8 32
R 58 R 58
5
R 11
16 ±1/16
8
Taper 1:20
+1/16
1 –0
R 29
32 Base line
3
R1 7 64 Tape line
8 Gage 1
16 31
R2 32
point
R 58

E +Y

D F
C G +X
B H
I
A

Gage
point
J
Railroad Geometry 131

Table 4.1 (Continued)

Node Coordinates Segment Details


Point X Y Segement Radius X Y
Line – Center – Center
A –1.1563 –0.6250 A–B Line 90°
B –1.1562 –0.4583 B–C 0.9063 –0.2500 –0.4583
C –1.1013 –0.1476 C–D 2.9688 1.6875 –1.1654
D –1.0438 –0.0019 D–E 0.6250 –0.4688 –0.2469
E –0.5313 0.3750 E–F 0.6250 –0.5781 –0.2482
F –0.0270 0.0465 F–G 1.8750 –1.6805 –0.8376
Gage Point 0.0000 0.0000
G 0.0751 –0.1790 G–H 0.6875 0.7188 0.0625
H 0.7188 –0.6250 H–I Line 1:20
I 3.7500 –0.7766 I–J 0.6250 3.7188 –1.4008
J 4.3438 –1.4008 Beyond J Line 90°

also use this position as a reference point. Another important dimension that is used to
ensure wheel strength is throat thickness. In determining this dimension, which depends
on the size of the wheel bearings, wheel/rail forces and axle loads must be considered.
The throat thickness is the distance between the root of the tread profile radius at the
flange root and the inner side of the flange-side wheel rim, measured at the narrowest
point (British Railway Board 1996). The minimum value for the throat thickness, which
depends on the type of wheel and its use, is in the range 30–50 mm. While more examples
of profiles and information on the wheel geometry and various wheel designs can be found
in the literature, Table 1 shows the complexity of the wheel profile geometry, which is
constructed using circular arcs. Because developing an analytical representation of the
wheel profile geometry can be difficult, a numerical description based on spline functions
is often used in the virtual prototyping of railroad vehicle systems.

Worn-Wheel Geometry For worn wheels, the wheel surface geometry can be deter-
mined from measurements. Point clouds can be created and used to define the profile func-
tion gwp numerically at tabulated data points for different values of the surface parameters
sw1 = yws and sw2 = 𝜃sw . With the advancement in scanning and imaging techniques and com-
puter technology, accurate point-cloud data can be obtained and used to define the geome-
try of worn wheels. The profile function depends on the two parameters sw1 and sw2 – that is,
gwp = gwp (sw1 , sw2 ) – and Eq. 1 can be replaced by the following more general equation:
⎡xowp + gwp (sw1 , sw2 ) sin sw2 ⎤
⎢ wp ⎥
uw (sw1 , sw2 ) = Rwp + uwp = ⎢ yo + sw1 ⎥ (4.2)
⎢ wp w⎥
⎣zo − g (s1 , s2 ) cos s2 ⎦
wp w w

[ ]T
In this equation, Rwp = xo yo zowp is the same as previously defined and is the vector
wp wp

that defines the origin of the wheel profile frame X wp Y wp Z wp with respect to the coordi-
[
nate system X w Y w Z w of the wheel or wheelset, and uwp = gwp (sw1 , sw2 ) sin sw2 sw1 −gwp (sw1 , sw2 )
]T
cos sw2 is the vector that defines the location of an arbitrary point on the profile frame
that may come into contact with the rail surface. Clearly, the case of worn wheels cannot
be considered the simple case of a surface of revolution as can be assumed for unworn
132 Mathematical Foundation of Railroad Vehicle Systems

wheels. When numerical methods are used to describe the profile function gwp (sw1 , sw2 ),
the value of this function at an arbitrary point on the wheel surface can be determined
using interpolations based on tabulated data points obtained from measurements and/or
scanning and imaging techniques.
Geometric tolerances are defined for new wheel profiles to estimate the level of wear of
wheels in service (British Railway Board, 1996). Profiles that do not meet these profile tol-
erances can lead to vehicle instability and derailments and can also result in significantly
increased wear. Specific measures are used in practice to define the profile wear, includ-
ing the tread run-out, which defines the change in the radius at a tread position in one
complete revolution about the wheel axle. The wear allowance is defined by measuring the
difference between the as-machined and fully worn flange height (British Railway Board
1996). In general, wear allowance is defined as the reduction allowed in the wheel radius
after profiling. While wear allowance can be in the range of 6.5 mm, tighter limits may be
adopted depending on wheel usage, loads, and operating speeds.

4.2 WHEEL CURVATURES AND GLOBAL VECTORS


It is clear from Eqs. 1 and 2 that only one function gwp is needed to define the coordi-
nates of arbitrary points on the wheel surface that come into contact with the rail. For
unworn wheels, this function depends only on the lateral surface parameter sw1 = yws : that
is, gwp = gwp (sw1 ). For worn wheels, gwp depends on the lateral surface parameter sw1 = yws
as well as the angular surface parameter sw2 = 𝜃sw : that is, gwp = gwp (sw1 , sw2 ). As previously
mentioned, the profile function gwp can be defined analytically or numerically using inter-
polation methods, including spline functions. Such a numerical approach offers generality
and allows for using measured data.
To formulate the contact conditions used to determine the locations of contact points on
the wheel, tangent and normal vectors must be evaluated. These tangent and normal vectors
are also required in order to define normal and tangential contact forces. By knowing the
profile function gwp , the tangent and normal vectors at an arbitrary point on the wheel sur-
w w
face can be determined, as discussed in Chapter 2. The two tangent vectors t1 and t2 define
the tangent plane and also define the normal vector nw in the wheel or wheelset coordi-
nate system. These tangent and normal vectors are defined, respectively, in the wheel or
wheelset coordinate system X w Y w Z w as
w 𝜕uw w 𝜕uw w w
t1 = w , t2 = w , nw = t1 × t2 (4.3)
𝜕s1 𝜕s2
where vector uw of Eq. 1 or 2 defines the location of an arbitrary point on the wheel surface
with respect to the wheel or wheelset coordinate system. If the unit normal vector must be
⁀ w w w w
computed, it is defined as n w = (t1 × t2 )∕|t1 × t2 |.

Higher Derivatives As discussed in later chapters of this book, it may be necessary to


compute higher derivatives with respect to the wheel surface parameters, particularly when
the contact conditions are solved numerically using the iterative Newton–Raphson proce-
dure. In this case, one needs to evaluate the curvature vectors as well as the derivative of
the normal vector. The curvature vectors at an arbitrary point on the wheel surface can be
written as
Railroad Geometry 133

w
𝜕 2 uw ∕𝜕(sw1 )2 = 𝜕t1 ∕𝜕sw1 ⎫
w ⎪
𝜕 2 uw ∕𝜕(sw2 )2 = 𝜕t2 ∕𝜕sw2 ⎬ (4.4)
w w w⎪
𝜕 2 uw ∕𝜕sw1 𝜕sw2 = 𝜕t1 ∕𝜕sw2
= 𝜕t2 ∕𝜕s1 ⎭
The derivatives of the normal vector with respect to the surface parameters can be deter-
mined using the derivatives of the two tangent vectors as
w w w w }
𝜕nw ∕𝜕sw1 = (𝜕t1 ∕𝜕sw1 ) × t2 + t1 × (𝜕t2 ∕𝜕sw1 )
w w w w (4.5)
𝜕nw ∕𝜕sw2 = (𝜕t1 ∕𝜕sw2 ) × t2 + t1 × (𝜕t2 ∕𝜕sw2 )
It is clear from the preceding two equations that the derivatives of the normal vector are
⁀ w w w
functions of the curvature vectors. When using the unit normal vector n w = (t1 × t2 )∕|t1 ×
w
t2 |, the derivatives can also be evaluated in closed form. For example, if a⁀ is a unit vector,
T T
one has the identity 𝜕a⁀∕𝜕a = (I − a⁀a⁀ )∕|a|, where a⁀a⁀ is the outer or dyadic product. It
T
follows that 𝜕a⁀∕𝜕s = (𝜕a⁀∕𝜕a)(𝜕a∕𝜕s): that is, 𝜕a⁀∕𝜕s = ((I − a⁀a⁀ )∕|a|)(𝜕a∕𝜕s) for any param-
eter s. Using this identity, one can write the derivative of the unit normal to the wheel
surface as
⁀ ⁀ ⁀ T
𝜕 n w ∕𝜕swk = ((I − n wn w )∕|nw |)(𝜕 nw ∕𝜕swk ), k = 1, 2 (4.6)
In this equation, 𝜕nw ∕𝜕swk , k
= 1, 2, can be determined using Eq. 5. Therefore, the deriva-
tives of the unit normal to the wheel surface can be systematically evaluated using the tan-
gent vectors and their derivatives. The numerical approach based on interpolations allows
using the equations developed in this section, as discussed in later chapters.
Table 2 shows a summary of the wheel surface position and derivative equations in the
cases of unworn and worn wheels where gwp = gwp (sw1 ) and gwp = gwp (sw1 , sw2 ), respectively.
The expressions presented in this table are used in the numerical implementation of the
formulations presented in this book. These expressions show how position and derivative
functions vary as a function of the profile function gwp and its derivatives.

Wheel Global Vectors Using the equation for the position vector of an arbitrary point
with respect to the wheel coordinate system X w Y w Z w , the position vector rw of arbitrary
points can be defined in the global coordinate system XYZ, as explained in the preceding
chapter, as
rw = Rw + Aw uw = Rw + Aw (Rwp + uwp ) (4.7)
where Rw = Rw (t) is the global position vector of the origin of the wheel coordinate sys-
tem X w Y w Z w , Aw is the transformation matrix that defines the orientation of the wheel
coordinate system, and t is time. Because Rw and Aw do not depend on the wheel surface
parameters, one can show that the tangent and normal vectors defined in the global coor-
dinate system XYZ are given by
( w)
𝜕rw 𝜕u w ⎫
w
t1 = w = A w
= Aw t1 ,
𝜕s1 𝜕sw1 ⎪
( w) ⎪
𝜕r w 𝜕u w w ⎬ (4.8)
tw2 = w = Aw = A t ,
𝜕s2 𝜕sw2 2 ⎪
w w w w ⎪
nw = tw1 × tw2 = (Aw t1 ) × (Aw t2 ) = Aw (t1 × t2 ) = Aw nw ⎭
134 Mathematical Foundation of Railroad Vehicle Systems

Table 4.2 Summary of the position and derivative equations of the wheel surface.

gwp = gwp (sw1 ) gwp = gwp (sw1 , sw2 )

⎡xwp + gwp sin sw ⎤ ⎡xwp + gwp (sw , sw ) sin sw ⎤


⎢ o 2⎥ ⎢ o 1 2 2⎥
uw (sw1 , sw2 ) = ⎢ wp
yo + sw1 ⎥ uw (sw1 , sw2 ) = ⎢ wp
yo + sw1 ⎥
⎢ ⎥ ⎢ ⎥
⎢zwp − gwp cos sw ⎥ ⎢zwp − gwp (sw , sw ) cos sw ⎥
⎣ o 2⎦ ⎣ o 1 2 2⎦

⎡ (𝜕gwp ∕𝜕sw ) sin sw ⎤ ⎡ (𝜕gwp ∕𝜕sw ) sin sw ⎤


⎢ 1 2 ⎥ ⎢ 1 2 ⎥
w 𝜕uw ⎢ ⎥ w 𝜕uw ⎢ ⎥
t1 = w = 1 t1 = w = 1
𝜕s1 ⎢ ⎥ 𝜕s1 ⎢ ⎥
⎢−(𝜕g ∕𝜕s ) cos s ⎥
wp w w ⎢−(𝜕g ∕𝜕s ) cos s ⎥
wp w w
⎣ 1 2⎦ ⎣ 1 2⎦

⎡gwp cos sw ⎤ ⎡ (𝜕gwp ∕𝜕sw ) sin sw + gwp cos sw ⎤


⎢ 2⎥ ⎢ 2 2 2 ⎥
w 𝜕uw ⎢ ⎥ w 𝜕uw ⎢ ⎥
t2 = w = 0 t2 = w = 0
𝜕s2 ⎢ ⎥ 𝜕s2 ⎢ ⎥
⎢ gwp sin sw ⎥ ⎢−(𝜕gwp ∕𝜕sw ) cos sw + gwp sin sw ⎥
⎣ 2⎦ ⎣ 2 2 2⎦

w w
nw = t1 × t2
⎡ gwp sin sw ⎤
⎢ 2 ⎥ ⎡−(𝜕gwp ∕𝜕sw ) cos sw + gwp sin sw ⎤
n = t1 × t2 = ⎢−gwp (𝜕gwp ∕𝜕sw1 )⎥
w w w
⎢ 2 2 2 ⎥
⎢ ⎥ = ⎢−gwp (𝜕gwp ∕𝜕sw1 ) ⎥
⎢ −gwp cos sw ⎥ ⎢ ⎥
⎣ 2 ⎦ ⎢−[(𝜕gwp ∕𝜕sw ) sin sw + gwp cos sw ]⎥
⎣ 2 2 2 ⎦

⎡ −(𝜕gwp ∕𝜕sw ) cos sw + gwp sin sw ⎤


⎢ 2 2 2 ⎥
⁀w nw ⁀w nw
n = w = n = w =⎢ −gwp (𝜕gwp ∕𝜕sw1 ) ⎥,
|n | |n | ⎢ ⎥
⎡ gwp sin sw ⎤ ⎢−[(𝜕gwp ∕𝜕sw ) sin sw + gwp cos sw ]⎥
⎣ 2 2 2 ⎦
⎢ 2 ⎥
1 ⎢ wp wp w ⎥
−g (𝜕g ∕𝜕s1 ) ,
|nw | ⎢ ⎥
⎢ −gwp cos sw ⎥
⎣ 2 ⎦
√ √
|nw | = gwp 1 + (𝜕gwp ∕𝜕sw1 )2 |nw | = (gwp )2 + (𝜕gwp ∕𝜕sw2 )2 + (gwp (𝜕gwp ∕𝜕sw1 ))2

w w
𝜕t1 ∕𝜕sw1 = 𝜕 2 uw ∕𝜕(sw1 )2 𝜕t1 ∕𝜕sw1 = 𝜕 2 uw ∕𝜕(sw1 )2
⎡ (𝜕 2 gwp ∕𝜕(sw )2 ) sin sw ⎤ ⎡ (𝜕 2 gwp ∕𝜕(sw )2 ) sin sw ⎤
⎢ 1 2 ⎥ ⎢ 1 2 ⎥
=⎢ 0 ⎥ =⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢ −(𝜕 2 gwp ∕𝜕(sw )2 ) cos sw ⎥ ⎢ −(𝜕 2 gwp ∕𝜕(sw )2 ) cos sw ⎥
⎣ 1 2 ⎦ ⎣ 1 2 ⎦

⎡−gwp sin sw ⎤ w
𝜕t2 ∕𝜕sw2 = 𝜕 2 uw ∕𝜕(sw2 )2
⎢ 2⎥
𝜕t2 ∕𝜕s2 = 𝜕 u ∕𝜕(s2 ) = ⎢
w w 2 w w 2
0 ⎥ ⎡ (𝜕 2 gwp ∕𝜕(sw )2 ) sin sw + 2(𝜕gwp ∕𝜕sw ) cos sw − gwp sin sw ⎤
⎢ ⎥ ⎢ 2 2 2 2 2 ⎥
⎢ gwp cos sw ⎥
⎣ 2 ⎦
=⎢ 0 ⎥
⎢ ⎥
⎢ −(𝜕 2 gwp ∕𝜕(sw )2 ) cos sw + 2(𝜕gwp ∕𝜕sw ) sin sw + gwp cos sw ⎥
⎣ 2 2 2 2 2 ⎦
w w w w
𝜕t1 ∕𝜕sw2 = 𝜕t2 ∕𝜕sw1 = 𝜕 2 uw ∕𝜕sw1 𝜕sw2 𝜕t1 ∕𝜕sw2 = 𝜕t2 ∕𝜕sw1 = 𝜕 2 uw ∕𝜕sw1 𝜕sw2
⎡ (𝜕gwp ∕𝜕sw ) cos sw ⎤ ⎡ (𝜕gwp ∕𝜕sw ) cos sw + (𝜕 2 gwp ∕𝜕sw 𝜕sw ) sin sw ⎤
⎢ 1 2 ⎥ ⎢ 1 2 1 2 2 ⎥
=⎢ 0 ⎥ =⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢ (𝜕g ∕𝜕s ) sin s ⎥
wp w w ⎢ (𝜕gwp ∕𝜕sw ) sin sw − (𝜕 2 gwp ∕𝜕sw 𝜕sw ) cos sw ⎥
⎣ 1 2 ⎦ ⎣ 1 2 1 2 2 ⎦
Railroad Geometry 135

Table 4.2 (Continued)


w w w w w w w w
𝜕nw ∕𝜕sw1 = (𝜕t1 ∕𝜕sw1 ) × t2 + t1 × (𝜕t2 ∕𝜕sw1 ) 𝜕nw ∕𝜕sw1 = (𝜕t1 ∕𝜕sw1 ) × t2 + t1 × (𝜕t2 ∕𝜕sw1 )
⎡ (𝜕gwp ∕𝜕sw1 ) sin sw2 ⎤ ⎡ −(𝜕 2 gwp ∕𝜕sw 𝜕sw ) cos sw + (𝜕gwp ∕𝜕sw ) sin sw ⎤
⎢ ⎥ ⎢ 1 2 2 1 2 ⎥
= ⎢ −(𝜕gwp ∕𝜕sw1 )2 − gwp (𝜕 2 gwp ∕𝜕(sw1 )2 ) ⎥ =⎢ −(𝜕gwp ∕𝜕sw1 )2 ⎥
⎢ ⎥ ⎢ ⎥
⎢ −(𝜕gwp ∕𝜕sw1 ) cos sw2 ⎥ ⎢ −(𝜕 2 gwp ∕𝜕sw 𝜕sw ) sin sw − (𝜕gwp ∕𝜕sw ) cos sw ⎥
⎣ ⎦ ⎣ 1 2 2 1 2 ⎦
w w w w w w w w
𝜕nw ∕𝜕sw2 = (𝜕t1 ∕𝜕sw2 ) × t2 + t1 × (𝜕t2 ∕𝜕sw2 ) 𝜕nw ∕𝜕sw2 = (𝜕t1 ∕𝜕sw2 ) × t2 + t1 × (𝜕t2 ∕𝜕sw2 )
⎡gwp cos sw ⎤ ⎡ −(𝜕 2 gwp ∕𝜕(sw )2 ) cos sw ⎤
⎢ 2⎥ ⎢ 2 2 ⎥

=⎢ 0 ⎥ = ⎢ −(𝜕gwp ∕𝜕sw1 )(𝜕gwp ∕𝜕sw2 ) ⎥


⎢ ⎥ ⎢ ⎥
⎢gwp sin sw ⎥ ⎢ −(𝜕 2 gwp ∕𝜕(sw )2 ) sin sw ⎥
⎣ 2 ⎦ ⎣ 2 2 ⎦
⎡2(𝜕gwp ∕𝜕sw ) sin sw + gwp cos sw ⎤
⎢ 2 2 2 ⎥
+ ⎢−gwp (𝜕 2 gwp ∕𝜕sw1 𝜕sw2 ) ⎥
⎢ ⎥
⎢−2(𝜕gwp ∕𝜕sw ) cos sw + gwp sin sw ⎥
⎣ 2 2 2⎦

Similarly, the curvature vectors can be defined in the global coordinate system XYZ as
( 2 w) ( w)
𝜕 2 rw 𝜕tw1 w 𝜕 u w 𝜕t1 ⎫
= = A = A ⎪
𝜕(sw1 )2 𝜕sw1 𝜕(sw1 )2 𝜕sw1
( ) ( w ) ⎪
𝜕 2 rw 𝜕tw2 w 𝜕 2 uw w 𝜕t2 ⎪
w 2 = w =A =A ⎬ (4.9)
𝜕(s2 ) 𝜕s2 𝜕(s2 )
w 2
𝜕s2w

( 2 w ) ( w) ( w )⎪
𝜕 2 rw 𝜕tw1 𝜕tw2 𝜕 u 𝜕t1 𝜕t2
= = = A w
= A w
= A w

𝜕sw1 𝜕sw2 𝜕sw2 𝜕sw1 𝜕sw1 𝜕sw2 𝜕sw2 𝜕sw1 ⎭
The derivatives with respect to the surface parameters of the wheel surface normal vector
defined in the global XYZ coordinate system can be developed in a similar manner.

4.3 SEMI-ANALYTICAL APPROACH FOR RAIL


GEOMETRY
This section introduces the semi-analytical approach for describing rail geometry. The
ANCF interpolation approach is discussed in the following section. To allow for indepen-
dent movement of the rails of a track during dynamic simulations in both approaches,
the kinematics of each rail is defined separately to allow accounting for scenarios that
include gage variations, vertical rail movements, different motion constraints imposed on
the rails, and rail rotations. For this reason, each rail is assumed to have its own coordinate
system X r Y r Z r that is different from the track coordinate system X t Y t Z t , as shown in
Figure 4. Here, superscript r stands for the right rail (rr) or left rail (lr): that is, r = rr,
lr. As in the case of the wheel, the rail surface can be defined in a general parametric
form in terms of two surface parameters sr1 and sr2 . This rail surface description allows for
a general representation of the rail profile using analytical methods, tabulated data, or
measurements. As shown in Figure 5, and without any loss of generality, the two surface
parameters of the rail can be selected to be a longitudinal surface parameter sr1 = xsr and a
lateral surface parameter sr2 = yrs . The rail profile curve can be defined in a profile frame
X rp Y rp Z rp using the profile function grp . The origin of the profile frame is assumed to be
136 Mathematical Foundation of Railroad Vehicle Systems

lr rr
Z lr Z
X
lr rr
Z
t Y X
rr
Y
t t
Y X
O

Figure 4.4 Rail coordinate system.

located on the rail space curve, as shown in Figure 5. If the profile does not change as a
function of the longitudinal surface parameter, one can write grp = grp (sr2 ). If the profile
changes as a function of the longitudinal surface parameter, one has grp = grp (sr1 , sr2 ). These
two important cases are discussed in this section.

yrs
Z
rp xrs
rp
rp X
Y

Figure 4.5 Rail surface parameters.


Constant Rail Profile Curve If the rail profile curve does not change as a function of
the longitudinal surface parameter sr1 = xsr , the rail surface can be defined by a translation of
the profile function grp = grp (sr2 ) in the rail longitudinal direction. In this case, the location
of an arbitrary point on the rail surface can be defined in the rail coordinate system X r Y r Z r
using vector ur (sr1 , sr2 ) as
ur (sr1 , sr2 ) = Rrp + Arp urp (4.10)
[ rp rp rp ] T
where Rrp = Rrp (sr1 ) = xo (sr1 ) yo (sr1 ) zo (sr1 ) is the vector that defines the origin of the
rail profile frame with respect to the rail coordinate system X r Y r Z r , Arp = Arp (sr1 ) is the
matrix that defines the orientation of the profile frame X rp Y rp Z rp with respect to the rail
[ ]T
coordinate system X r Y r Z r , and urp = 0 sr2 grp (sr2 ) is the vector that defines the loca-
tion of an arbitrary point in the profile frame X rp Y rp Z rp . In making these definitions, it is
assumed that the X rp axis of the profile frame is along the longitudinal tangent of the rail
space curve. As discussed in Chapter 3, matrix Arp = Arp (sr1 ) in Eq. 10 depends only on the
arc length parameter sr1 of the rail space curve. This matrix, as explained in Chapter 3, can
be expressed in terms of three Euler angles that are not totally independent if some specific
conditions are imposed (Shabana and Rathod 2007). Euler angles are used in the design of
track geometry and in the track layout, as discussed in the preceding chapter and in later
sections of this chapter. Table 3 shows geometric details of rail sections, basic dimensions,
and rail profile geometries.
Railroad Geometry 137

Table 4.3 Rail geometry details.

Rail section
● Head is the portion of the rail
Running surface section that contains the profile
and running surface that
Side comes into contact with the
Head wheels of the railroad vehicle.
● Base (also called foot) is the
bottom section of the rail that
Fillet rests on the track structure.
Web
● Web is the section that connects
the head and the base.

Base
Flat-bottom rail

Flat-bottom rail
Basic dimensions

hw wh is the rail section height.


wb is the base width.
hd hw is head width.
ww is the web thickness at the
ww center point.
wh hfd hd is the head depth.
hfd is the fishing.
hbh
hbd hbd is the base depth.

wb hbh is the bolt-hole position.

Dimension example: 68 kg rail


74.6
73.4
Head width at gauge point
35.6
R14.3
16.0 gauge point
R254.0
1 IN 40

49.2
R31.8
1 IN 4
R7.9
R7.9 68 kg rail profile: head area
7.1 R203.0 3117 mm2 , web area 2349 mm2 ,
17.5 R19.1
foot area 3147 mm2
Centerline of 203 and 508 radii
Total area 8614 mm2 , estimated
185.7
Centerline bolt holes
Neutral axis mass 67.7 kg/m, maximum rail
length 25.0 m
R508.0

98.4
R19.1 78.6 85.0

1 IN 4
R3.2
30.2 R1.6

152.4

(Continued)
138 Mathematical Foundation of Railroad Vehicle Systems

Table 4.3 (Continued)

Other rail geometries

Crane rails are used in many


Bullhead rail was used metro areas and are available Guard rails (check rails) are
by the British railway from in sections that range from 104 placed parallel to running rails
the mid-nineteenth to mid- to 175 lb per yard. along restricted-clearance areas,
twentieth century. It has such as tunnels, bridges, and/or
been, with very few excep- level crossings to keep the rail
tions, replaced by the wheels in alignment and thus
flat-bottom rail. avoid derailment and damage to
the track structure.

Turnouts (Shabana et al. 2008)

Point section Lead section Crossing section Guard rail

A
B Tongue rail
C Lead rail
Wheelset

Tongue rail Stock rail

(a) (b) (c)

Track geometry is characterized by different types of discontinuities resulting from


non-smooth changes in the rail geometry. These geometric discontinuities, which produce
significant impulsive forces as the vehicle negotiates track sections, can be attributed
to misalignments, change in soil properties, a sudden change in the track structural
stiffness, or unavoidable structural designs, as is the case with switches and turnouts.
Discontinuities can also be encountered as a result of bridge abutments and road crossing.
A train direction of motion is changed using turnouts (switches and crossings), as shown in
Railroad Geometry 139

Table 3 (Schupp et al. 2004; Kassa et al. 2006). For the most part, a turnout is constructed
using a switch panel (point section), crossing panel (crossing section), and closure panel
(lead rail) that connect the switch panel with the crossing panel, as shown in Table 3
(Shabana et al. 2008). A movable swing nose is used for high-speed trains (Kono et al.
2005). When a railroad vehicle negotiates switches starting from the stock rail to the tongue
rail in the switch panel or the crossing panel, more than one wheel/rail contact point can
occur, as shown in Table 3, leading to significant impulsive forces that influence vehicle
dynamics and stability and may lead to derailments.
[ rp r ]T
The fact that vector Rrp = xorp (sr1 ) yrp r
o (s1 ) zo (s1 ) is assumed to depend on the
longitudinal surface parameter allows for modeling curved and spiral rail sections.
In the simpler case of a tangent track with fixed straight rail, this vector reduces to
[ rp ]T rp rp
Rrp = xorp (sr1 ) yrp
o zo , where yo and zo are constants and sr1 = xsr represents the longi-
tudinal position of an arbitrary point on the rail surface. The more general representation of
rp rp rp rp
Eq. 10, in which yo = yo (sr1 ) and zo = zo (sr1 ), allows describing systematically curved rail
sections. As in the case of wheels, the rail profile function grp can be described analytically
using tabulated data or measurements. In the two latter cases, interpolation methods can
be used to define the location of an arbitrary point on the rail surface. As discussed in
the following section, using ANCF finite elements allows for obtaining a much simpler
alternate expression to Eq. 10.

Non-Constant Rail Profile Curve For worn rails or when using track measurement
data, the rail profile curve may vary as a function of the rail longitudinal surface parameter
sr1 = xsr as well as the lateral rail surface parameter sr2 = yrs . If the rail profile curve changes
as a function of the longitudinal surface parameter sr1 = xsr , the rail surface can be defined
by the profile function grp = grp (sr1 , sr2 ) in the rail longitudinal direction. In this case, the
location of an arbitrary point on the rail surface can be defined in the rail coordinate system
X r Y r Z r using vector ur (sr1 , sr2 ) as
ur (sr1 , sr2 ) = Rrp (sr1 ) + Arp (sr1 )urp (sr1 , sr2 ) (4.11)
[ rp rp rp ]T
where Rrp = Rrp (sr1 ) = xo (sr1 ) yo (sr1 ) zo (sr1 ) is the vector that defines the origin of the
rail profile frame with respect to the rail coordinate system X r Y r Z r , Arp = Arp (sr1 ) is the
matrix that defines the orientation of the profile frame X rp Y rp Z rp with respect to the rail
[ ]T
coordinate system X r Y r Z r , and urp = 0 sr2 grp (sr1 , sr2 ) is the vector that defines the loca-
tion of an arbitrary point in the profile frame X rp Y rp Z rp . As with a constant rail profile
curve, it is assumed that the X rp axis of the profile frame is along the tangent of the rail
[ rp r ]T
space curve. Furthermore, the form of vector Rrp = xorp (sr1 ) yrp r
o (s1 ) zo (s1 ) , as in the case
of the constant rail profile curve, allows for modeling curved and spiral rail sections. For
a tangent track with fixed straight rail, this vector, as previously mentioned, reduces to
[ rp ]T rp rp
Rrp = xorp (sr1 ) yrp o zo , where yo and zo are constants, and sr1 = xsr represents the longitu-
dinal coordinate of an arbitrary point on the rail surface. The rail profile function grp (sr1 , sr2 )
can also be described analytically, using tabulated data, or using measurements. When tab-
ulated data or measurements are used to define the rail surface, interpolation methods can
be used to determine the location of an arbitrary point on the rail surface. Similar to the
140 Mathematical Foundation of Railroad Vehicle Systems

case of constant profile, using the ANCF interpolation approach allows for obtaining a more
general, and much simpler, alternate expression to Eq. 11.

Tangent Plane and Normal Vector For rigid rails, vector Rrp that defines the loca-
tion of the origin of the profile frame X rp Y rp Z rp with respect to the rail or track coordinate
system X r Y r Z r can be, without any loss of generality, assumed constant. Consequently,
the derivatives of this vector with respect to the surface parameters are equal to zero, and
this vector does not appear in the definition of the rail surface tangent and normal vec-
tors. However, this is not the case when the semi-analytical approach is used to describe
the geometry of the rails since vector Rrp that defines the location of the origin of the
profile frame X rp Y rp Z rp with respect to the rail coordinate system X r Y r Z r cannot, in gen-
eral, be assumed constant due to its dependence on the rail longitudinal surface param-
eter sr1 = xsr : that is, Rrp = Rrp (sr1 ). Therefore, as is clear from Eqs. 10 and 11, both vec-
tor function Rrp and scalar function grp are needed in order to define the coordinates of
arbitrary points on the rail surface as well as the derivatives of these coordinates at these
points. For unworn rails, scalar function grp depends only on the lateral surface parame-
ter sr2 = yrs : that is, grp = grp (sr2 ). For worn rails, or in case of using measured track data, grp
depends on the longitudinal surface parameter sr1 = xsr as well as the lateral surface parame-
ter sr2 = yrs : that is, grp = grp (sr1 , sr2 ). In both cases, profile function grp can be defined analyti-
cally or numerically using interpolation methods, including spline functions. As previously
mentioned, the numerical approach offers generality and allows for using measured track
data.
Formulating the wheel/rail contact conditions requires determining the tangent and nor-
mal vectors of the rail surface at contact points. The rail surface tangent and normal vectors
are also required in order to define the normal and tangential wheel/rail contact forces. The
r r
two tangent vectors t1 and t2 define the rail tangent plane and also define normal vector nr
in the rail coordinate system X r Y r Z r . Using the more general definition of the rail profile
function grp = grp (sr1 , sr2 ), the tangent and normal vectors are defined as
( rp )
𝜕ur 𝜕Rrp 𝜕A 𝜕urp
urp + Arp r ,⎫
r
t1 = r = +
𝜕s1 𝜕sr1 𝜕sr1 𝜕s1 ⎪

r 𝜕u r 𝜕R rp
rp 𝜕u
rp
⎬ (4.12)
t2 = r = + A ,
𝜕s2 𝜕sr2 𝜕sr2 ⎪
r r ⎪
nr = t1 × t2 ⎭
where vector ur of Eq. 10 or 11 defines the location of an arbitrary point on the rail
[ ]T
surface with respect to the rail coordinate system X r Y r Z r , 𝜕urp ∕𝜕sr1 = 0 0 𝜕grp ∕𝜕sr1 ,
[ ]T
and 𝜕urp ∕𝜕sr2 = 0 1 𝜕grp ∕𝜕sr2 . If Rrp is assumed to depend only on the longitudinal
rail surface parameter sr1 – that is, Rrp = Rrp (sr1 ) – then in the preceding equation, one has
r r
𝜕Rrp ∕𝜕sr2 = 𝟎, and in this case, the lateral tangent vector t2 reduces to t2 = Arp (𝜕urp ∕𝜕sr2 ).
Furthermore, all Euler angles that define the transformation matrix of the profile frame
X rp Y rp Z zp can be expressed in terms of the rail longitudinal surface parameters sr1 , as
discussed in the preceding chapter. If the rail unit normal vector must be computed, the
⁀ r r r r
unit normal to the rail surface is defined as n r = (t1 × t2 )∕|t1 × t2 |. The case in which the
Railroad Geometry 141

rail profile curve does not change as a function of the rail longitudinal surface parameter
sr1 can be obtained as a special case of Eq. 12 in which 𝜕grp ∕𝜕sr1 = 0 and, therefore,
𝜕urp ∕𝜕sr1 = 𝟎.

Higher Derivatives The computations of higher derivatives with respect to rail surface
parameters are necessary, particularly when solving nonlinear contact conditions numeri-
cally using the iterative Newton–Raphson procedure. Determining higher-order derivatives
requires evaluating the curvature vectors as well as the derivative of the vector normal to the
rail surface. The curvature vectors at an arbitrary point on the rail surface can be written as
r ( 2 rp ) ( rp ) rp
𝜕 2 ur 𝜕t1 𝜕 2 Rrp 𝜕 A rp 𝜕A 𝜕u rp 𝜕 u ⎫
2 rp
r 2 = r = r 2 + u +2 r +A
𝜕(s1 ) 𝜕s1 𝜕(s1 ) 𝜕(s1 )
r 2
𝜕s1
r
𝜕s1 𝜕(sr1 )2 ⎪
r

𝜕 2 ur 𝜕t2 𝜕 2 Rrp rp 𝜕u
rp ⎪
= r = +A ⎬ (4.13)
𝜕(sr2 )2 𝜕s2 𝜕(sr2 )2 𝜕(sr2 )2 ⎪
𝜕 2 ur 𝜕t1
r
𝜕t2
r
𝜕Rrp 𝜕Arp 𝜕urp ⎪
rp 𝜕u
rp
= = = + + A ⎪
𝜕sr1 𝜕sr2 𝜕sr2 𝜕sr1 𝜕sr1 𝜕sr2 𝜕sr1 𝜕sr2 𝜕sr1 𝜕sr2 ⎭
[ ] T [ ]T
where 𝜕 2 urp ∕𝜕(sr1 )2 = 0 0 𝜕 2 grp ∕𝜕(sr1 )2 , 𝜕 2 urp ∕𝜕(sr2 )2 = 0 0 𝜕 2 grp ∕𝜕(sr2 )2 , and
[ ]T
𝜕 2 urp ∕𝜕sr1 𝜕sr2 = 0 0 𝜕 2 grp ∕𝜕sr1 𝜕sr2 . The derivatives of the normal vector with respect
to the rail surface parameters can be determined using the derivatives of the two tangent
vectors as
r r r r }
𝜕nr ∕𝜕sr1 = (𝜕t1 ∕𝜕sr1 ) × t2 + t1 × (𝜕t2 ∕𝜕sr1 )
r r r r (4.14)
𝜕nr ∕𝜕sr2 = (𝜕t1 ∕𝜕sr2 ) × t2 + t1 × (𝜕t2 ∕𝜕sr2 )
These two equations show that, as in the case of a wheel, the derivatives of the normal
vector are functions of the rail curvature vectors. When using the unit normal vector
⁀ r r r r
n r = (t1 × t2 )∕|t1 × t2 |, the derivatives can also be evaluated in closed form, as previously
discussed:
⁀ ⁀ ⁀ T
𝜕n r ∕𝜕srk = ((I − n rn r )∕|nr |)(𝜕nr ∕𝜕srk ), k = 1, 2 (4.15)
In this equation, 𝜕nr ∕𝜕srk , k = 1, 2, can be determined using Eq. 14. Therefore, the deriva-
tives of the unit normal to the rail surface can be systematically evaluated using the tangent
vectors and their derivatives. While, as in the case of wheels, using a numerical approach
based on interpolations allows using the rail equations developed in this section, a much
simpler and more straightforward approach for describing rail geometry is to use ANCF
interpolations. Using ANCF geometry significantly simplifies the rail surface equations, as
discussed in the following section.

Rail Global Vectors In the formulations used in this book, the rail and/or track can have
arbitrarily large rigid-body displacements, including finite rotations. Using these formula-
tions may be necessary in some railroad scenarios that include foundation movements or
the failure of some rail segments. Using the equation for the position vector of an arbitrary
point with respect to the rail coordinate system X r Y r Z r , the position vector rr of arbitrary
142 Mathematical Foundation of Railroad Vehicle Systems

points can be defined in the global coordinate system XYZ as


rr = Rr + Ar ur (4.16)
where Rr = Rr (t) is the position vector of the origin of the selected rail coordinate sys-
tem X r Y r Z r in the global coordinate system XYZ, Ar is the transformation matrix that
defines the orientation of the rail coordinate system, t is time, and ur is defined by Eq. 11 as
ur (sr1 , sr2 ) = Rrp (sr1 ) + Arp (sr1 )urp (sr1 , sr2 ). Because Rr and Ar do not depend on the rail surface
parameters, one can show that the rail tangent and normal vectors defined in the global
coordinate system XYZ are given by
( r)
𝜕rr 𝜕u r ⎫
tr1 = r = Ar = Ar t1 ,
𝜕s1 𝜕sr1 ⎪
( r) ⎪
𝜕r r 𝜕u r ⎬ (4.17)
tr2 = r = Ar = A r
t ,
𝜕s2 𝜕sr2 2 ⎪
r r r r ⎪
nr = tr1 × tr2 = (Ar t1 ) × (Ar t2 ) = Ar (t1 × t2 ) = Ar nr ⎭
where the derivatives of ur with respect to surface parameters sr1 and sr2 are as previously
defined in this section. Similarly, the curvature vectors can be defined in the global coordi-
nate system XYZ as
( 2 r ) ( r)
𝜕 2 rr 𝜕tr1 r 𝜕 u r 𝜕t1 ⎫
= = A = A ⎪
𝜕(sr1 )2 𝜕sr1 𝜕(sr1 )2 𝜕sr1
( ) ( r ) ⎪
𝜕 2 rr 𝜕tr2 r 𝜕 2 ur r 𝜕t2 ⎪
r 2 = r =A =A ⎬ (4.18)
𝜕(s2 ) 𝜕s2 𝜕(s2 )
r 2
𝜕s2r

( 2 r ) ( r) ( r )⎪
𝜕 2 rr 𝜕tr1 𝜕tr2 𝜕 u 𝜕t1 𝜕t2
= = = A r
= A r
= A r

𝜕sr1 𝜕sr2 𝜕sr2 𝜕sr1 𝜕sr1 𝜕sr2 𝜕sr2 𝜕sr1 ⎭
The derivatives with respect to the surface parameters of the rail surface normal vector
defined in the global XYZ coordinate system can be developed in a similar manner.

4.4 ANCF RAIL GEOMETRY


It is clear from the analysis presented in the preceding section that the computer implemen-
tation of rail geometry equations can be cumbersome as compared to wheel geometry, even
for a rigid rail. The complexities arise from the dependence of more vectors and matrices
in the rail equations on rail surface parameters. Furthermore, the rail is constructed using
different section types that have different geometries, requiring the use of a more general
approach. Even in the case of a rigid rail, a data file that defines the rail geometry must be
developed at a preprocessing stage. It is necessary to have this data file containing the posi-
tion of points as well as rotations at these points in order to perform dynamic simulations
using general multibody system (MBS) algorithms. Rotations at discrete points are used
to define the geometry of the rail, including curvature, grade (elevation), and supereleva-
tion. Therefore, using the semi-analytical approach described in the preceding section can
significantly complicate the computer implementation of general rail geometry equations.
Railroad Geometry 143

Furthermore, such a semi-analytical approach does not lend itself easily to cases in which
the rail deforms. In such cases, using more general flexible-body computational approaches
is necessary.

ANCF Geometry One effective and general method to describe the rail surface geome-
try in railroad vehicle system applications is to use the ANCF finite elements introduced
in Chapter 2 (Shabana 2018). ANCF finite elements have been used effectively to integrate
geometry and analysis in railroad vehicle system applications (Berzeri et al. 2000). These
elements, which have been used to describe both rail and catenary geometries, are
well-suited for describing arbitrary rail geometry with different section types such as
tangent, curve, and spiral sections. ANCF elements allow for developing the solid model
geometry and using the same geometry model in the analysis without any adjustments.
Furthermore, these ANCF elements can be used for both rigid-rail and flexible-rail
analysis. For example, ANCF fully parameterized beam elements can be used to develop
the geometry of the rail space curve and superimpose on this curve geometry the profile
geometry to define the rail surface, as described in this section.
The main reason ANCF elements can significantly simplify describing the rail geom-
etry and its computer implementation is that ANCF elements use position gradients as
nodal coordinates. By changing the orientations of position gradient vectors at element
nodal points, curved elements can be easily constructed and used to define the rail space
curve for different section types. ANCF elements are also related to computational geometry
and computer-aided design (CAD) methods – such as B-splines (basis splines) and NURBS
(non-uniform rational B-splines) – by a linear mapping and therefore can be used to gener-
ate the same geometries developed by these CAD methods. Additionally, ANCF elements
are mechanics-based and do not require using control points, which are not material points,
as is the case with B-spline and NURBS representation.
For a node k of an ANCF rail mesh, the vector of the nodal coordinates at the node is
defined as
[ T T T T
]T
erk = rrk rrk x rrk
y rrk
z , k = 1, 2, … , nrn (4.19)

In this equation, rrk is the position of node k of rail r, which can be defined in any coordinate
system including the global, track, or rail coordinate system; rrk x = 𝜕r ∕𝜕x, ry = 𝜕r ∕𝜕y,
rk rk rk

and rz = 𝜕r ∕𝜕z are the position vector gradients; x, y, and z are the ANCF element param-
rk rk

eters or spatial coordinates; and nrn is the number of nodes in the rail ANCF mesh. For a
long track, the number of nodes nrn can be relatively large if geometric variations are to be
captured accurately.
When rail deformation is considered, the same ANCF mesh can be used as the analysis
mesh without any change. The ANCF equations of motion of the rail can be formulated
and solved to determine the nodal coordinates as functions of time. These coordinates can
be used with the ANCF displacement field to determine the change in the geometry as well
as the elastic forces that result from rail deformations. For a rigid rail, which is the focus of
the discussion in this chapter, the ANCF position vector gradients remain constant and can
be defined using three orthogonal unit vectors expressed in terms of the three Euler angles
144 Mathematical Foundation of Railroad Vehicle Systems

𝜓 rt (sr1 ), 𝜃 rt (sr1 ), and 𝜙rt (sr1 ) that define, respectively, the curvature, grade, and superelevation,
where sr1 is the rail arc length parameter. The Euler angle sequence used in the railroad lit-
erature is Z r , − Y r , and X r . As explained in the preceding chapter, these three Euler angles
completely define the rail geometry and can be calculated using specific track information.
Also as discussed in Chapter 3, the three Euler angles 𝜓 rt (sr1 ), 𝜃 rt (sr1 ), and 𝜙rt (sr1 ) are, in gen-
eral, independent unless additional conditions are enforced, as discussed in the literature
(Shabana and Rathod 2007). Nonetheless, these angles can be defined in terms of the arc
length parameter sr1 because a space curve is uniquely defined using one parameter, which
here is the arc length sr1 . This arc length parameter can also be used to conveniently define
the distance traveled by a wheel rolling and/or sliding on the rail. It was shown in Chapter 3
that an orthogonal transformation matrix that describes the rail geometry can be expressed
in terms of the three angles 𝜓 rt , 𝜃 rt , and 𝜙rt as
rt
⎡cos 𝜓 cos 𝜃 − sin 𝜓 cos 𝜙 + cos 𝜓 sin 𝜃 sin 𝜙 − sin 𝜓 sin 𝜙 − cos 𝜓 sin 𝜃 cos 𝜙⎤
Art = ⎢ sin 𝜓 cos 𝜃 cos 𝜓 cos 𝜙 + sin 𝜓 sin 𝜃 sin 𝜙 cos 𝜓 sin 𝜙 − sin 𝜓 sin 𝜃 cos 𝜙 ⎥
⎢ ⎥
⎣ sin 𝜃 − cos 𝜃 sin 𝜙 cos 𝜃 cos 𝜙 ⎦
(4.20)

By imposing additional conditions, one can show that the second column of this matrix rep-
resents the unit curvature vector, and this matrix reduces to the transformation matrix that
defines the Serret–Frenet frame (Shabana and Rathod 2007). Further discussion of these
additional conditions, however, is not required for the analysis presented in this chapter.
As discussed in later sections of this chapter, the track data file produced by a track pre-
processor computer program can be designed, based on a specified simple track geometry
description, to provide information about rail segments (elements) defined by nodes. The
information includes the positions of the rail nodes as well as the three Euler angles at
these nodes. Therefore, for each node k on the rail, the track data file defines the following
six-dimensional vector in a global, track, or rail coordinate system:
[ T ]T
prk = rrk 𝜓 rtk 𝜃 rtk 𝜙rtk , k = 1, 2, … , nrn (4.21)

Using this information, the ANCF gradient vectors in Eq. 19 for a rigid fixed rail can be
defined by equating the gradient vectors with the columns of the transformation matrix of
Eq. 20 as

⎡ cos 𝜓 rtk cos 𝜃 rtk ⎤ ⎡ − sin 𝜓 rtk cos 𝜙rtk + cos 𝜓 rtk sin 𝜃 rtk sin 𝜙rtk ⎤ ⎫
⎢ ⎥ ⎢ ⎥ ⎪
x = ⎢ sin 𝜓 ⎥, ⎥ ,⎪
rtk cos 𝜃 rtk
y = ⎢ cos 𝜓
rrk rrk rtk cos 𝜙rtk + sin 𝜓 rtk sin 𝜃 rtk sin 𝜙rtk

⎢ ⎥ ⎢ ⎥ ⎪
⎣ sin 𝜃 rtk ⎦ ⎣ − cos 𝜃 rtk sin 𝜙rtk ⎦ ⎪

⎡ − sin 𝜓 rtk sin 𝜙rtk − cos 𝜓 rtk sin 𝜃 rtk cos 𝜙rtk ⎤ ⎪
⎢ ⎥ ⎪
z = ⎢ cos 𝜓 ⎥, k = 1, 2, … , nrn
rrk rtk sin 𝜙rtk − sin 𝜓 rtk sin 𝜃 rtk cos 𝜙rtk

⎢ ⎥ ⎪
⎣ cos 𝜃 rtk cos 𝜙rtk ⎦ ⎭
(4.22)

Using this definition of the ANCF gradient vectors, the vector of geometry coordinates of
node k in Eq. 19 can be fully defined numerically. It is important to point out that in the
Railroad Geometry 145

ANCF interpolation approach, Euler angles are interpreted as geometry parameters used to
define the elements of the ANCF position-gradient vectors. This is a more accurate and gen-
eral interpretation of these angles, which are treated as field variables that depend on the rail
longitudinal surface parameter sr1 and can be used to define rail segment geometries. As will
be seen, using Euler angles to define the position vector gradients at the node allows using
a higher-order interpolation for the position coordinates and also avoids differentiating the
angles with respect to the arc length parameter sr1 of the rail space curve.

4.5 ANCF INTERPOLATION OF RAIL GEOMETRY


Using fully parameterized ANCF elements to define rail surface geometry significantly
simplifies the definitions of the position, tangent, and normal vectors as well as their deriva-
tives. This is because Euler angles are used to define the ANCF gradient coordinates at
discrete nodes, and these angles are no longer considered field variables. Consequently,
they are not differentiated with respect to the arc length parameter sr1 of the rail space curve.
Furthermore, when the ANCF interpolation approach is used, using vectors Rrp and matrix
Arp of the semi-analytical approach is not necessary. In addition to avoiding angle differen-
tiations, higher-order cubic position interpolation can be used based only on position and
gradient coordinates at the two nodes of the elements; this is mainly due to the depen-
dence of the ANCF position field on the Euler angles at the discrete nodal points. Using
the ANCF position vector gradients leads to a 2-node element with 24 nodal coordinates;
therefore, such an element is based on a cubic interpolation for the position coordinates, as
explained in this section.
For a fully parameterized three-dimensional ANCF element, the displacement field that
defines the position of an arbitrary point on an element j on the rail with respect to the rail
coordinate system X r Y r Z r can be written as
urj (x, y, z, t) = Srj (x, y, z)erj (t) (4.23)
In this equation, x, y, and z are the element volume parameters, t is time, urj is the position
vector of an arbitrary point on an element in the rail coordinate system, Srj is the element
shape function matrix, and erj is the element vector of nodal coordinates. If ANCF elements
are used to describe only geometry without considering rail deformations, then erj is con-
stant regardless of the rail rigid-body displacement, and the preceding equation becomes
independent of time. In this case, the rigid-body motion of the rail is described using the
coordinates of the rail or the track reference described by time-dependent vector Rr and rail
orientation matrix Ar . In this case of a rigid rail, vector urj at an arbitrary point becomes
independent of time. In this special and important case, the preceding equation reduces to
urj (x, y, z) = Srj (x, y, z)erj (4.24)
Since the focus in this chapter is on a numerical description of rail geometry, Eq. 24 is used
in this chapter. However, Eq. 23 can be used for a flexible rail to update the geometry due
to rail deformation. As discussed in Chapter 2, for a fully parameterized ANCF beam ele-
ment, the shape function matrix Srj and the vector of nodal coordinates erj can be defined
using interpolating polynomials that are cubic along x and linear along y and z. In the
146 Mathematical Foundation of Railroad Vehicle Systems

development presented in this section, the gradient vectors allow for defining arbitrarily
curved geometry; there is no loss of generality in assuming that within an element, sr2 = y.
The contact conditions can be used to determine sr1 ; and for an element or rail segment
defined by two nodes k and k − 1, the element longitudinal coordinates can be determined
rj(k−1) rj(k−1)
as x = sr1 − s1 , where s1 is the arc length parameter at the first node (k − 1) of ele-
ment j. Replacing the 24 polynomial coefficients with 24 element nodal coordinates, as
explained in Chapter 2, the shape function matrix Srj can be written as
[ ]
Srj = s1 I s2 I s3 I s4 I s5 I s6 I s7 I s8 I (4.25)
where shape functions si , i = 1, 2, … , 8 are defined as (Yakoub and Shabana 2001)
}
s1 = 1 − 3𝜉 2 + 2𝜉 3 , s2 = l(𝜉 − 2𝜉 2 + 𝜉 3 ), s3 = l(𝜂 − 𝜉𝜂), s4 = l(𝜍 − 𝜉𝜍),
s5 = 3𝜉 2 − 2𝜉 3 , s6 = l(−𝜉 2 + 𝜉 3 ), s7 = l𝜉𝜂, s8 = l𝜉𝜍
(4.26)
rj(k−1)
where l is the length of the element. 𝜉 = x∕l = − 𝜂 = y∕l =
(sr1 s1 and 𝜍 = z/l.
)∕l, sr2 ∕l,
To define the rail surface, parameter z must be expressed as a function of x and y.

Surface Parameterization The surface of an element can assume an arbitrary shape


by writing coordinate z as a function of the two other coordinates x and y. To this end, one
can write z = f (x, y), where f is a function that defines the shape of the surface. For a rail
surface, for example, one has z = f (y) if the rail profile does not depend on the longitudinal
parameter x. Using the equation z = f (x, y) implies that an element surface is defined using
two parameters sr1 = x and sr2 = y. That is, on the surface of the element, one has
urj (x, y, z) = urj (x, y, f (x, y)) = urj (x, y) = urj (sr1 , sr2 ) = Srj (sr1 , sr2 )erj (4.27)
The function z = f (x, y) can assume any form and can also be represented numerically in
a tabulated form. In the special and important case z = f (y), a cubic spline function can
be used to describe this function based on tabulated data that can be obtained for the rail
profile using MiniProf device measurements.

4.6 ANCF COMPUTATION OF TANGENTS AND


NORMAL

In Eq. 27, a volume representation is converted to a surface representation using the


function z = f (x, y). Using this ANCF geometric approach, much simpler expressions –
compared to the expressions of the semi-analytical approach previously presented in this
chapter – can be developed for the tangent and normal vectors at arbitrary points on the
rail surface.
Using Eq. 27, one can write durj = (𝜕urj ∕𝜕x)dx + (𝜕urj ∕𝜕y)dy + (𝜕urj ∕𝜕z)dz. Using the
functional relationship z = f (x, y), one has dz = (𝜕f /𝜕x)dx + (𝜕f /𝜕y)dy. It follows that
durj = (𝜕urj ∕𝜕x)dx + (𝜕urj ∕𝜕y)dy + (𝜕urj ∕𝜕z)dz
= ((𝜕urj ∕𝜕x) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕x))dx + ((𝜕urj ∕𝜕y) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕y))dy (4.28)
Railroad Geometry 147

This equation defines two tangent vectors at an arbitrary point on the surface of an element
as
rj }
t1 = (𝜕urj ∕𝜕sr1 ) = (𝜕urj ∕𝜕x) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕x)
rj
(4.29)
t2 = (𝜕urj ∕𝜕sr2 ) = (𝜕urj ∕𝜕y) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕y)

It is important to recognize the difference between x and y when used with and without the
relationship z = f (x, y). In the case when the relationship z = f (x, y) is not used, the vol-
ume parameters x, y, and z are independent, and the three tangent vectors at any point x,
y, and z of an element (𝜕urj ∕𝜕x), (𝜕urj ∕𝜕y), and (𝜕urj ∕𝜕z) are independent. However, if the
surface of an element is specified by the function z = f (x, y), one has only two independent
parameters s1 = x and s2 = y, and the interpretation of the parameters x and y in this case is
different from their interpretation when z is not specified on the surface. When the function
relationship z = f (x, y) is used, variations are allowed only on a surface. This is clear from the
rj
definition of the tangent vectors given by Eq. 29 in which t1 = (𝜕urj ∕𝜕x) = (𝜕urj ∕𝜕sr1 ) with
rj
x = sr1 and t2 = (𝜕urj ∕𝜕y) = (𝜕urj ∕𝜕sr2 ) with y = sr2 have different directions, magnitudes,
and interpretations as compared to the case when the relationship z = f (x, y) is not used.
Using Eq. 29, the unit normal vector to the surface of an element can be defined as

⁀ (𝜕urj ∕𝜕sr1 ) × (𝜕urj ∕𝜕sr2 )


n rj =
|(𝜕urj ∕𝜕sr1 ) × (𝜕urj ∕𝜕sr2 )|
((𝜕urj ∕𝜕x) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕x)) × ((𝜕urj ∕𝜕y) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕y))
= (4.30)
|((𝜕urj ∕𝜕x) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕x)) × ((𝜕urj ∕𝜕y) + (𝜕urj ∕𝜕z)(𝜕f ∕𝜕y))|

As in the analytical approach, one can use the equation for position vector urj of an arbitrary
point with respect to the rail coordinate system X r Y r Z r to determine position vector rrj of
arbitrary points on ANCF element j in the global coordinate system XYZ as

rrj = Rr + Ar urj (4.31)

where Rr is the position vector of the origin of the selected rail coordinate system X r Y r Z r
in the global coordinate system XYZ, Ar is the transformation matrix that defines the ori-
entation of the rail coordinate system, and urj is defined by Eq. 27, which is much simpler
than Eq. 11 given by ur (sr1 , sr2 ) = Rrp (sr1 ) + Arp (sr1 )urp (sr1 , sr2 ) in the semi-analytical approach.
Because Rr and Ar do not depend on the rail surface parameters, one can write the rail
tangent and normal vectors defined in the global coordinate system XYZ as

( rj ) ⎫
rj 𝜕rrj 𝜕u rj ⎪
t1 = = A r
= Ar t1 ,
𝜕sr1 𝜕sr1 ⎪
( rj ) ⎪
rj 𝜕rrj 𝜕u rj ⎬ (4.32)
t2 = r = Ar = Ar t2 ,
𝜕s2 𝜕sr2 ⎪
rj rj rj rj rj rj ⎪
nrj = t1 × t2 = (Ar t1 ) × (Ar t2 ) = Ar (t1 × t2 ) = Ar nrj ⎪

where the derivatives of urj with respect to the surface parameters sr1 and sr2 are as previ-
ously defined in this section. Similarly, the curvature vectors can be defined in the global
148 Mathematical Foundation of Railroad Vehicle Systems

coordinate system XYZ as


( ) ( )
𝜕 2 rrj 𝜕t1
rj
𝜕 2 urj
rj
𝜕t1 ⎫
r 2 = = Ar =A r

𝜕(s1 ) 𝜕sr1 𝜕(sr1 )2 𝜕sr1 ⎪
( ) ( rj ) ⎪
rj
𝜕 2 rrj 𝜕t2 r 𝜕 2 urj r 𝜕t2 ⎪
r 2 =
𝜕(s2 ) r =A
𝜕s2 𝜕(s2 )
r 2 =A
𝜕s2
r ⎬ (4.33)

( 2 rj ) ( rj
) ( )
rj ⎪
rj rj
𝜕 r
2 rj 𝜕t 𝜕t 𝜕 u 𝜕t1 𝜕t2 ⎪
= 1r = 2r = Ar = Ar = Ar
𝜕sr1 𝜕sr2 𝜕s2 𝜕s1 𝜕sr1 𝜕sr2 𝜕sr2 𝜕sr1 ⎪⎭
The derivatives with respect to the surface parameters of the rail surface normal vector
defined in the global XYZ coordinate system can be developed in a similar manner. The
rj rj
tangent vectors t1 = (𝜕rrj ∕𝜕x) = (𝜕rrj ∕𝜕sr1 ) with x = sr1 and t2 = (𝜕rrj ∕𝜕y) = (𝜕rrj ∕𝜕sr2 ) with
y = sr2 can be used to define the coefficients of the first fundamental form. The definition
of the normal vector given in Eq. 32 and the second derivatives of vector rrj with respect
to x = sr1 and y = sr2 can be used to define the coefficients of the second fundamental form.
These derivatives can be systematically evaluated using the element shape function matrix
and vector of nodal coordinates of Eq. 27. The coefficients of the first and second fundamen-
tal forms of surfaces obtained using ANCF geometry can be used to systematically define
the principal curvatures and directions of surfaces at wheel/rail contact points using the
procedure discussed in Chapter 2.
Based on a simple track description, a track preprocessor computer program can be used
to determine Euler angles at a number of discrete nodal points specified by the program
user. Two choices can be made when discrete-node data are used in the algorithm of the
main processor. The first choice is to convert discrete Euler angles to field variables by using
linear interpolation within a segment and then use this angle field representation with the
semi-analytical approach to define the position, tangent, normal, and curvature vectors as
previously described in this chapter. This approach requires interpolating vector Rrp (sr1 )
and defining matrix Arp of the semi-analytical approach. In the second approach, Euler
angles at the discrete nodal points are used to evaluate ANCF position gradients; in this case,
the angles are not treated as field variables and are not interpolated during the dynamic
simulations. Furthermore, there is no need to define vector Rrp (sr1 ) and matrix Arp . This
ANCF interpolation approach also allows for having a cubic interpolation for the position
coordinates despite the fact that only two pieces of information are provided at the element
nodes for each coordinate.

4.7 TRACK GEOMETRY EQUATIONS


As shown in the preceding section, using the ANCF interpolation approach can signifi-
cantly simplify the calculations of the position, tangent, normal, and curvature vectors that
are required to define rail geometry. Because position vector gradients are used as nodal
coordinates, ANCF elements allow for describing complex geometry with a high degree of
accuracy. Euler angles at selected nodes on a rail or track space curve can be used to define
ANCF position-gradient vectors. These Euler angles completely define the rail space curve
geometric properties.
Railroad Geometry 149

This section summarizes some of the basic equations obtained in Chapter 3, which
are the basis for computing track geometry. These equations are used to explain how
simple track input data can be used to develop a track file to be used in dynamic computer
simulations of railroad vehicle system models. This track file defines the position of and
Euler angles at selected nodes on the track space curves. This information, which defines
the track geometry, is used as an input to an MBS computer program that automatically
constructs and numerically solves the nonlinear dynamic equations of complex railroad
vehicle systems. In the development presented in the remaining sections of this chapter, to
develop general railroad vehicle system models in which two rails can move independently,
it is assumed that the track file has three different meshes that correspond to three different
space curves: the track space curve, sometimes referred to as the track centerline, which is
used as a reference to define vehicle locations and forward velocities; the right rail space
curve; and the left rail space curve. The locations and geometric properties of these space
curves are completely defined by the position of and Euler angles at the nodes of their
space curves that are constructed using simple industry inputs as described in this chapter.
Since the procedures described are applicable to any space curve (track, and right and left
rails), the superscripts are dropped for the sake of simplicity and generality.
In practice, track geometry is defined using simple, specific data provided at the points
where the geometry changes. Only three inputs are used to define each rail segment
(tangent, curve, and spiral) regardless of the length of the segment (Dukkipati and Amyot
1988; Berzeri et al. 2000; Rathod and Shabana 2006a). These simple inputs, as discussed in
Chapter 2, are a projection of the rail space curve on the horizontal plane to define a planar
curve with curvature CH , called the horizontal curvature; a development or grade that defines
a vertical-development angle 𝜃 about the rail or track Y axis; and a superelevation that defines
a bank angle 𝜙 about the longitudinal tangent to the rail or track space curve. The three vari-
ables CH , 𝜃, and 𝜙 completely and uniquely define the position coordinates and geometry
of any track segment. Table 4 shows an example of the data used for defining the geometry
of the track shown in Figure 6. These data are also used as input to the track preprocessor.

387.76
10
Global Y coordinate

301.04 9
214.32
8
127.60 7
5 6
4
40.88 1 2 3

0 149.58 448.75 747.92 1047.10


Global X coordinate

Figure 4.6 Track described in the data in Table 4.

As discussed in Chapter 3, a track or rail space curve can be defined by consid-


ering the first column in the transformation matrix of Eq. 20 as the unit tangent
vector. In this case, one can write a segment of the space curve as dr = (𝜕r∕𝜕s)ds =
[ ]T
cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 sin 𝜃 ds. The projection of this segment on the horizontal plane is
150 Mathematical Foundation of Railroad Vehicle Systems

Table 4.4 Example of input data to a track preprocessor. Source:


Courtesy of Shabana, A.A., Zaazaa, K.E., and Sugiyama, H.

Distance Curvature Superelevation Grade


Node # (ft) (∘ ) (in.) (%)

1 0 0 0 0
2 100 0 0 0
3 150 5 1.5 0
4 450 5 1.5 0
5 500 −3 −1 0
6 650 −3 −1 0
7 720 7 2 0
8 1020 7 2 0
9 1145 0 0 0
10 1195 0 0 0

[ ]T
defined by the equation drH = cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 0 ds. By factoring out cos𝜃, one
[ ]T
can write drH = cos 𝜓 sin 𝜓 0 cos 𝜃ds. This equation, as discussed in Chapter 2, defines
vector tH tangent to and arc length dS of the projected planar curve, respectively, as
[ ]T
tH = cos 𝜓 sin 𝜓 0 , dS = cos 𝜃ds (4.34)
Using the relationship between space curve arc length s and arc length S of the projected
planar curve, one has 𝜕𝜓/𝜕S = (𝜕𝜓/𝜕s)(𝜕s/𝜕S), where 𝜕s/𝜕S = 1/ cos 𝜃. The curvature vector
of the projected planar curve is defined as
⎡− sin 𝜓 ⎤
𝜕tH ⎢ 𝜕𝜓
CH = = cos 𝜓 ⎥ (4.35)
𝜕S ⎢ ⎥ 𝜕S
⎣ 0 ⎦
This equation defines the horizontal curvature of the projected planar curve, which is the
magnitude of curvature vector CH , as CH = |CH | = 𝜕𝜓/𝜕S, or
dS
d𝜓 = CH dS = (4.36)
RH
In this equation, RH is the radius of curvature of the projected planar curve. It is clear from
this equation that angle 𝜓 can be determined by integration if the horizontal curvature CH
is given. The other two Euler angles 𝜃 and 𝜙 are assumed known from the grade (eleva-
tion) and superelevation (bank) data. Therefore, knowing the Euler angles, the preceding
equations can be used to define angle 𝜓 over a domain defined by endpoints So and S1 as
S1
𝜓 = 𝜓(S) = 𝜓o + (CH (S) )dS (4.37)
∫So
Railroad Geometry 151

C1 s

dr
C2
Z

Y
O
X
C1H

C2H
drH

Figure 4.7 Horizontal curvature.

where 𝜓 o is the value of 𝜓 at So . Figure 7 shows an illustration of the curve projection just
discussed. In this figure, the segment dr of a space curve defined by endpoints C1 and C2 and
whose arc length is s is projected on the horizontal plane to define a planar curve segment
drH that has endpoints CH1 and CH2 and arc length S.
Recall that the sequence Z, − Y , − X of Euler angles is selected to have positive definitions
for the horizontal curvature, grade, and superelevation. Curvature is considered positive if it
is the result of a positive rotation 𝜓 about the Z axis; vertical development is considered pos-
itive if it is the result of rotation 𝜃 about the −Y axis; and superelevation is considered
positive if it is the result of rotation 𝜙 about the −X axis. This convention is consistent
with measurements made in practice. It is also important to point out that the equations
presented in this section remain applicable when the track is not fixed. In such cases, the
equations use the horizontal plane before displacements.
In addition to determining angles at arbitrary points on the space curve from track input
data, one needs to determine the position coordinates of these points. Using the first column
of the transformation matrix of Eq. 20 as the tangent vector to the space curve allows writing
[ ]T
the curve equation in the differential form dr = cos 𝜓 cos 𝜃 sin 𝜓 cos 𝜃 sin 𝜃 ds. Using
Eq. 34, which defines the relationship between s and S, the equation for dr can be written
as
⎡cos 𝜓dS⎤
dr = ⎢ sin 𝜓dS ⎥ (4.38)
⎢ ⎥
⎣ sin 𝜃ds ⎦
152 Mathematical Foundation of Railroad Vehicle Systems

[ ]T
This equation, upon integration, can be used to define the curve r = r1 r2 r3 in its para-
metric form as

⎡ro1 ⎤ ⎡∫So cos 𝜓dS⎤ ⎡r1 (S)⎤


S1

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
r = ⎢ro2 ⎥ + ⎢ ∫S 1 sin 𝜓dS ⎥ = ⎢r2 (S)⎥
S
(4.39)
⎢ ⎥ ⎢ s1 ⎢ o ⎥ ⎢ ⎥

⎣ro3 ⎦ ⎣ ∫s sin 𝜃ds ⎦ ⎣ r3 (s) ⎦
o

[ ]T
In this equation, ro = ro1 ro2 ro3 is vector r at the start of the segment This equation
shows clearly that the coordinates of points on the space curve can be determined using
the two Euler angles, which are assumed to be known from the horizontal curvature and
grade. That is, if CH and 𝜃 are known from the track inputs, the locations of the points on
the space curve and Euler angles that define the curve geometry can be completely defined
for tangent, curve, and spiral rail sections. The tangent section has zero curvature, the curve
section has constant curvature, and the spiral section is designed to have linearly varying
horizontal curvature, as discussed in the following section.

4.8 NUMERICAL REPRESENTATION OF TRACK


GEOMETRY

Using the three parameters – horizontal curvature CH , vertical-development angle 𝜃, and


bank angle 𝜙 – a numerical procedure based on simple interpolation can be used to deter-
mine the position of the nodes and the three Euler angles at these nodes for the track, right
rail, and left rail space curves. Using the numerical procedure is necessary since it is not
possible to obtain a closed-form solution for the track equations presented in the preced-
ing section. The nodal position and angle data, which can be determined in a preprocessor
computer program prior to dynamic simulations, are required to determine the locations
of wheel/rail contact points when the governing dynamic equations are numerically solved
by the main processor computer program.
As discussed in Chapter 3, track consists of a few segment types. In the design of the
track segments, the rail is divided into segments, and geometric variables are assumed to
vary linearly within a segment. Using linear interpolation for the geometry variables allows
obtaining a solution with end conditions that are provided in the description of the track.
Using higher-order interpolation requires additional data that are generally not available.
Linear interpolation can be used to define the position coordinates of and Euler angles at
an arbitrary number of nodes specified by the user.

Rail Segment Types As discussed in Chapter 3, the track segment types include a
tangent segment that represents a straight section with zero horizontal curvature: that is,
CH = 0; a curve segment with constant horizontal curvature CH = 1/RH ; a tangent-to-curve
entry spiral section with horizontal curvature CH = 1/RH that varies linearly as a function of
the projected arc length S from zero at the first point to the value of the horizontal curvature
of the curve at the segment endpoint; a curve-to-tangent exit spiral with horizontal curva-
ture CH = 1/RH that varies linearly as a function of the projected arc length S from the value
Railroad Geometry 153

of the horizontal curvature of the curve at the first point to zero at the segment endpoint;
and a curve-to-curve spiral section with horizontal curvature CH = 1/RH that varies linearly
as a function of the projected arc length S from the value of the horizontal curvature of
the first curve at the first point to the value of the horizontal curvature of the second curve
at the segment endpoint. The linear variation of vertical-development angle 𝜃 and bank
angle 𝜙 within the segments is also assumed in most algorithms used to construct the track
layout.

Linear Interpolation of Angles Assuming that horizontal curvature CH varies lin-


early within a rail segment, the following linear interpolation can be used for CH within
the segment
CH = (1 − 𝜉S )CH1 + 𝜉S CH2 (4.40)
where 𝜉 S = (S − S1 )/LS , LS = (S2 − S1 ), and CH1 , CH2 , S1 , and S2 are the values of the hor-
izontal curvature CH and the projected arc length S at the first and second nodes of the
segment, respectively. Using the linear interpolation of horizontal curvature CH defined by
the preceding equation and Eq. 37, which defines angle 𝜓, one can obtain a closed-form
expression of 𝜓 at an arbitrary point S within the segment in terms of the values of hori-
zontal curvatures CH1 and CH2 at the two segment endpoints as
1
𝜓(S) = 𝜓1 + [C (S − S1 )2 + CH1 ((LS )2 − (S − S2 )2 )] (4.41)
2LS H2
In this equation, 𝜓 1 defines angle 𝜓 at the first node of the segment. The preceding equation
can be used to define angle 𝜓 at the desired discrete nodal points used in the track file
produced by the track preprocessor.
Vertical-development angle 𝜃 is assumed to vary with respect to arc length s according to
the equation
d𝜃 = CV ds (4.42)
where CV is a constant referred to in the railroad literature as the vertical curvature. As
shown in Chapter 3, the curvature vector of a space curve can be written in terms of Euler
angles and their derivatives as
⎡−𝜓 ′ sin 𝜓 cos 𝜃 − 𝜃 ′ cos 𝜓 sin 𝜃 ⎤
c = ⎢ 𝜓 ′ cos 𝜓 cos 𝜃 − 𝜃 ′ sin 𝜓 sin 𝜃 ⎥ (4.43)
⎢ ⎥
⎣ 𝜃 ′ cos 𝜃 ⎦
where a′ = 𝜕a/𝜕s. This curvature vector can be written as a linear combination of two
orthogonal unit vectors as
⎡− sin 𝜓 ⎤ ⎡− cos 𝜓 sin 𝜃 ⎤
c = 𝜓 ′ cos 𝜃 ⎢ cos 𝜓 ⎥ + 𝜃 ′ ⎢ − sin 𝜓 sin 𝜃 ⎥ (4.44)
⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦ ⎣ cos 𝜃 ⎦
The two orthogonal unit vectors in this equation are also orthogonal to the unit
tangent to the space curve defined by the first column of the transformation matrix
in Eq. 20. Therefore, the curvature vector has two components: 𝜓 ′ cos 𝜃 along
[ ]T
the unit vector − sin 𝜓 cos 𝜓 0 and 𝜃 ′ = 𝜕𝜃/𝜕s along the unit vector
154 Mathematical Foundation of Railroad Vehicle Systems

[ ]T [ ]T
− cos 𝜓 sin 𝜃 − sin 𝜓 sin 𝜃 cos 𝜃 . The vector − cos 𝜓 sin 𝜃 − sin 𝜓 sin 𝜃 cos 𝜃
[ ]T
is the same as the vertical vector 0 0 1 before performing the two rotations 𝜓 and 𝜃;
therefore, this vector is perpendicular to the plane of the track before the superelevation,
justifying calling CV the vertical curvature. Assuming that 𝜃 varies linearly with respect
to s, the vertical curvature becomes constant within the segment and can be written as
d𝜃 𝜃 − 𝜃1
CV = = 2 (4.45)
ds s 2 − s1
where 𝜃 1 and 𝜃 2 are the vertical-development angles at the segment endpoints defined by
the arc length parameters s1 and s2 , respectively. It follows that vertical-development angle
𝜃 at an arbitrary point within the segment can be written as
𝜃(s) = 𝜃1 + CV (s − s1 ) (4.46)
This equation can be used to define angle 𝜃 at the desired discrete nodal points used in the
track file produced by the track preprocessor.
Similarly, bank angle 𝜙 can be defined using a function that is linear in the projected arc
length S as
𝜙 = 𝜙(S) = (1 − 𝜉S )𝜙1 + 𝜉S 𝜙2
𝜙 (S − S1 ) − 𝜙1 (S − S2 )
= 2 (4.47)
(S2 − S1 )
where 𝜉 S = S/LS , LS = S2 − S1 , and 𝜙1 and 𝜙2 are the bank angles at the two end nodes of
the segments defined, respectively, by the projected arc length coordinates S1 and S2 . As
in the case of the other two angles 𝜓 and 𝜃, the preceding equation can be used to define
angle 𝜙 at the desired discrete nodal points used in the track file produced by the track pre-
processor. The distance between the nodes can be selected to ensure achieving the desired
accuracy of the interpolation during computer simulations. Using a 1 ft distance between
nodes in a track file is common, so the track file can be very large.
Recent investigations have demonstrated that the three Euler angles 𝜓, 𝜃, and 𝜙 used to
describe the geometry of a curve cannot all be treated as independent geometric variables.
Therefore, the rail industry numerical procedure described in this section for determining
the angles must be carefully evaluated for spiral geometry. The tangent and curve segments
of a track have zero and constant curvature, respectively. A curve with constant curvature
should, in general, be superelevated using a constant bank angle to achieve balance speed
requirements. This is not the case, however, for a spiral segment that has varying curvature.
Bank angle 𝜙, as explained in the rail industry procedure for numerically constructing track
geometry, is used with the other angles 𝜓 and 𝜃 to determine the orientation of the track
(body trajectory) frames, but such a linearly interpolated bank angle does not enter into
the definition of the spiral geometry, which is completely defined using the angles 𝜓 and 𝜃
(Shabana and Ling, 2019).

Numerical Evaluation of Position Coordinates As shown in Table 4, track geom-


etry is described by using simple inputs in a preprocessor computer program to define the
positions and Euler angles at discrete points on the rail or track space curve. Regardless of
the length of a track or rail segment (tangent, curve, or spiral), only inputs at track locations
Railroad Geometry 155

where the geometry changes can be provided. That is, for a curve section of the track, which
can be miles long, information is provided only at the beginning and end of the section.
Linear interpolations discussed in this section are used in practice in the track layout. How-
ever, to determine the positions of the track nodes, numerical integration is necessary since
closed-form solutions cannot be obtained.
Recall from Eq. 39 that position coordinates are defined in terms of angles using the
S S
integral forms r1 (S) = ro1 + ∫S 1 cos 𝜓dS, r2 (S) = ro2 + ∫S 1 sin 𝜓dS, and r3 (s) = ro3 +
o o
s
∫s 1 sin 𝜃ds. While angle 𝜓 is defined in terms of projected arc length S, angle 𝜃 is defined
o
in terms of the space curve arc length s. Equation 34 shows that S and s are related by the
equation dS = cos 𝜃ds, and Eq. 42 shows that d𝜃 = CV ds: that is, ds = d𝜃/CV . It follows that
dS = cos 𝜃(d𝜃/CV ): that is,
d𝜃 C
= V (4.48)
dS cos 𝜃
which defines the derivative of vertical-development angle 𝜃 with respect to projected arc
length S. The equation dS = (cos𝜃/CV )d𝜃, upon integration, can be used to define the pro-
jected arc length within a segment as (Berzeri et al. 2000)
{
S1 + C1 (sin 𝜃(s) − sin 𝜃1 ) if CV ≠ 0
S(s) = V
(4.49)
S1 + (s − s1 ) cos 𝜃(s) if CV = 0
The definition of S = S(s) from this equation can be used to evaluate the angles in Eqs. 41 and
47 as well as the coordinates of the nodal points on the space curve defined using numer-
S S
ical integration of the equations r1 (S) = ro1 + ∫S 1 cos 𝜓dS, r2 (S) = ro2 + ∫S 1 sin 𝜓dS, and
o o
s
r3 (s) = ro3 + ∫s 1 sin 𝜃ds.
o

Summary It is clear from the analysis presented in this section that knowing the horizon-
tal curvature CH , vertical-development angle 𝜃, and bank angle 𝜙 at specific points where
track geometry changes allows developing a numerical procedure based on interpolation to
determine the angles, which can be used with numerical integration to define the position
coordinates and angles at an arbitrary number of points on the rail or track space curve
as a function of arc length parameter s. This is typically accomplished, as previously men-
tioned, in a track preprocessor that generates a track file that has, for a given space curve,
a number of nodes specified by the user. For each node of a space curve, the file provides
the corresponding entries s, r 1 , r 2 , r 3 , 𝜓, 𝜃, and 𝜙. Additional information, such as hori-
zontal curvature CH at the nodes, can also be printed out by the track preprocessor. An
organization of such a track preprocessor is discussed in the following section.

4.9 TRACK DATA


Before performing dynamic simulations of railroad vehicle systems, a data file that
describes the geometry of the track space curve and the right and left rail space curves is
often be created in a preprocessor computer program. In this track data file, the track and
rail space curve geometries are described using the positions of and Euler angles at discrete
nodes determined using the equations presented in the preceding section. During dynamic
156 Mathematical Foundation of Railroad Vehicle Systems

simulations, wheel/rail contact conditions are used to solve for surface parameters, as
explained in Chapter 6. These surface parameters define points on the rails that come into
contact with the wheels. Using the value of the rail longitudinal surface parameters s at
the contact point, the segment of rail within which contact occurs can be identified. The
positions and Euler angles at the two ends of this segment can be used with interpolation
methods to determine the tangent, normal, and curvature vectors previously defined in
this chapter. By using ANCF finite elements, interpolation of the angles is avoided, and the
position vectors that define the tangent, normal, and curvature vectors are defined using
cubic interpolations.

Track Preprocessor Using a track preprocessor to develop a track data file prior to
dynamic simulations is recommended for the computational efficiency of the solution algo-
rithm. The input to the track preprocessor, as shown by the data in Table 4, is simple and
defines horizontal curvature CH ; the grade defining vertical-development angle 𝜃; and the
superelevation, which defines bank angle 𝜙. As described in the preceding section, these
three inputs can be used to completely define the track nodal positions and Euler angles
at an arbitrary number of points on the rail space curve specified by the user. More nodal
points lead to more accurate interpolations; therefore, the track file for long tracks can be
relatively large, as previously mentioned. It is also important to note that the three inputs
(horizontal curvature, grade, and superelevation) need to be provided only at track points
where the geometry changes, as previously mentioned. Therefore, regardless of the num-
ber of nodal points selected by analysts, the input file to the track preprocessor contains a
relatively small amount of data.
The track preprocessor can be designed to read a set of simple standard input data that are
sufficient for creating discrete-point information used in nonlinear dynamic simulations of
a vehicle/track system. It is clear from the analysis presented in the preceding section that
the track preprocessor is required to perform the following specific functions (Berzeri et al.
2000):
1. Read the input data that define the geometry of the track segments along the track center-
line. These input data define horizontal curvature CH , the grade, and the superelevation
only at points where the track geometry changes. The grade and superelevation are used
to define vertical-development angle 𝜃 and bank angle 𝜙, respectively, at the specified
points.
2. Use horizontal curvature CH , vertical-development angle 𝜃, and bank angle 𝜙 with linear
interpolation to create field variables CH (S), 𝜃(s), and 𝜙(S). Using horizontal curvature
CH , angle 𝜓(S) can be defined using Eq. 41. Therefore, the values of the three Euler
angles 𝜓, 𝜃, and 𝜙 can be determined at the discrete nodal points selected by the user
using linear interpolation in the track preprocessor.
3. Use Euler angles 𝜓(S) and 𝜃(s) and the methods of numerical integration to compute
coordinates r 1 , r 2 , and r 3 of the nodes selected by the user on the track centerline, as
discussed in the preceding section. The nodal positions can be defined with respect to the
track body coordinate system in order to develop more general algorithms for simulation
scenarios in which the track is not fixed.
4. Use track space curve data to define the right and left rail space curves in terms of discrete
points. The right and left rail space curves can have different position coordinates and
Railroad Geometry 157

different Euler angles at their nodes to allow for an independent description of the rail
geometries. The different nodes of the different space curves can be due to supereleva-
tion, track irregularities, curvature, etc. There is no loss of generality in assuming that the
track and rail body coordinate systems initially coincide, since the dynamic simulation
algorithm can allow these coordinate systems to move independently.
5. Produce an output file that can be read by the main processor and used in dynamic sim-
ulations. This output file can define the track and rail geometry mesh at discrete nodal
points. For node
[ k on space curve j, the]track preprocessor can provide, at minimum, the
information sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk , where j = t, rr, lr, and t, rr, and lr refer, respec-
tively, to the track, right rail, and left rail space curves. Additional information, such as
the horizontal curvature, can also be included in the output file. ANCF interpolation
can be used during dynamic simulations to obtain the tangent, normal, and curvature
vectors at points within the segments, as previously discussed in this chapter.

Input Data to the Track Preprocessor As presented in Table 4, the input data to
the track preprocessor that generates the track file meshes of the space curves are simple
entries. These entries are provided at the points that define track segments. For a tangent
section, which can be many miles long, only entries at the beginning and end of the section
need to be provided because the section geometry remains constant. The same is for a
curve, which has a constant curvature. The input at the beginning and end of a segment
can be used to completely define the geometry of the segment. In addition to location s at
the beginning and end of a segment, three specific geometry inputs are provided: the hori-
zontal curvature, superelevation, and grade. This information is provided in Table 4, which
shows an example of the data used as input to the track preprocessor: the node number,
distance at which the node is located, horizontal curvature, superelevation, and grade at
points where the geometry changes. In North America, the distance at which the node is
located along the track centerline from a given origin is measured in feet. The horizontal
curvature is defined as the number of degrees traversed by a 100 ft track: that is, the num-
ber of degrees for a 100 ft-chord arc length that encompasses angle 𝜓, as shown in Figure 8.
As discussed before, the horizontal curvature can be used to determine angle 𝜓. For a tan-
gent track, the horizontal curvature is zero. The fourth column in Table 4 provides values
of superelevation h at the nodes in inches. These values are used to determine bank angle
𝜙. The last column in Table 4 shows the grade, which is used to define the climb between
two points and vertical-development angle 𝜃. A grade value of 5.89% that defines a track
rise of 413 ft over a distance of 7012 ft is considered large. Some railroad tracks in North
America have a 4.7% grade value, which is also considered large since the grade of rail track
rarely exceeds 1%. Very steep grades are not normally used in track design, to avoid wasting
energy in braking scenarios.
The values given in Table 4, which correspond to the track space curve shown in Figure 6,
can be used as examples to explain how simple track inputs relate to segment geometry.
Based on the definitions given in the preceding sections, the first segment is a tangent,
the second segment is a tangent-to-curve entry spiral, the third segment is a curve, the
fourth segment is a curve-to-curve spiral, the fifth segment is a curve, the sixth segment
is again a curve-to-curve spiral, the seventh segment is a curve, the eighth segment is a
curve-to-tangent exit spiral, and the ninth segment is a tangent.
158 Mathematical Foundation of Railroad Vehicle Systems

100

P1

RH

P2
ψ

Figure 4.8 Horizontal curvature.

Computing Angles and Coordinates Given entries like those in Table 4, the Euler
angles at a node can be determined and used to define nodal positions as previously
described. The value of the horizontal curvature is often given using the following
equation:
sin(𝜓∕2)
CH = (4.50)
50′
It is clear from this equation that Euler angle 𝜓 is given using the table entries as
𝜓 = 2sin−1 (50CH ). The values of superelevation h at the track nodes given in the fourth
column in Table 4 and the value of gage G can be used to define bank angle 𝜙 as
h
sin 𝜙 = (4.51)
d
where d, shown in Figure 9, can be determined using the gage values. Finally, the grade,
which describes the track climb and is defined in the fifth column of Table 4, can be used
to calculate vertical-development angle 𝜃 at the nodes. Because the grade is the number
of units the track rises vertically divided by the number of units traveled horizontally, the
grade can be used to define sin 𝜃.

w
G

ϕ Bank angle h

Horizontal plane

Figure 4.9 Bank angle.


Railroad Geometry 159

As previously explained in this chapter, once the values of Euler angles 𝜓, 𝜃, and 𝜙 are
determined using the input to the track preprocessor at a relatively small number of points,
these angles can be determined at an arbitrary number of points specified by the user using
the interpolation described previously in this chapter. Interpolation of angles as functions
of the arc length can also be used to determine the position coordinates of nodes specified by
the user using Eq. 39, which cannot be evaluated in a closed form. Therefore, these position
coordinates are evaluated using numerical integration methods, including the trapezoidal
and Simpson rules, and Gaussian quadrature (Atkinson 1978).

Lengths of the Right and Left Rails Figure 10 illustrates that for a curved track, the
length of the track space curve used in developing the entries in Table 4 is not the same as
the lengths of the space curves of the right and left rails. The differences between the lengths
of these three space curves can be significant and should not be ignored when computing
position coordinates or angles at the discrete points specified by the user. For this reason,
the lengths of the right and left rails must be properly calculated. Using Figure 10, one can
show that the change in the radius of curvature of the curve projected on the horizontal
plane can be written as

ΔRH = d cos 𝜙 (4.52)


R
Δ

Br
B
Bl
2
R/
Δ

Track centerline

H
R

Left rail space curve

Right rail space curve


Δψ

sl
s
sr
RH
RH
+
Δ

Al
R/
2

Ar

Figure 4.10 Right and left rail lengths.


160 Mathematical Foundation of Railroad Vehicle Systems

where d is defined using the gage shown in Figure 9. One can then write
dSj = (RH ± (ΔRH ∕2))d𝜓
= (RH ± (d∕2) cos 𝜙)d𝜓 = (1 ± CH (d∕2) cos 𝜙)dS, j = rr, lr (4.53)
In this equation, superscripts rr and lr refer, respectively, to the right and left rails; the
positive sign corresponds to the right rail, while the negative sign corresponds to the left
rail for a positive curvature angle 𝜓. Using the equation dS = (cos𝜃)ds, which relates the
projected arc length of the track space curve to the actual arc length, one can then write the
projected arc length of a segment starting at a point defined by s1 as
s
S(s) = S1 + cos 𝜃(s)ds (4.54)
∫s1
For the right and left rails, one can also write
dSj
dsj = = (1 ± CH (d∕2) cos 𝜙)ds, j = rr, lr (4.55)
cos 𝜃
This equation, which defines the difference between ds and dsj , j = rr, lr, can be used to
compute the change in the length Δlj , j = rr, lr, of the segments of the right and left rails
using the data of the track space curve as
s
1
Δlj = ± C d cos 𝜙ds, j = rr, lr (4.56)
2 ∫s1 H
This equation can be used to evaluate the lengths of the right and left rail segments using
the equation
lj = l ± Δlj , j = rr, lr (4.57)
By using this length-adjustment scheme, it is clear that the values of the position coordi-
nates and Euler angles at the corresponding nodes of the track, right rail, and left rail space
curves can vary significantly. This variation must be taken into account in order to develop
an accurate algorithm for predicting wheel/rail contact forces.

Track Data File Based on the simple track description given in this section, a track pre-
processor can be designed to create space curve meshes for the track centerline and right
and left rail space curves. The track preprocessor produces a track data file that contains, at
minimum, the location of the mesh track nodes and Euler angles at these nodes. As previ-
ously mentioned,
[ for node k on space curve ] j, the track file can have the information given
jk jk jk
in the array s r1 r2 r3 𝜓 𝜃 𝜙 , where j = t, rr, lr, and t, rr and lr refer, respec-
jk jk jk jk

tively, to the track, right rail, and left rail space curves. Euler angles completely define the
track geometry, as explained in this chapter. Nodal locations can be defined in the track
body coordinate system or the rail body coordinate system. The latter case allows modeling
motion scenarios in which the rails move independently and are treated as separate bod-
ies during nonlinear simulations. Without any loss of generality, the track and rail body
coordinate systems can be assumed to coincide initially.
Railroad Geometry 161

Two approaches were discussed in this chapter for evaluating tangent, normal, and curva-
ture vectors. In both approaches, the global position vector of an arbitrary point on the rail is
defined using Eq. 16 as rr = Rr + Ar ur , where the superscript r stands for rail (right or left),
Rr is the position vector of the origin of the selected rail coordinate system X r Y r Z r in the
global coordinate system XYZ, Ar is the transformation matrix that defines the orientation
of the rail coordinate system, and ur is the location of an arbitrary point with respect to the
rail coordinate system. Vector Rr and matrix Ar do not depend on rail surface parameters.
In the first approach, referred to as the semi-analytical approach, used to determine the tan-
gent, normal, and curvature vectors, ANCF interpolation is not used, and the location of an
arbitrary point on the rail is defined using Eq. 11 as ur (sr1 , sr2 ) = Rrp (sr1 ) + Arp (sr1 )urp (sr1 , sr2 ).
Using this approach, evaluating the tangent, normal, and curvature vectors requires differ-
entiating vector Rrp (sr1 ) and matrix Arp (sr1 ) with respect to rail surface parameters; this is
in addition to differentiating vector urp (sr1 , sr2 ). Using such an approach requires evaluating
higher derivatives of Euler angles, including third derivatives, which are required when
solving wheel/rail contact conditions. Therefore, if this semi-analytical approach is used,
two additional sets of information must be provided at the nodal points for each space
curve: (i) the longitudinal tangent and its first and second derivatives with respect to the
curve arc length; and (ii) the first, second, and third derivative of Euler angles with respect
to the curve arc length. Clearly, using the semi-analytical approach can lead to more exten-
sive track-preprocessor and main-solver implementations that require a very large track
file and many more arithmetic operations, which can have an adverse effect on the com-
putational efficiency of the algorithm. The accuracy of the algorithm is also compromised
when using the semi-analytical approach because only low-order interpolation for position
coordinates can be used during nonlinear dynamic simulations.
Implementations can be significantly simplified, and the size of the track file can be
reduced, by using ANCF interpolation [ as previously discussed in ]this chapter. In this case,
only the data defined by the vector sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk needs to be included in
the track data file. This is mainly due to the fact that for ANCF interpolation, vector ur
is simply defined by Eq. 24 as ur = Sr er , where in this case, coordinate vector er includes
position gradients evaluated using the values of Euler angles at discrete nodal points. In
this case, differentiation to obtain the tangent and normal vectors is carried out by differ-
entiating shape function matrix Sr with respect to the surface parameters. Therefore, when
using ANCF interpolation, there is no need to differentiate Euler angles with respect to the
surface parameters of the rails; this can lead to significant simplification of the implemen-
tation in both the track preprocessor and the main solver and a significant reduction in the
size of the track data file produced by the track preprocessor.

Example 4.1
For the track described in Table 4, determine the location and orientation of the profile
frame at the nodes of the track centerline, assuming that the track has a gage value equal
to 1.42 m. Neglect the effect of the head width (Shabana et al. 2008).

(Continued)
162 Mathematical Foundation of Railroad Vehicle Systems

Solution For an arbitrary node i, one can write


( )
0.0254hi
𝜙 = sin
i −1
rad, i = 1, 2, … , 10
G
where hi is the given superelevation in inches at node i and G is the gage. Angle 𝜓 i can
be written as
𝜋cir
𝜓i = rad, i = 1, 2, … , 10
180
where cir is the given curvature in degrees at the node. Angle 𝜃 i can be defined as
( i )
GR
𝜃 i = sin−1 rad, i = 1, 2, … , 10
100
where GiR is the given grade percentage. The horizontal and vertical curvatures are given,
respectively, by
sin(𝜓 i ∕2) 𝜃i − 𝜃i−1
i
CH = 1∕m, Cv i = 1∕m, i = 1, 2, … , 10
50 × 0.3048 si − si−1
Since the grade is equal to zero for all nodes, one has
i i−1
CH + CH
Si = Si + (si − si−1 ) cos 𝜃 i , 𝜓 i = 𝜓 i−1 + (Si − Si−1 ), i = 1, 2, … , 10
2
By integrating Eq. 55 numerically, the locations and orientations of the profile frame at
each node can be determined and are summarized in the following table (Shabana et al.
2008):

Node # Distance s (m) C H (m−1 ) 𝜽 (rad) 𝝓 (rad) 𝝍 (rad) x (m) y (m) z (m)

1 0 0.0 0.0 0.0 0.0 0.0 0.0 0.0


2 30.48 0.0 0.0 0.0 0.0 30.48 0.0 0.0
3 45.72 0.00286 0.0 0.02932 0.02181 45.72 0.11 0.0
4 137.16 0.00286 0.0 0.02932 0.28353 135.84 13.98 0.0
5 152.40 −0.00172 0.0 −0.01953 0.29225 150.43 18.39 0.0
6 198.12 −0.00172 0.0 −0.01953 0.21372 194.68 29.83 0.0
7 219.46 0.00401 0.0 0.03909 0.23813 215.52 34.40 0.0
8 310.90 0.00401 0.0 0.03909 0.60442 298.50 71.58 0.0
9 349.0 0.0 0.0 0.0 0.68073 328.70 94.79 0.0
10 364.24 0.0 0.0 0.0 0.68073 340.54 104.38 0.0

4.10 IRREGULARITIES AND MEASURED TRACK DATA


Track preprocessors are designed to allow for including track deviations and using mea-
sured track data to perform more realistic virtual tests to identify the causes of derailments.
Before discussing the implementation of the analytical geometry variations in the track
Railroad Geometry 163

preprocessor and using measured track data, this section first discusses track quality
and classes.

Track Classes Railroad vehicle speed limits are dictated by track quality. Higher track
quality is required by federal regulators for passenger trains that travel at higher speeds in
order to ensure safety and avoid deadly accidents. Having higher track quality for passenger
routes is also necessary for ride comfort and to control the noise level. For high-speed trains,
track quality must be even higher to avoid deadly accidents. Track conditions can rapidly
deteriorate, particularly for freight trains, which have high axle loads. Therefore, tracks
that are owned by railroad companies in North America are regularly monitored to improve
their condition. The track maintenance cost for a single railroad company in North America
is measured in billions of dollars annually. The Federal Railroad Administration (FRA) in
the United States sets a limit on maximum train speed based on the quality of the track
sections. The FRA track-quality classification, shown in Table 5, is based on the maximum
allowable deviations. The data presented in this table, which show both speed and gage
limits, are for tangent segments of the track. Train speeds can be significantly reduced when
a train negotiates a curve or approaches a rail crossing.

Table 4.5 Speed and gage limits for different track classes.

Minimum Maximum
Track class Freight trains Passenger trains gage (in) gage (in)

Class 1 16 km/h ≈ 10 mile/h 24 km/h ≈ 15 mile/h 56 58


Class 2 40 km/h ≈ 25 mile/h 48 km/h ≈ 30 mile/h 56 57.75
Class 3 64 km/h ≈ 40 mile/h 97 km/h ≈ 60 mile/h 56 57.75
Class 4 97 km/h ≈ 60 mile/h 129 km/h ≈ 80 mile/h 56 57.50
Class 5 129 km/h ≈ 80 mile/h 145 km/h ≈ 90 mile/h 56 57.50
Class 6 177 km/h ≈ 110 mile/h 177 km/h ≈ 110 mile/h 56 57.25
Class 7 201 km/h ≈ 125 mile/h 201 km/h ≈ 125 mile/h 56 57.25
Class 8 257 km/h ≈ 160 mile/h 257 km/h ≈ 160 mile/h 56 57.25
Class 9 354 km/h ≈ 220 mile/h 354 km/h ≈ 220 mile/h 56.25 57.25

The data in Table 5 demonstrate that, in general, lower classes that correspond to lower
track quality are used for freight trains, which generally operate at lower speeds compared
to passenger trains. Nonetheless, passenger trains, in some cases, can run on lower-class
tracks, and freight trains are allowed in some cases to use higher-class tracks such as classes
6–9 if certain conditions are met (US Dept. of Transportation, Federal Railroad Administra-
tion, Office of Safety 2004, 2005). While Table 5 shows the gage limits for different classes,
the actual distance between the right and left rails can vary from original gage values due
to track irregularities, as discussed in this section.
Because track conditions can vary based on usage, measured track data are often used
in computer simulations of railroad vehicle systems. The measured data account for the
effect of all irregularities and can be used to better evaluate the response of a vehicle under
more realistic operating conditions. In addition to using measured data, standard analytical
164 Mathematical Foundation of Railroad Vehicle Systems

track deviations are often used in computer simulations to evaluate vehicle stability using
virtual prototyping simulation tests. Recorded track measurements can be used to develop
analytical geometry variation functions (Hamid et al. 1983). These analytical track devi-
ations, which take different shapes and can be used to represent both vertical and lateral
irregularities, can be systematically integrated with track data to test different scenarios and
develop guidelines for avoiding derailments. For example, new vehicle designs are virtually
analyzed and evaluated using simulation test scenarios that include standard deviations.
These tests are designed to excite certain modes that can influence vehicle stability and
cause derailments. In the remainder of this section, track deviations and measured track
data are discussed.

Analytical Geometry-Variation Functions Rail tracks are constructed using specific


geometries including tangent, curve, and spiral. Due to dynamic loads resulting from
wheel/rail interaction forces as well as environmental conditions that may include sig-
nificant temperature variations, the track geometry can change from its original design.
Changes in track geometry, referred to as track deviations, can include local geometry
variations, gage widening, rail buckling, cant change, etc. These geometric variations,
which are considered in vehicle design and when developing operation guidelines, can
lead to serious accidents or derailments; therefore, their effect needs to be evaluated.
Analytical track deviations that define specific smooth track geometry variations can
be systematically implemented in a track preprocessor. Recorded track measurements can
also be used to define other functions using interpolation schemes. Track deviations that
describe vertical or lateral geometry variations are known in a closed form. Vertical geom-
etry variations are called profile deviations, and lateral geometry variations are called align-
ment deviations. These two types of deviations are illustrated in Figure 11. Analytical profile
and alignment deviations, which can have arbitrary forms, are described using functions as
shown in Table 6. These functions, which include cusp, bump, jog, plateau, trough, sinusoid,
and damped sinusoid, are defined based on common irregularities observed in measured
track data (Hamid et al, 1983).

Alignment
Chord length

Profile (vertical deviation)


Z
X

X
Chord length

Y
(a) (b)

Figure 4.11 Track irregularities.


Railroad Geometry 165

Table 4.6 Examples of analytical track deviations. Source: Courtesy of Hamid, A., Rasmussen, K.,
Baluja, M., and Yang, T-L.

Deviation type Deviation shape Function f d (sr ) Occurrence

Cusp Y(s) y = Ae−k|s| Joints, turnouts,


interlocking, sun kinks,
buffer rails, insulated
A joints in continuous
welded rail (CWR), splice
bar joint in CWR, piers at
1/k s bridges
2
Bump Y(s) y = Ae−(1∕2)(ks) Soft spots, washouts, mud
spots, fouled ballast, joists,
spirals, grade crossings,
A bridges, overpasses, loose
bolts, turnouts,
1/k s interlinking
Aks
Jog Y(s) y= √ Spirals, bridges, crossings,
(1 + 4k2 s2 ) interlocking, fill-cut
transitions
0.5A
s

1/k
Plateau Y(s) √( )
Bridges, grade crossings,
A2 areas of spot maintenance
y=
1 + (ks)8
A 1/k

s
Trough Y(s) √ Soft spots, soft and
y = Ak [(1∕k)2 − s2 ] unstable subgrades, spirals
1/k
s
A

Sinusoid Y(s) y = A sin 𝜋ks Spirals, soft spots, bridges

1/k s

Damped Y(s) y = Ae−ks cos 𝜋ks Spirals, turnouts, localized


sinusoid soft spots

A
s
1/k
166 Mathematical Foundation of Railroad Vehicle Systems

The functions presented in Table 6 can be used as vertical (profile) or lateral (alignment)
deviations. Some of the analytical deviations shown in Table 6 are defined using exponen-
tial functions that assume a large value as arc length s increases. These functions can be
used to cover a large section of a deflected rail that assumes the shape of this particular
deviation. Hamid et al. (1983) provided a range for the values of parameters A and k used in
Table 6; these values are presented in Table 7. Table 6 also shows a summary of the possible
occurrences for each deviation. As discussed in the literature, a single cusp occurs as a pro-
file deviation due to the joint between two welded rail sections. Bump deviations, which are
smoother and typically cover longer rail segments, can be alignment or profile deviations
and can be found on both the right and left rails simultaneously. Bump deviations may be
considered profile irregularities in the case of bridges or alignment irregularities in the case
of spirals, where they occur at a certain distance from the start of the irregularities. A jog
can occur as the result of track stiffness variations, as in an interface between soft track
and a bridge. A plateau is generally caused by variations in track stiffness or wear of the
high rail. A cusp/plateau combination, which can be found before a spiral section of track
in some cases, can be dangerous because it leads to a sudden change in the wheel load.
Trough irregularities can be the result of poor drainage or localized soft subgrades. Sinu-
soid irregularities occur for bridges or reverse curves connected without a tangent segment
to separate them. Damped sinusoidal irregularities, which usually occur on a single rail,
are the result of significant changes in track stiffness due to switches, grade crossings, and
curves (Shabana et al. 2008).
Track irregularities, which can have adverse geometric effects, are taken into consider-
ation when developing operation guidelines. For example, Table 5 shows the gage limits
for different track classes. Nonetheless, lateral deviations can alter the distance between
the right and left rails; and in some cases, as the result of changes in the load across the
ties, the track cross-level changes. This cross-level change, called warp (twist), is defined
more accurately as the rate of change in the cross-level. To ensure safety, the gage change
for track classes 6 and higher, which are used for higher-speed trains, is limited to a maxi-
mum value of 0.5 in. within 31 ft. Due to safety requirements, limits are also imposed on the
maximum allowable vertical and lateral deviations for track classes used for higher-speed
trains. Tables 8 and 9 show the maximum allowable alignment and profile deviations for
track classes 6 and higher, respectively. In these tables, h1 , h2 , and h3 are, respectively, the
maximum allowable deviations from the designed geometry at the middle of 31-, 62-, and
124-ft chords.

Updating Track Preprocessor Data A track preprocessor can be designed to account


systematically for deviation functions that are used to modify the rail space curve geometry
regardless of the method used in the main solver (semi-analytical approach or ANCF inter-
polation) to describe the track geometry. In the track preprocessor, the function f d (sr ) is used
to define the geometry of the space curve in the region covered by a deviation. For a profile
deviation, the space curve over the domain of the deviation function is updated by vector
[ ]T
function fd = 0 0 fd (sr ) ; and for an alignment deviation, the space curve over the domain
[ ]T
of the deviation function is updated by vector function fd = 0 fd (sr ) 0 . Vector function fd
jk jk jk
can be used to define the nodal coordinates r1 , r2 , and r3 in the track-preprocessor output
Table 4.7 Parameters of track deviations. Source: Courtesy of Hamid, A., Rasmussen, K., Baluja, M., and Yang, T-L.

Parameter range
Gage Alignment Cross level Profile
−1 −1 −1
Deviation A (in) k (ft ) A (in) k (ft ) A (in) k (ft ) A (in) k (ft−1 )

Cusp 0.8–1.4 0.016–0.061 0.5–0.3 0.011–0.103 0.9–3.0 0.031–0.095 0.9–3.0 0.016–0.095


Bump 0.8–1.4 0.031–0.040 0.5–2.8 0.009–0.083 1.0–3.0 0.017–0.031 0.5–4.0 0.013–0.065
Jog – – 0.5–3.3 0.006–0.025 1.6–2.8 0.020–0.05 0.5–5.0 0.008–0.045
Plateau 0.8–1.3 0.029–0.08 1.2–1.6 0.025–0.027 0.6–1.0 0.026–0.04 0.9–3.0 0.009–0.033
Trough – – 1.4–2.2 0.013–0.029 – – 0.7–2.0 0.020–0.025
Sinusoid – – 0.8–1.2 0.033–0.020 – – 1.0–1.5 0.020–0.025
Damped sinusoid 0.5–1.0 – 1.0–2.2 0.013–0.015 0.9–1.2 0.051–0.061 – –
168 Mathematical Foundation of Railroad Vehicle Systems

[ ]
array sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk for a given node k of space curve j, j = rr, lr. Euler
angles 𝜓 jk and 𝜃 jk can also be adjusted due to lateral and vertical deviations, respectively.
The adjustment can be made easily by updating the angles, since the change in rotation due
to the deviation is assumed to be about a single axis. The change in the angle can be calcu-
lated using the equation 𝜓 jk = 𝜓 jk + Δ𝜓 jk = 𝜓 jk + tan−1 (df (sr )/dsr ) for lateral (alignment)
deviation and 𝜃 jk = 𝜃 jk + Δ𝜃 jk = 𝜃 jk + tan−1 (df (sr )/dsr ) for vertical (profile) deviation.

Table 4.8 Alignment limits. Source: Courtesy of Shabana, A.A., Zaazaa, K.E., and Sugiyama, H.

Three or more
Single deviation non-overlapping deviations

Track class h1 (in.) h2 (in.) h3 (in.) h1 (in.) h2 (in.) h3 (in.)

6 0.5 0.75 1.5 0.375 0.5 1.0


7 0.5 0.5 1.25 0.375 0.375 0.875
8 0.5 0.5 0.75 0.375 0.375 0.5
9 0.5 0.5 0.75 0.375 0.375 0.5

Table 4.9 Profile limits. Source: Courtesy of Shabana, A.A., Zaazaa, K.E., and Sugiyama, H.

Three or more
Single deviation non-overlapping deviations

Track class h1 (in.) h2 (in.) h3 (in.) h1 (in.) h2 (in.) h3 (in.)

6 1.0 1.0 1.75 0.75 0.75 1.25


7 1.0 1.0 1.50 0.75 0.75 1.0
8 0.75 1.0 1.25 0.50 0.75 0.875
9 0.5 0.75 1.25 0.375 0.50 0.875

Measured Track Data In virtual prototyping, analysis, and design of railroad vehicle
systems, it is important to examine vehicle nonlinear dynamics and stability when using
actual track data obtained from field measurements. This is particularly important during
accident investigations and when developing operation guidelines. The track preprocessor
must be designed to allow for reading and using the measured track data when constructing
the track geometry data used by the main solver. In accident investigations, in particular, it is
important to use the measured track data when performing dynamic simulations to develop
more realistic and credible computer models that can identify the causes of accidents. While
measured track data can be provided in various formats, these data must include the coordi-
nates of points along the track; and the gage, curvature, and superelevation at these points.
In addition to these data, information about the right and left rail profile and alignment
deviations is needed. Clearly, measured track data have the same information required by
the track preprocessor to define the track geometry, except that for measured track data,
Railroad Geometry 169

information is required at a much larger number of points in order to capture all the irregu-
larities recorded by the measurements. Given measured data, a simple computer procedure
can be developed to define standard inputs to the track preprocessor based on the data.
Using the measured track input data, which provide the same information as for analytical
track models, the track[ preprocessor can be designed ] to produce the required nodal data
jk jk jk
defined by the vector s r1 r2 r3 𝜓 𝜃 𝜙 , as previously discussed.
jk jk jk jk

When measured data are used, the curvature, grade, and superelevation are provided at
measurement points. Because the value of the gage varies and is defined at measurement
points, one can assume a mean gage value and use the measured gage value at the measure-
ment points to define alignment deviations that account for gage variations. Gage variation
involves both the right and left rails, so alignment deviations designed to represent the gage
variation can be equally divided between the two rails. Another important issue is using
smoothing techniques to filter out noise from measured track data, which is often provided
as raw data. For the gage and superelevation, filtering can be accomplished using a moving
average window with a width that depends on the data format.

4.11 COMPARISON OF THE SEMI-ANALYTICAL AND


ANCF APPROACHES

As explained in this chapter, a track preprocessor employs linear interpolation to calculate


Euler angles that describe the shape of the rail and the track space curve. Using numerical
integration, the position coordinates at nodes selected by the user can be determined.
Each node k of space curve j in the output of the track preprocessor is defined using
its
[ arc length location, position ] coordinates, and Euler angles given in the node array
jk jk jk
sjk r1 r2 r3 𝜓 jk 𝜃 jk 𝜙jk . During dynamic simulations, contact conditions are used
to define the location of points of contact between the wheels and rails. These contact
points can be any points inside the rail segments defined by the track preprocessor.
Therefore, it is necessary to use an accurate interpolation scheme to determine the
locations of the wheel/rail contact points between two nodes of a rail segment when the
surface parameters at these points become known from solving the wheel/rail contact
conditions.
This chapter discussed two geometric approaches for determining the locations of points
in a rail segment: semi-analytical and ANCF interpolation. In both approaches, the global
position of a point in a rail segment is defined using Eq. 16, given by rr = Rr + Ar ur ,
where superscript r stands here for rail (right or left), Rr is the position vector of the
origin of the selected rail coordinate system X r Y r Z r in the global coordinate system
XYZ, Ar is the transformation matrix that defines the orientation of the rail coordinate
system, and ur = ur (sr1 , sr2 ) is the location of an arbitrary point with respect to the rail
coordinate system. As previously mentioned, vector Rr and matrix Ar do not depend
on the rail surface parameters; therefore, they do not enter into the differentiation with
respect to the surface parameters when the tangent, normal, and curvature vectors are
evaluated. The semi-analytical and ANCF interpolation approaches differ in the way
vector ur = ur (sr1 , sr2 ) is defined. The definitions of ur = ur (sr1 , sr2 ) in the semi-analytical
170 Mathematical Foundation of Railroad Vehicle Systems

approach have two major computational disadvantages: the derivatives of Euler angles
with respect to rail surface parameters are required; and the degree of interpolation used
for the position [coordinates is limited to linear
] interpolation when the nodal information
given in vector sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk is used, as explained in this section. These
major computational disadvantages are avoided by using ANCF interpolation, which does
not require differentiation of Euler angles and also allows using a higher interpolation
order [(third-order) for the position ]coordinates based on nodal information given in the
array sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk .

Semi-Analytical Approach In the semi-analytical approach used to determine the tan-


gent, normal, and curvature vectors during dynamic simulations based on data produced
by a track preprocessor, ANCF interpolation is not used, and the location of an arbitrary
j j j j j j
point on the rail is defined using Eq. 11, given as uj (s1 , s2 ) = Rjp (s1 ) + Ajp (s1 )ujp (s1 , s2 ),
where j can be used for the right or left rail. It is clear from this equation that evaluating the
j
tangent, normal, and curvature vectors requires differentiating vector Rjp (s1 ) and matrix
j
Ajp (s1 ) with respect to the rail surface parameters, in addition to differentiating vector
[ ]T
j j j j
ujp (s1 , s2 ). In general, one can write ujp (s1 , s2 ) = 0 sj2 f (sj1 , sj2 ) . As previously shown
in this chapter, using this approach leads to the following definitions of the tangent and
j j j j j j
normal vectors, respectively: t1 = 𝜕rj ∕𝜕s1 = Aj (𝜕uj ∕𝜕s1 ), t2 = 𝜕rj ∕𝜕s2 = Aj (𝜕uj ∕𝜕s2 ), and
j j j jp j jp j
n = t1 × t2 , which require differentiating both R (s1 ) and matrix A (s1 ) with respect to the
rail surface parameters. Calculating the curvature vectors, as previously discussed, requires
evaluating higher derivatives. To determine the position coordinates [ of a point inside a]
rail segment defined by nodes k and (k + 1) using the data s r1 r2 r3jk 𝜓 jk 𝜃 jk 𝜙jk
jk jk jk

for node k of rail j, only linear interpolation can be used for the components of vector
j j j jk
uj (s1 , s2 ) since for a coordinate ul , l = 1, 2, 3 of vector uj , only two coordinates rl and
j(k+1)
rl are available. Therefore, with the information available from the track preprocessor,
j jk j(k+1) j
one is limited to the linear interpolation rl = (1 − 𝜉)rl + 𝜉rl , where 𝜉 = s1 ∕lk,(k+1) and
lk, (k + 1) is the length of the segment defined by the two nodes k and k + 1. Using linear
interpolation, which is not accurate for curved rails, fails to capture gradient and curvature
continuity and can lead to incorrect force predictions if a very fine mesh is not used. A
similar comment applies to the interpolation of Euler angles 𝜓 j , 𝜃 j , and 𝜙j , which can
only be interpolated linearly because only two coordinates 𝛼 jk and 𝛼 j(k + 1) , 𝛼 = 𝜓, 𝜃, 𝜙, are
available at the segment ends for each of the angles. Therefore, one is limited to linear
interpolation 𝛼 j = (1 − 𝜉)𝛼 jk + 𝜉𝛼 j(k + 1) . Furthermore, due to the non-commutativity of finite
rotations, and because interpolations imply addition, using the semi-analytical approach
is not recommended.
The fact that evaluating the tangent, normal, and curvature vectors requires differenti-
ation of Euler angles with respect to the arc length parameter when the semi-analytical
approach is used is the result of the dependence of the geometric description on matrix
j j j
Ajp (s1 ). The derivative of matrix Ajp (s1 ) with respect to the rail longitudinal arc length s1
can be written as
j j j j
𝜕Ajp ∕𝜕s1 = (𝜕Ajp ∕𝜕𝜓 j )(𝜕𝜓 j ∕𝜕s1 ) + (𝜕Ajp ∕𝜕𝜃 j )(𝜕𝜃 j ∕𝜕s1 ) + (𝜕Ajp ∕𝜕𝜙j )(𝜕𝜙j ∕𝜕s1 )
(4.58)
Railroad Geometry 171

Sj(k+1) j
Equation 37 shows that for space curve j, 𝜓 j (Sj ) = 𝜓 jk + ∫Sjk (CH (Sj ) )dSj , which can be
used with the second equation in Eq. 34 to evaluate 𝜕𝜓 j /𝜕sj as
j
𝜕𝜓 j ∕𝜕sj = (𝜕𝜓 j ∕𝜕Sj )(𝜕Sj ∕𝜕sj ) = CH cos 𝜃 j (4.59)
j
In this equation, CH at an arbitrary point can be evaluated using the linear interpo-
j jk j(k+1)
lation of Eq. 40 given by CH = (1 − 𝜉S )CH + 𝜉S CH . Similarly, Eq. 42 can be used to
j j
evaluate 𝜕𝜃 j /𝜕sj as 𝜕𝜃 j ∕𝜕sj = CV , where vertical curvature CV is assumed to be constant
determined using Eq. 45. Equation 47 defines the linear interpolation of angle 𝜙 as
𝜙j (Sj ) = (1 − 𝜉 S )𝜙jk + 𝜉 S 𝜙j(k + 1) , which shows that
1
𝜕𝜙j ∕𝜕sj = (𝜕𝜙j ∕𝜕Sj )(𝜕Sj ∕𝜕sj ) = (𝜙j(k+1) − 𝜙jk ) cos 𝜃 j (4.60)
Lk,k+1
S

where 𝜉S = Sj ∕Lk,k+1
S and Lk,k+1
S = Sj(k+1) − Sjk . Following a similar procedure, higher
angle derivatives, which are required to evaluate the curvature vectors, can be developed
[using the track preprocessor
] output data defined at the nodes by entries in the array
jk jk jk
sjk r1 r2 r3 𝜓 jk 𝜃 jk 𝜙jk . The angles and their derivatives are presented in Table 10.
Interpolating the angles, and evaluating and interpolating their derivatives, can be avoided
by using the second approach based on ANCF interpolation.

Table 4.10 Euler angles and their derivatives.

Angle First derivative Second derivative Third derivative

𝜕2 𝜓 j 𝜕3 𝜓 j
= =
𝜕(sj )2 𝜕(sj )3
𝜓 j (Sj ) = 𝜓 jk 𝜕𝜓 j j j
j
= CH cos 𝜃 j 𝜕CH 𝜕C j
Sj(k+1) j
+ ∫Sjk (CH (Sj ) )dSj 𝜕sj cos2 𝜃 j −2 Hj CV sin 𝜃 j cos 𝜃 j
𝜕Sj 𝜕S
j j j j
− CH CV sin 𝜃 j − CH (CV )2 cos 𝜃 j

𝜃 j (sj ) = 𝜃 jk 𝜕𝜃 j j 𝜕2 𝜃j 𝜕3 𝜃j
= CV =0 =0
j
+ CV (sj −s )jk 𝜕sj (𝜕sj )2 (𝜕sj )3

𝜕𝜙j 𝜕 2 𝜙j 𝜕 3 𝜙j
= = =
𝜕sj (𝜕sj )2 (𝜕sj )3
𝜙j (Sj ) = (1 − 𝜉S )𝜙jk Δ𝜙 Δ𝜙 j Δ𝜙
cos 𝜃 j , − k,k+1 CV sin 𝜃 j ,
j
− k,k+1 (CV )2 cos 𝜃 j ,
+ 𝜉S 𝜙j(k+1) LSk,k+1 LS LS
Δ𝜙 = 𝜙j(k+1) − 𝜙jk Δ𝜙 = 𝜙j(k+1) − 𝜙jk Δ𝜙 = 𝜙j(k+1) − 𝜙jk

ANCF Interpolation Approach With ANCF interpolation during dynamic simula-


tions performed by the main processor, a simple procedure can be used to obtain the local
rail geometry required for wheel/rail contact analysis. In this ANCF procedure, the deriva-
tives of angles are not required, and higher-order interpolation can be used for position
coordinates. Recall that the primary function of Euler angles in a track preprocessor is to
172 Mathematical Foundation of Railroad Vehicle Systems

define geometry. Therefore, these angles function primarily as elements of the position
vector gradients that can be used to shape a rail segment. By changing the orientation of
the position vector gradients, different geometry can be created. The main idea behind
ANCF interpolation is to use angles to develop the elements of position vector gradients at
nodal points. During dynamic simulations, wheel/rail contact conditions are used to solve
for wheel and rail surface parameters at potential contact points. Using the value of a rail
longitudinal surface parameter, the rail segment within which the potential contact point
lies can be determined. The two nodes of rail segments can be considered the two nodes of
an ANCF beam [ element. One can, therefore, ] use the elements of the track-preprocessor
output array sjk r1jk r2jk r3jk 𝜓 jk 𝜃 jk 𝜙jk to define, for each node of an ANCF ele-
ment, 12 coordinates: 3 position coordinates and 9 gradient coordinates. The gradient
coordinates can be defined using the columns of the transformation matrix of Eq. 20.
Therefore, for node k, one can define the following vector of nodal coordinates of the ANCF
element:
jk jk
⎡ e1 ⎤ ⎡ r1 ⎤ ⎡ jk
r1 ⎤
⎢ jk ⎥ ⎢ jk ⎥ ⎢ ⎥
⎢ e2 ⎥ ⎢ r2 ⎥ ⎢ jk
r2 ⎥
⎢ jk ⎥ ⎢ jk ⎥ ⎢ ⎥
e r
⎢ 3⎥ ⎢ 3⎥ ⎢ jk
r3
⎢ jk ⎥ ⎢ jk ⎥ ⎢ ⎥
⎢ e4 ⎥ ⎢rx1 ⎥ ⎢ cos 𝜓 jk cos 𝜃 jk ⎥
⎢ ejk ⎥ ⎢r jk ⎥ ⎢ ⎥
⎢ 5 ⎥ ⎢ x2 ⎥ ⎢ sin 𝜓 jk cos 𝜃 jk ⎥
⎢ ejk ⎥ ⎢r jk ⎥ ⎢ ⎥
sin 𝜃 jk ⎥
e = ⎢ jk ⎥ = ⎢ jk ⎥ = ⎢
jk 6 x3
⎢ e ⎥ ⎢ry1 ⎥ ⎢− sin 𝜓 jk cos 𝜙jk + cos 𝜓 jk sin 𝜃 jk sin 𝜙jk ⎥⎥
(4.61)
⎢ ⎥ ⎢ ⎥ ⎢
7
⎢ ejk ⎥ ⎢r jk ⎥ ⎢ cos 𝜓 jk cos 𝜙jk + sin 𝜓 jk sin 𝜃 jk sin 𝜙jk ⎥⎥
⎢ 8jk ⎥ ⎢ y2 ⎥ ⎢ ⎥
⎢ e9 ⎥ ⎢ry3jk
⎥ − cos 𝜃 jk sin 𝜙jk
⎢ jk ⎥ ⎢ jk ⎥ ⎢⎢ ⎥
jk sin 𝜙jk − cos 𝜓 jk sin 𝜃 jk cos 𝜙jk ⎥
⎢e10 ⎥ ⎢rz1 ⎥ ⎢ − sin 𝜓
⎢ jk ⎥ ⎢ jk ⎥ ⎢ cos 𝜓 jk sin 𝜙jk − sin 𝜓 jk sin 𝜃 jk cos 𝜙jk ⎥⎥
⎢e11 ⎥ ⎢rz2 ⎥ ⎢ ⎥
⎢ jk ⎥ ⎢ jk ⎥ ⎣ cos 𝜃 jk cos 𝜙jk ⎦
⎣e12 ⎦ ⎣rz3 ⎦
[ ]T
jk jk jk jk jk
In this equation, the gradient vectors at node k are rx = rx1 rx2 rx3 , ry =
[ ]T [ ]T
jk jk jk jk jk jk jk
ry1 ry2 ry3 , and rz = rz1 rz2 rz3 . Having 12 coordinates at the nodes of the ANCF
element allows using cubic interpolation for the position coordinates using the ANCF dis-
placement field of Eq. 24, given for rail segment l by the equation ujl (x, y, z) = Sjl (x, y, z)ejl ,
where vector ejl has 24 elements that have the coordinates of the 2 nodes determined
using the track preprocessor output data. The tangent, normal, and curvature vectors can
be determined by simply differentiating the element displacement field. As previously
mentioned, on the surface of the rail, element spatial coordinate z is written in terms
of the other spatial coordinates x and y. Using this ANCF approach eliminates the need
for the derivatives of the angles; instead, the angles at the nodes are used to determine
Railroad Geometry 173

the position gradients that define the element geometry. Furthermore, using this ANCF
approach allows for using a cubic interpolation of the position coordinates instead of using
less accurate linear interpolation. This approach for the integration of computer-aided
design and analysis (I-CAD-A) was adopted in 2000 for developing a general procedure for
the virtual prototyping of railroad vehicle systems. Such an approach can be systematically
generalized for the analysis of flexible rails, as has been demonstrated in the literature.
175

Chapter 5

CONTACT PROBLEM

One of the most fundamental problems in the study of railroad vehicle system dynamics
and stability is formulating wheel/rail contact forces. These forces must be accurately eval-
uated in order to obtain credible results that shed light on system behavior. When a wheel
is pressed against the rail, a contact region, referred to as a contact area or contact patch,
is formed. The shape of the region of contact between two solids depends on many factors
that include the material properties, contact pressure, solid geometries in the contact area,
etc. If the contact region covers a finite area that cannot be approximated by a point or a
line, one has the case of conformal contact. In the case of wheel/rail contact, the dimensions
of the contact area are small compared to the dimensions of the wheel and rail; therefore,
a localized concentrated contact is often assumed when wheel/rail contact forces are eval-
uated. This assumption of non-conformal contact can be justified because of the shapes of
the wheel and rail surfaces in the contact area.

Creep Forces Wheel/rail contact forces are the result of a combination of relative sliding
and rolling. As the vehicle attains a certain speed and due to the effect of friction forces, the
motion of the wheel with respect to the rail becomes predominantly rolling with a small
amount of slipping that gives rise to tangential creep forces as well as spin moments that have
a significant effect on vehicle dynamics and stability. In the case of traction and braking,
for example, the relative velocity between the wheel and rail at the contact point increases,
causing significant sliding, a case known as full saturation in which the tangential forces can
be approximated using the Coulomb’s law of friction. Below a certain relative velocity value,
slipping as a result of the creep phenomenon produces creep forces that can be expressed
in terms of normalized velocities called creepages. The creep phenomenon, attributed to
the elasticity of the two solids in the contact area, is the source of tangential creep forces,
which are functions of creepages. When two solids come into contact and are subjected to
external pressure, some points on the contact surfaces may slip while other points may stick;
therefore, the contact region consists of slip and adhesion areas. The small relative slip and
spin between the two solids can be the result of the difference between the deformations in
the contact region, and they lead to creep forces and spin moments, respectively. Different
linear and nonlinear wheel/rail tangential contact force and spin moment models expressed

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
176 Mathematical Foundation of Railroad Vehicle Systems

in terms of velocity creepages are used in the railroad vehicle system literature, as discussed
in this chapter.

Constraint and Elastic Contact Formulations Predicting the locations of the


contact points online is necessary for the generality of the wheel/rail dynamic algorithm
and the accuracy of predicting the wheel/rail interaction forces and moments. Funda-
mentally different approaches are used in formulating the wheel/rail dynamic interaction.
In some of these approaches, referred to as constraint contact formulations (CCFs), the
wheel is assumed to remain in contact with the rail: that is, wheel/rail separation is not
allowed. In some other formulations, referred to in this book as elastic contact formulations
(ECFs), wheel/rail separation is allowed. Both formulations can be used to determine the
normal contact force, which is the wheel/rail interaction force in the direction normal
to the surfaces of contact. To develop efficient computational algorithms for predicting
wheel/rail contact point locations and forces, it is often assumed that wheel/rail contact
is non-conformal: that is, the contact is assumed to cover a very small region such that
using a point contact can be justified. Once the normal contact force is determined, the
tangential creep forces can be computed using linear or nonlinear creep force models, as
discussed in this chapter.

Hertz Contact Theory As previously mentioned, in the case of non-conformal con-


tact, the wheel/rail contact area is assumed small in comparison with the wheel and rail
dimensions. When the constraint contact formulation is used, the normal contact force
is determined as a reaction force. In the elastic contact formulation, on the other hand,
the normal contact force is determined using a compliant force model with assumed stiff-
ness and damping coefficients. Experimental observations have shown that the wheel/rail
contact area can be approximated using an elliptical shape. The dimensions of the con-
tact ellipse, which enter into formulating the wheel/rail creep forces, can be determined
using Hertz contact theory (Hertz 1882). The contact ellipse dimensions, wheel and rail
material properties, creepages, and normal contact force predicted using the constraint or
elastic contact formulation can be used to formulate the tangential creep force and spin
moment. In Hertz contact theory, the geometry of two surfaces in contact is used to deter-
mine the principal curvatures, which are used to determine the dimensions of the contact
ellipse.

Maglev Trains In this chapter, Hertz contact theory, wheel/rail contact formulations,
and creep force models are discussed. While most of this chapter is devoted to formulat-
ing the wheel/rail contact problem, discussions of the forces used in magnetically levitated
(maglev) trains are also presented. In the case of maglev trains, in which there is no con-
tact between the vehicle and the guideway, vehicles are lifted using magnetic forces. This
allows for a significant increase in train speed due to the elimination of restrictions that are
the result of wheel/rail and pantograph/catenary contacts. Nonetheless, maglev train tech-
nology is still being researched and developed and for the most part, with a few exceptions,
Contact Problem 177

remains in an experimental phase. For this reason, passenger maglev trains are not widely
used for passenger transportation, and their use for freight transportation remains in doubt
for the foreseeable future.

5.1 WHEEL/RAIL CONTACT MECHANISM


Efficient dynamic simulations and virtual prototyping of complex railroad vehicle systems
require making simplifying assumptions in the formulation and computer implementation
of wheel/rail contact models. The simplifying assumptions often made are reasonable, are
based on acceptable scientific explanation, and lead to accurate prediction of wheel/rail
interaction forces. In this section, some of these assumptions are discussed and the basic
steps used for computing wheel/rail interaction forces are presented.

Conformal and Non-Conformal Contact Wheel/rail dynamic interaction is one of


the most important problems in modeling the nonlinear dynamic behavior of railroad vehi-
cle systems. As mentioned in the introduction of this chapter, different models can be used
to describe the interaction between two solids in contact. The contact between two solids
can be concentrated in a small area, such as in the case of two stiff spheres pressed against
each other. If the dimensions of the contact region are small compared to the dimensions
of the two solids in contact, the case of concentrated contact, referred to as non-conformal,
can be assumed. In the case of non-conformal contact, the interaction between the two
solids in the contact region can be described using a concentrated force vector and pos-
sibly a concentrated moment vector that is defined at the center of the contact area. The
calculations of the force and moment vectors can still take into account the dimensions
of the contact region, effect of normal pressure, material properties, and surface geometry
of the two solids in contact. Therefore, using the assumptions of non-conformal contact
does not imply that the dimensions of the contact region are totally ignored in comput-
ing wheel/rail contact forces. Using this assumption can be viewed as an approximation
in which distributed contact forces are replaced by concentrated force and moment vec-
tors. An example of this non-conformal contact approach is Hertz theory (1882), in which
the dimensions of the contact region are taken into consideration. Hertz theory has been
widely used to solve contact problems in a wide class of engineering applications, and its
assumptions have been tested and accepted by the scientific community in a large num-
ber of applications in which these simplifying assumptions can be justified. In addition to
numerical verification, experimentation has also been conducted to validate the results of
Hertz contact theory.
If, on the other hand, the contact region cannot be assumed small in comparison with
the dimensions of the two solids in contact, as in the case of a block pressed against a metal
sheet or a cylinder pressed against a solid plate, then the assumptions of non-conformal
contact cannot be justified. In these situations, one has the case of conformal contact. In
the case of conformal contact, interaction between the two solids in the contact region
178 Mathematical Foundation of Railroad Vehicle Systems

cannot be accurately described using force and moment vectors concentrated at a point.
Accurately predicting the forces of interaction between the two solids requires using a
continuum-based approach, such as the finite element (FE) method. The FE approach
allows for the discretization of the contact region to create point meshes that make it
possible to develop a general procedure for detecting the contact points between the
two solids. The FE approach is also better suited for modeling semi-conformal contact in
which only one dimension of the contact region cannot be assumed small compared to the
dimensions of the two solids in contact.
Clearly, using the continuum-based FE approach to model the contact between two solids
is more general and employs fewer simplifying assumptions. However, using this approach
can be computationally prohibitive, and in some applications, such an FE approach may
not lead to significant improvements in model accuracy, particularly in most railroad vehi-
cle system applications. When a wheel is pressed against a rail in most railroad applications,
the contact area is very small, and experimental observations have shown that the contact
area can be approximated by an elliptical shape that has very small dimensions compared to
the dimensions of the wheel and rail. Furthermore, numerical experimentation has shown
that using assumptions of non-conformal contact with partially nonlinear tangential creep
contact formulations leads to accurate results when compared with formulations based on
fully nonlinear approaches. For this reason, assumptions of non-conformal contact can
be justified in most cases and have been widely used in railroad vehicle dynamics algo-
rithms. Nonetheless, there are cases, particularly near-flange contact or worn wheels and
rails, where contact is conformal. In these cases, using non-conformal contact assumptions
may not lead to accurately predicting wheel/rail interaction forces. Efficient and accurate
conformal contact formulations for wheel/rail interaction forces are still being researched
in order to test the feasibility of using these approaches in the analysis of complex railroad
vehicle systems.

Slip, Adhesion, and Coulomb’s Friction When two solids come into contact, some
particles on the surfaces of the two solids slide relative to each other while others stick,
forming slip and adhesion regions. Therefore, the contact area can be covered by slip and
adhesion areas, which are difficult to determine precisely. Due to this difficulty, several the-
ories have been proposed, and different shapes and areas of slip and adhesion regions have
been assumed. Part of the slip region can be due to relative deformation between particles of
the two solids in the contact area. Pure sliding between two rigid surfaces is characterized
by jump discontinuity in tangential friction force as the transition from zero to non-zero
relative velocity occurs. Furthermore, tangential friction force, according to Coulomb’s the-
ory of dry friction (Coulomb 1785), assumes a constant value of 𝜇F n in the case of sliding,
where 𝜇 is the coefficient of friction and F n is the normal contact force. This tangential fric-
tion force, which has a direction opposite to the direction of the relative velocity between
the two bodies, has a value that is independent of the relative velocity between the two
bodies at the contact point when Coulomb’s theory of dry friction is used. While the accu-
racy of Coulomb’s friction law has been subjected to extensive debate for many years, it is
still widely used in many engineering applications to describe the friction between solids in
the case of finite relative velocities. This law, however, should not be used when the relative
Contact Problem 179

motion between two solids is predominantly due to deformations, as in the case of the creep
phenomenon.
If a local deformation is assumed in the small contact area, the change in tangential force
from a zero value at zero relative velocity to saturation value 𝜇F n at nonzero velocity can
be described using a smooth transition, and such a change is not characterized by the force
jump discontinuity that characterizes the force-velocity relationship in the case of contact
between two rigid surfaces. Some of the particles on the two surfaces deform with respect
to each other, leading to the creep phenomenon, and this gives rise to small slip that is not
accounted for in Coulomb’s theory of dry friction. As the relative motion between the two
solids evolves with time, the relative velocity between the two solids can be attributed to a
combination of small slip and rolling motion. Before the tangential friction force reaches
the saturation limit 𝜇F n , the relative velocity at the contact points is primarily due to the
creep phenomenon, which is a result of the relative deformation between particles on the
surfaces of the two solids. At these small values of the relative velocity in the contact area,
Coulomb’s friction law of the sliding between two rigid surfaces does not apply. To compute
the tangential forces resulting from normal contact pressure, normalized velocity coeffi-
cients called creepages are evaluated. Three normalized relative velocities have been found
to have a significant effect on wheel/rail interaction forces: longitudinal creepage 𝜁 x , lateral
creepage 𝜁 y , and spin creepage 𝜁 s . The equations of these normalized velocity creepages,
which enter into formulating the tangential creep force and spin moment, are presented
later in this chapter.

Steps in the Wheel/Rail Contact Analysis To determine wheel/rail contact forces,


several basic steps must be followed. The geometry and mechanics concepts discussed in
the preceding chapters play a fundamental role in all of these steps for developing a gen-
eral three-dimensional wheel/rail contact model. These steps, which need to be followed in
order and which demonstrate the importance of integrating geometry and mechanics con-
cepts, can be summarized as follows: (i) solve for the system configuration by numerical
integration of the nonlinear dynamic equations of motion to determine the system coor-
dinates and velocities; (ii) determine the location of the contact point on the wheel and
rail surfaces online using the coordinates of the wheel and rail obtained using numerical
integration of the equations of motion; (iii) compute the normal wheel/rail contact forces;
(iv) determine the geometry of the wheel and rail surfaces at the contact point; (v) calcu-
late the tangential creep or friction forces at the contact point; and (vi) use the normal and
tangential contact forces to determine the generalized contact forces associated with the
generalized coordinates of the wheel and rail. These six basic steps are described in more
detail as follows:
1. Solve for the system configuration. The system configuration at the beginning of the
simulation is assumed to be known from the initial coordinates and velocities. At any
other instant of time, the system coordinates and velocities are assumed to be known
from the numerical integration of the nonlinear equations of motion, developed in
[ T T ]T
Chapter 6. Therefore, for a given wheel w and rail r, coordinates qw = Rw 𝛉w and
[ r T r T ]T
r
q = R 𝛉 are assumed to be known when searching for the wheel/rail contact
180 Mathematical Foundation of Railroad Vehicle Systems

points, where Rk and 𝜽k , k = w, r, are, respectively, the vector that defines the global
position of the origin of the body coordinate system and the set of rotation parameters
that define the body orientation in the global coordinate system. Similarly, vectors
[ T T ]T [ T T ]T
q̇ w = Ṙ w 𝛉̇ w and q̇ r = Ṙ r 𝛉̇ r are assumed to be known from numerical
integration of the system equations of motion.
2. Determine the contact point. At this step, it is assumed that the configurations of the wheel
and rail are already determined from numerical integration of the system equations of
motion. That is, vectors of coordinates qw and qr of the wheel and rail, respectively, as
well as their time derivatives q̇ w and q̇ r , are assumed known. Using these known wheel
and rail configurations, two methods can be used to define the location of the contact
points on the wheel and rail online. These are the constraint and elastic contact formu-
lations, as previously mentioned in this chapter. As will be shown, the geometry and
mechanics concepts discussed in the preceding chapters represent the foundation for
the analysis performed in this step to determine the locations of the wheel/rail contact
points. The constraint contact formulation is described in Section 5.3, while the elastic
contact formulation is described in Section 5.4. Both formulations are based on a general
three-dimensional analysis and allow for predicting the location of the contact points
online. In both formulations, a set of algebraic equations is solved for the surface param-
eters that define the coordinates of the contact points on the wheel and rail surfaces.
In both approaches, the solution for the surface parameters requires using an iterative
Newton–Raphson algorithm because of the nonlinearity of the algebraic equations used.
The basic difference between the two contact formulations, however, is in the number of
degrees of freedom used. In the constraint contact formulation, wheel/rail separations
are not allowed, and therefore, the wheel has five degrees of freedom with respect to
the rail. In the elastic contact formulation, on the other hand, wheel/rail separations are
allowed, and therefore, the wheel has six degrees of freedom with respect to the rail. In
the constraint contact formulation, the wheel is not allowed to move with respect to the
rail in a direction normal to the contact surfaces at the contact point, while this rela-
tive motion is allowed in the elastic contact formulation. Some other and fundamentally
different approaches, such as lookup tables, are used in the literature to determine the
locations of the wheel/rail contact points. By using such simplified approaches, how-
ever, it is difficult to account for the three-dimensional nature of the wheel/rail contact
problem using a manageable set of precomputed data. Furthermore, using lookup tables
requires using interpolation methods that have orders that depend on the number of data
entries at each table point. A high degree of accuracy may require using very large data
files, particularly if a lower order of interpolation between the table points is the option to
be used. For three-dimensional contact analysis, the interpolation can be more demand-
ing computationally. Using a three-dimensional approach for predicting the location of
the contact points is necessary in many scenarios that include wheel climb at a large
angle of attack.
3. Compute the normal contact force. Having determined the location of the contact points
on the wheel and rail, the constraint or elastic contact formulation can be used to deter-
mine the normal contact force. In the constraint contact formulation, the normal contact
force is determined as a constraint (reaction) force because, in this CCF approach, one
degree of freedom is eliminated. As discussed later in this chapter, the local coordinates
of the contact point can be used to formulate the contact constraint conditions using a set
Contact Problem 181

of nonlinear algebraic equations. Chapter 6 explains how the contact constraints can be
used to determine the normal contact force systematically in terms of the constraint Jaco-
bian matrix and vector of Lagrange multipliers. If, on the other hand, the elastic contact
formulation is used, the wheel/rail penetration at the contact point is determined from
the known system configuration. The penetration and its time derivative are used to for-
mulate a compliant force model with assumed stiffness and damping coefficients. In the
ECF approach, wheel/rail separations are allowed; therefore, if there is no penetration,
the normal contact force is assumed to be zero.
4. Determine the surface geometry at the contact point. When two solids come into contact,
the dimensions of the contact area depend on the geometric properties of the surfaces
of the two solids in the contact region, their material properties, and the normal con-
tact force. Determining the dimensions of the contact area is necessary in order to be
able to evaluate the tangential creep forces. In Hertz contact theory, widely used in rail-
road vehicle dynamics, the contact area is assumed to be elliptical. To determine the
dimensions of the contact ellipse in Hertz theory, the principal curvatures and principal
directions of the wheel and rail surfaces at the contact point must be determined using
the approach discussed in Chapter 2. To determine the principal curvature and principal
directions, the geometry of the wheel and rail surfaces must be defined in terms of the
surface parameters and used to evaluate the coefficients of the first and second funda-
mental forms of the wheel and rail surfaces. These coefficients are used to determine
the principal curvatures and directions, as discussed in Chapter 2. It is explained in this
chapter how the principal curvatures enter into the definition of the dimensions of the
contact ellipse when Hertz theory is used.
5. Calculate the tangential creep forces. Tangential creep forces, which play a fundamental
role in railroad vehicle dynamics and stability, can be computed using the normal contact
force, dimensions of the contact ellipse, wheel and rail material properties, and relative
velocities between the wheel and rail at the contact point. If the coordinates of the con-
tact points on the wheel and rail are known in their respective body coordinate systems
and defined by vectors uPk , k = w, r, the global position vectors of the contact points can
be written as rkP = Rk + Ak uPk , k = w, r, where Ak , k = w, r is the transformation matrix
that defines the orientation of body k, as discussed in Chapter 3. Using this definition of
the global position vector, the absolute velocity vector of the contact points on the wheel
( )
and rail can be obtained, as discussed in Chapter 3, as vPk = Ṙ k + 𝛚k × Ak uPk , k = w, r,
where 𝛚k , k = w, r, is the absolute angular velocity vector of body k. Using these defi-
nitions of the absolute velocity vector and absolute angular velocity vector, the relative
velocity between the wheel and rail in the tangential plane as well as the relative angu-
lar velocity in the direction of the normal to the surfaces at the contact point can be
determined. These relative velocities are used to define the creepages that enter into the
definition of the creep force models. Linear and nonlinear creep force models have been
used in the literature. Some of these models are discussed in this chapter.
6. Evaluate the generalized forces. As discussed in Chapter 6, the equations of motion of
the wheel and rail can be formulated using Newton–Euler equations or the general-
ized form of those equations. Newton–Euler equations are written in terms of angu-
lar acceleration vectors, while their generalized form is written in terms of the second
time derivatives of the orientation parameters. Having determined the normal contact
forces, tangential creep forces, and spin moments at the contact points, either form of
182 Mathematical Foundation of Railroad Vehicle Systems

the Newton–Euler equations can be used to formulate the nonlinear system equations
of motion. For railroad vehicle systems subjected to motion constraints, it is more con-
venient to use the generalized form of the Newton–Euler equations. In this case, the
generalized contact forces associated with the generalized coordinates of the wheel and
rail must be evaluated before numerically integrating the accelerations, as explained in
Chapter 6. In general, the generalized wheel/rail contact forces are highly nonlinear
functions of the wheel and rail coordinates. For this reason, a fully nonlinear approach
that does not employ small-rotation assumptions needs to be used in order to be able to
predict the dynamics of railroad vehicle systems accurately.

Summary of the Wheel and Rail Surface Equations The contact force and con-
straint formulations discussed in this chapter, which require an understanding of the geom-
etry and mechanics concepts discussed in the preceding chapters, will be used to formulate
the nonlinear dynamic equations of complex railroad vehicle systems developed in Chapter
6. For the three-dimensional analysis procedure developed in this book, the wheel surface is
parameterized using surface parameters sw1 and sw2 , while the surface parameters of the rail
are sr1 and sr2 . Therefore, the vector of the wheel and rail surface parameters can be written
[ ]T
as s = sw1 sw2 sr1 sr2 , where superscripts w and r refer, respectively, to the wheel and
rail. The locations of points on the wheel and rail in the respective body coordinate system
can be written using these surface parameters, respectively, as
( ) [ ( ) ( ) ( )]T ⎫
uw sw1 , sw2 = xw sw1 , sw2 yw sw1 , sw2 zw sw1 , sw2 ,⎪
( ) [ ( ) ( ) ( )] T ⎬ (5.1)
ur sr1 , sr2 = xr sr1 , sr2 yr sr1 , sr2 zr sr1 , sr2 ⎪

The tangents and normal to the surfaces at the contact point are defined in the body coor-
dinate system as
𝜕uk 𝜕uk
tk1 = , tk2 = , nk = tk1 × tk2 , k = w, r (5.2)
𝜕sk1 𝜕sk2
Equations 1 and 2 are used in the following sections to define the wheel/rail contact
conditions. It is important, however, to recall from the analysis presented in Chapter 4
that the preceding two equations are highly nonlinear functions of the surface parameters.
For example, in the case of the wheel, vector uw , which is defined in the wheel coordinate
( ) [ wp ]T
system X w Y w Z w can be written as uw sw1 , sw2 = Rwp + uwp , where Rwp = xowp ywp o zo
is the vector that defines the origin of the wheel profile frame X wp Y wp Z wp with
respect to the origin of the coordinate system X w Y w Z w of the wheel or wheelset,
[ ( ) ( ) ]T
uwp = gwp sw1 sin sw2 sw1 −gwp sw1 cos sw2 is the vector that defines the location of the
arbitrary point in the profile frame, and gwp is the profile function, which can be a highly
nonlinear function in the wheel lateral surface parameter sw1 . Therefore, vector uw is a highly
nonlinear function of the wheel surface parameters, as discussed in more detail in Chapter
4. Similarly, in the case of the rail, vector ur is a highly nonlinear function in the rail surface
( )
parameters. This vector was defined in Chapter 4 in its most general form as ur sr1 , sr2 =
( ) ( ) ( ) ( ) [ ( ) ( r ) rp ]T
Rrp sr1 + Arp sr1 urp sr1 , sr2 , where Rrp = Rrp sr1 = xorp sr1 yrp o s1 zo is the
vector that defines the origin of the rail profile frame X rp Y rp Z rp in the rail coordinate
( )
system X r Y r Z r , Arp = Arp sr1 is the matrix that defines the orientation of the profile frame
Contact Problem 183

[ ( )]T
X rp Y rp Z rp in the rail coordinate system X r Y r Z r , urp = 0 sr2 grp sr1 , sr2 is the vector
that defines the location of the arbitrary point in the profile frame X rp Y rp Z rp , and grp is the
rail profile function. As discussed in Chapter 4, the use of the ANCF interpolation leads to
significant simplifications of the rail kinematic equations.

5.2 CONSTRAINT CONTACT FORMULATION (CCF)


In the wheel/rail constraint contact formulation, the wheel is assumed to remain in con-
tact with the rail, and therefore, separation is not allowed. In this contact formulation, the
freedom of the wheel to translate with respect to the rail along the direction normal to
the surface at the contact point is eliminated by enforcing a set of nonlinear algebraic con-
straint equations at the position, velocity, and acceleration levels. In general, one constraint
equation is sufficient to eliminate one degree of freedom. However, in the case of wheel/rail
contact in which the contact point is not fixed and slides on the surfaces, the four surface
[ ]T
parameters s = sw1 sw2 sr1 sr2 are introduced in order to be able to determine the loca-
tion of contact points on the wheel and rail surfaces online. Since four additional parameters
are used, one must enforce five constraint equations; four of them can be used to solve for
[ ]T
surface parameters s = sw1 sw2 sr1 sr2 that define the location of the contact points given
by Eq. 1. For a wheel/rail contact, the five nonlinear algebraic constraint equations used in
the non-conformal constraint contact formulation can be written as
⎡ rwr
P ⎤
( ) ⎢ wT r ⎥
C qw , qr , s w , s r
= ⎢t1 n ⎥ = 𝟎 (5.3)
⎢ wT r ⎥
⎣t2 n ⎦
These equations, in which rwr P = rP − rP and rP = R + A uP , k = w, r, is the global posi-
w r k k k k

tion vector of contact point P on body k, ensure that there are no wheel/rail separations and
the tangent planes to the wheel and rail surfaces at contact point P are parallel. The first
three scalar equations rwr P = rP − rP = 𝟎 ensure that there is no separation, while the last
w r
w• r w• r
two equations t1 n = 0 and t2 n = 0 ensure the parallelism of the tangent planes for the
wheel and rail surfaces at the contact point: a necessary non-conformal contact condition.
Figure 1, which shows a wheel and a rail in contact at point P, illustrates the parallelism of
the two tangent planes at the contact point.
To better explain the meaning of the constraint conditions of Eq. 3, these equations can
be rewritten using another coordinate system as
( ) [( T ) ( rT wr ) ( rT wr ) ( wT r ) ( wT r )]T
C qw , qr , sw , sr = tr1 rwr
P t2 rP n rP t1 n t2 n = 𝟎 (5.4)

While Eqs. 3 and 4 are equivalent, the first three scalar equations in Eq. 4 show more
clearly that the longitudinal, lateral, and normal relative position of the contact point on
the wheel and rail must remain equal to zero. As discussed at the end of the preceding
section, the vectors that appear in Eqs. 3 and 4 are highly nonlinear functions of the surface
parameters; therefore, the constraint conditions in these two equations are highly nonlinear
functions in these surface parameters as well as the wheel and rail generalized coordinates.
For this reason, an analytical solution for these constraint equations cannot be obtained
184 Mathematical Foundation of Railroad Vehicle Systems

rc rc
Z X wc
X

rc
Y wc
P Y

wc
Z

Figure 5.1 Contact conditions.

in practical applications, and the equations must be solved numerically using an iterative
[ ]T
procedure. For example, to solve for the four surface parameters s = sw1 sw2 sr1 sr2 , one
can choose the first two equations and last two equations in Eq. 4 to form the vector of
[ T ]T
constraint equations Cd (qw , qr , sw , sr ) = tr1 rwr tw1 nr tw2 nr = 𝟎. If the gen-
T T T
P tr2 rwr
P
eralized coordinates of the wheel and rail qw and qr , respectively, are assumed to be known
from the numerical integration, these four nonlinear algebraic equations can be solved
using an iterative Newton–Raphson algorithm to determine the four surface parameters
that ensure there is no longitudinal and lateral shift between the contact points on the wheel
and rail and also ensure that the tangent planes to the wheel and rail surfaces at the contact
T
point are parallel. The remaining equation nr rwr P = 0 imposes a constraint on the gener-
alized coordinates to ensure that there is no relative displacement between the wheel and
rail along the normal to the contact surfaces. In this CCF approach, the constraint con-
ditions must be satisfied at the position, velocity, and acceleration levels, and the normal
contact force is considered a reaction force. The first and second derivatives of the constraint
equations with respect to time must be evaluated and used with the equations of motion
to solve for the system accelerations and constraint forces. Chapter 6 discusses computa-
tional algorithms for solving the constraint conditions of Eq. 4 with the dynamic equations
of motion to determine the surface parameters, system coordinates, and constraintforces.

5.3 ELASTIC CONTACT FORMULATION (ECF)


Another formulation that can be used to determine the locations of the wheel/rail contact
points online is the elastic contact formulation. Unlike the constraint contact formulation,
the elastic contact formulation does not eliminate any degrees of freedom and allows for
wheel/rail separations. Therefore, a rigid wheel is assumed to have six degrees of freedom
with respect to the rail. This is despite the fact that a set of algebraic equations similar to the
Contact Problem 185

ones used in the constraint contact formulation are used in the elastic contact formulation.
In the ECF approach, the algebraic equations used to solve for the surface parameters do not
impose any constraints on the motion of the wheel or wheelset, and no constraint (reaction)
forces are associated with these equations.
Using the algebraic contact conditions to determine the locations of the contact points
on the wheel and rail is not the only approach that can be used. Other alternate approaches
that can be used include nodal search methods and lookup tables (Shabana et al. 2008). In
the nodal search approach, point meshes are created on the wheel and rail surfaces, and
the distances between these points are checked to determine whether contact occurs. In
the lookup table approach, precomputed wheel and rail point data are tabulated and used
with an interpolation scheme to determine whether there is contact. For surfaces that have
a certain degree of smoothness, there is no clear advantage of using the nodal search and
lookup table approaches. Furthermore, implementing these two approaches can be cum-
bersome, particularly in the case of three-dimensional contacts that cannot be ignored in
important motion scenarios, including wheel climbs and derailments. Therefore, the focus
in this section is on using algebraic equations to determine the locations of the contact
points. This approach has proven to be robust and allows for the implementation of efficient
algorithms for the wheel/rail contact analysis.
In all the elastic contact formulations, such as using algebraic equations, nodal search,
and lookup tables, distances between points on the wheel and rail are checked along the
normal to the contact surfaces to determine whether penetration 𝛿 occurs. If there is pene-
tration, penetration 𝛿 and its time derivative are computed and used to compute the normal
contact forces using a compliant force model with assumed stiffness and damping coeffi-
cients.
In the elastic contact formulation discussed in this section, four algebraic equations
are solved for each wheel/rail contact to determine the four surface parameters
[ ]T ( ) ( )
s = sw1 sw2 sr1 sr2 that define the locations of contact points uwp sw1 , sw2 and urp sr1 , sr2
on the wheel and rail surfaces, respectively. It is assumed that the wheel and rail config-
urations are known from the integration of the system equations of motion. Because the
contact algebraic equations are nonlinear functions of the surface parameters, an iterative
Newton–Raphson algorithm is used to solve the contact conditions for the four surface
parameters. The four equations used in the ECF approach are the same as the first two and
last two equations of Eq. 4 used in the CCF approach. These equations are rewritten as
T
⎡g1 (sw , sr )⎤ ⎡tr1 rwr
P ⎤
⎢ ⎥ ⎢ rT wr ⎥
( ) ⎢g2 (s , s )⎥ ⎢t2 rP ⎥
w r
g sw , sr = ⎢ =⎢ T ⎥=𝟎
w r ⎥
(5.5)
⎢g3 (s , s )⎥ ⎢ t1 n ⎥
w r

⎢ ⎥ ⎢ T ⎥
⎣g4 (sw , sr )⎦ ⎣ tw2 nr ⎦
These four nonlinear algebraic equations can be solved for surface parameters
[ ]T [ ]T
sw = sw1 sw2 and sr = sr1 sr2 using an iterative Newton–Raphson algorithm. Toward
this end, the set of algebraic equations (𝜕g/𝜕s)Δs = − g are iteratively solved for each
[ T T ]T [ ]T
contact, where Δs = Δsw Δsr = Δsw1 Δsw2 Δsr1 Δsr2 is the vector of Newton
186 Mathematical Foundation of Railroad Vehicle Systems

differences and
𝜕g [ 𝜕g ] [ 𝜕g 𝜕g 𝜕g 𝜕g
]
𝜕g
= =
𝜕s 𝜕sw 𝜕sr 𝜕sw1 𝜕sw2 𝜕sr1 𝜕sr2

⎡ 𝜕g1 𝜕g1 𝜕g1 𝜕g1 ⎤


⎢ 𝜕sw1 𝜕sw2 𝜕sr1 𝜕sr2 ⎥
⎢ 𝜕g 𝜕g2 𝜕g2 𝜕g2 ⎥
⎢ w2 ⎥
⎢ 𝜕s 𝜕sw2 𝜕sr1 𝜕sr2 ⎥
=⎢ 1
𝜕g3 ⎥⎥
(5.6)
⎢ 𝜕g3 𝜕g3 𝜕g3
⎢ 𝜕sw1 𝜕sw2 𝜕sr1 𝜕sr2 ⎥
⎢ 𝜕g 𝜕g4 𝜕g4 𝜕g4 ⎥⎥
⎢ 4
⎣ 𝜕sw1 𝜕sw2 𝜕sr1 𝜕sr ⎦
2
This 4 × 4 Jacobian matrix can be written more explicitly as
(( )T ) (( )T )
⎡ 𝜕tr1 ∕𝜕sr2 rwr − tr1 tr2 ⎤
T T T T
tr1 tw1 tr1 tw2 𝜕tr1 ∕𝜕sr1 rwr − tr1 tr1
⎢ (( ) (( )⎥
⎢ )T )T
𝜕tr2 ∕𝜕sr2 rwr − tr2 tr2 ⎥⎥
T T T T
𝜕g ⎢ tr2 tw1 tr2 tw2 𝜕tr2 ∕𝜕sr1 rwr − tr2 tr1
=
𝜕s ⎢( w w )T r ( w w )T r ( r r )T w ( r r )T w ⎥
⎢ 𝜕t1 ∕𝜕s1 n 𝜕t1 ∕𝜕s2 n 𝜕n ∕𝜕s1 t1 𝜕n ∕𝜕s2 t1 ⎥
⎢( w w )T r ( w w )T r ( r r )T w ( r r )T w ⎥
⎣ 𝜕t2 ∕𝜕s1 n 𝜕t2 ∕𝜕s2 n 𝜕n ∕𝜕s1 t2 𝜕n ∕𝜕s2 t2 ⎦
(5.7)
In the case of non-conformal contact, this Jacobian matrix is nonsingular; therefore, the
system of equations (𝜕g/𝜕s)Δs = − g can be solved for the Newton differences, which can
be used to update the surface parameters. The iterative process continues until convergence
is achieved. Implementing this method showed that using the values of the surface param-
eters from the previous time step leads to a good initial guess for the Newton–Raphson
iterations and fast convergence. It is important to emphasize again that while the four non-
linear ECF algebraic equations used to solve for the surface parameters are the same as four
of the algebraic equations used in the constraint contact formulation, the ECF approach
cannot be considered a constraint formulation because the ECF algebraic equations are not
imposed at the velocity and acceleration levels and do not give rise to constraint forces.
Furthermore, the ECF procedure allows for wheel/rail penetrations and separations.
[ ]T
The converged solution obtained for surface parameters s = sw1 sw2 sr1 sr2 can
( ) ( )
be used to define the locations of potential contact points uwp sw1 , sw2 and urp sr1 , sr2 on
the wheel and rail surfaces, respectively. These local position vectors can be used to define
the global position vectors of the potential contact points as rkP = Rk + Ak uPk , k = w, r,
which can be used in turn to define vector rwr w r
P = rP − rP . The distance between the two
potential contact points on the wheel and rail surfaces along the normal can be evaluated
using an equation similar to the third equation of Eq. 4. Toward this end, the wheel/rail
penetration can be computed as
T ( )T r
𝛿 = rwr
P n ̂ r = rwP − rrP n ̂ (5.8)
(r )
where n r | r r |
̂ = t1 × t2 ∕ |t1 × t2 | is the unit normal to the rail surface at the potential contact
r

point. The sign of the penetration in the preceding equation can be used to define whether
there is wheel/rail contact or separation.
Contact Problem 187

5.4 NORMAL CONTACT FORCES

The fundamental differences between the CCF and ECF approaches will become clearer
from the development of the governing dynamic equations and computational procedures
presented in Chapter 6. While a CCF algorithm does not allow for wheel/rail separations
and is not well-suited for the analysis of derailments, such an algorithm can shed more
light on stability and critical speed issues without making assumptions in the force model.
For this reason, implementing the constraint contact formulation is recommended for the
study of motion scenarios that require precise calculations of some parameters under nor-
mal operating conditions and for investigating phenomena without making assumptions
related to the nature of the wheel/rail interaction. The ECF implementation, on the other
hand, is necessary for the analysis of motion scenarios in which the wheels are separated
from the rails. These scenarios include derailments, and for this reason, the ECF implemen-
tation is required for credible accident investigations.
In the constraint contact formulation, the non-conformal wheel/rail contact conditions
presented in this chapter can be used to identify a dependent coordinate and a constraint
force for each contact. As shown in Chapter 6, the constraint reaction force, which defines
the wheel/rail normal contact force, can be systematically determined using the Jacobian
matrix of one of the contact constraint equations, which is taken to be the third equation in
T ( w )
r T r
Eq. 4, rwr r
P n = rP − rP n = 0. Therefore, after the surface parameters and generalized
coordinates are determined from a position analysis, the constraint equations can be for-
mulated and used to define the constraint Jacobian matrix, which is used in the Lagrangian
approach to define the wheel/rail normal contact force F n . This normal contact force can
be used to compute the tangential creep forces and spin moment, as discussed later in this
chapter.
In the case of the elastic contact formulation, on the other hand, the normal contact
force can be calculated using a compliant force model if the wheel and rail surfaces
penetrate. This compliant force model is based on assumed stiffness and damping coeffi-
( )T r
̂ r = rwP − rrP n̂
T
cients. The penetration can be determined from the equation 𝛿 = rwr P n
(Eq. 8), and the relative velocity along the normal to the rail surface can be written
( )T r
as vrn = ṙ wr ̂ r = ṙ wP − ṙ rP n̂ with the understanding that the unit normal to the rail
T
P n
surface is assumed
( T ) fixed when evaluating the normal component of the relative velocity. In
wr ̂ r
general, d rP n ∕dt = ṙ P n ̂ r + rwr ̂̇ r ; therefore, vrn = ṙ wr ̂ r is the component of the
wr T T T
P n P n
relative velocity along the normal to the rail surface at the contact point and is not exactly
the time derivative of 𝛿, unless an assumption is made that the normal to the rail surface
is fixed because the generalized coordinates and surface parameters, as well as their time
derivatives, are assumed known. It is also clear that vector n ̂̇ r is perpendicular to vector
̂ ; therefore, the velocity component rP n
n r wr T
̇
̂ is not along a direction normal to the contact
r

surface. For this reason, the equation 𝛿̇ = vrn = ṙ wr ̂ r is often used to measure the time
T
P n
rate of the penetration along the normal to the contact surfaces. Using the expressions for
𝛿 and vrn , the normal contact force F n can be defined as the sum of an elastic force F ns and
a damping force F nd as (Shabana et al. 2004):

Fn = Fns + Fnd = −K(𝛿)1.5 − Cvrn |𝛿| (5.9)


188 Mathematical Foundation of Railroad Vehicle Systems

where F ns = −K𝛿 1.5 is the elastic contact force used in Hertz theory, K is the Hertzian
stiffness coefficient that depends on the material properties and the geometry of the two
surfaces in contact, F nd = −Cvrn |𝛿| is the component of the damping force, and C is an
assumed damping coefficient. The absolute value of penetration |𝛿| is included in the damp-
ing force to ensure that the normal contact force assumes a zero value when penetration
is zero.

5.5 CONTACT SURFACE GEOMETRY


One computational method that can be used to predict the locations of the contact point
to study wheel/rail interaction forces is to create point meshes on the wheel and rail sur-
faces. Using the generalized coordinates of the wheel and rail, which are assumed to be
known from the numerical integration of the equations of motion, the distances between
these mesh points along a direction normal to the surface can be computed. The surface dis-
cretization can be performed using the FE method or a set of arbitrary points selected by the
analyst. In this case, the forces can be calculated directly without the need to specify a con-
tact area or have a certain degree of smoothness of the two surfaces in contact. In the case of
non-smooth surfaces, an approximation of the normal to the contact surfaces can be made if
the differential geometry smoothness requirements cannot be applied. While this approach
is more general and allows for modeling both conformal and non-conformal contacts, its
use for describing the contact surface geometry can lead to inefficient implementation of
computational algorithms designed for solving railroad vehicle system problems. For this
reason, a different semi-analytical geometry approach has been used in the railroad vehicle
dynamics literature. This approach, which requires computing the principal curvatures and
principal directions of the wheel and rail surfaces at the contact point, has proven to capture
the nonlinear dynamics of the wheel/rail contact accurately. It also allows for defining the
dimensions of the wheel/rail contact area, which is assumed to be elliptical based on exper-
imental observations (Hertz 1882; Johnson 1985). In Hertz contact theory, non-conformal
contact is assumed. This assumption is justified in most railroad vehicle system applica-
tions due to the shapes of the solid wheel and rail. Because the dimensions of the wheel/rail
contact ellipse are small in comparison with the wheel and rail dimensions and the relative
radii of curvatures, the assumption of half-space is often used (Johnson 1985). This assump-
tion allows for treating contact stresses separately from the stress distributions in the wheel
and rail due to other forces. Hertz contact theory, widely used in railroad vehicle dynam-
ics, is based on static considerations, does not account for dynamic and friction effects, and
employs the assumptions of small deformations and linear elasticity.

Surface Parametric Forms Figure 2 shows two solids in contact. The surfaces of these
two solids, denoted as i and j, can be represented in parametric forms as previously men-
tioned in this chapter and as explained in more detail in Chapter 2. If two different coor-
dinate systems are used to define the two surfaces, one can write the equations of the two
Contact Problem 189

Body i

i
Z
i
X
Y
j P

i j
Y j
X
Z

Body j

Figure 5.2 Contact between two solids.

surfaces, respectively, as
⎡r1i ⎤ ⎡ xi ⎤
⎢ ⎥ ⎢ ⎥
ri = ⎢r2i ⎥ = ⎢ yi ⎥ (5.10)
⎢ i ⎥ ⎢ i ( i )2 ( i )2 ( i i) ⎥
⎣r3 ⎦ ⎣a1 x + a2 y + a3 x y + · · ·⎦
i i

and
j
⎡r1 ⎤ ⎡ xj ⎤
j ⎢ j ⎥ ⎢ j ⎥
r = ⎢r2 ⎥ = ⎢ y ⎥ (5.11)
⎢ j ⎥ ⎢ j ( j )2 ( )2 ( ) ⎥
⎣r3 ⎦ ⎣a1 x + aj2 yj + aj3 xj yj + · · ·⎦
In these equations, the surface parameters are assumed to be Cartesian coordinates: that is,
[ ]T [ ]T
sk = sk1 sk2 = xk yk , k = i, j. A polynomial with coefficients akl , l = 1, 2, … , nkp is used
to define the shapes of the two surfaces, where the number of coefficients nkp , k = i, j, can be
used to define the order of the polynomial. Because static and small deformation assump-
tions are made in Hertz theory – that is, no rigid body motion of the surfaces with respect
to their coordinate systems – the absence of the linear terms of the polynomials of the pre-
ceding equations can be justified. The two curve equations can also be defined in the same
coordinate system using one set of parameters and two different sets of coefficients that
define the different geometries of the two surfaces. Using this representation, and assum-
ing that the third axis of the coordinate system is normal to the contact surfaces as shown
in Figure 2, one can define the gap between the two surfaces in the neighborhood of the
j
origin of the coordinate system as hij = r3i − r3 , which can be written using the parametric
equations of the two surfaces defined in the same coordinate system as
hij = a1 (x)2 + a2 (y)2 + a3 (xy) + · · · (5.12)
where al , l = 1, 2, …, np , are the resulting polynomial coefficients and np is the number of
coefficients that can be used to define the polynomial order. The gap hij = hij (x, y) is used
to define the area of contact between the two solids.
190 Mathematical Foundation of Railroad Vehicle Systems

Surface Curvatures Because the contact area is assumed to be small, a quadratic


interpolation is sufficient for the surface representation of the two solids in the
contact region. Therefore, Eqs. 10 and 11 can be written, respectively, as rk =
[ ( ( ) ( )2 ( ))]T
2
xk yk ak1 xk + ak2 yk + ak3 xk yk , k = i, j. Furthermore, the coordinate
system can be selected such that coefficient ak3 is equal to zero. This leads to the surface
[ ( ( ) ( )2 )]T
2
equation rk = xk yk ak1 xk + ak2 yk , k = i, j. Using the coefficients of the first
and second fundamental forms of surfaces presented in Chapter 2, one can determine the
principal curvatures 𝜅1k and 𝜅2k , which define the maximum and minimum curvatures. For
[ ( ( ) ( )2 )]T
2
a surface described by the parametric equation rk = xk yk ak1 xk + ak2 yk , one
can show that this equation at the origin yields
( )2 }
𝜕 2 r3k ∕𝜕 xk = 2ak1 = 𝜅1k = 1∕Rk1
( )2 (5.13)
𝜕 2 r3k ∕𝜕 yk = 2ak2 = 𝜅2k = 1∕Rk2 , k = i, j

where Rk1 and Rk2 are the principal radii of curvature at the origin, which must be given the
appropriate signs depending on the signs of the principal curvatures, which can be positive
or negative. It is clear from Figure 2, for example, that one surface has positive or negative
curvatures, and the other surface has curvatures with opposite signs. Therefore, the surface
equation can be written in terms of the principal radii of curvatures for the two solids as

⎡ xk ⎤
⎡r1k ⎤ ⎢ ⎥
⎢ k⎥ ⎢ yk ⎥
( k )2 ⎥ ,
k
r = ⎢ r2 ⎥ = ⎢ ( ) k = i, j (5.14)
2
⎢ k⎥ ⎢ x k y ⎥
⎣ r3 ⎦ ⎢ + k ⎥
⎣ 2R k
1 2R 2 ⎦

Regardless of the coordinate system used, the principal curvatures and principal radii of cur-
vature are unique. Similarly, the gap function of Eq. 12 can be represented using a quadratic
function and can be written, by selecting a coordinate system such that a3 = 0, as

(x)2 (y)2
hij = a1 (x)2 + a2 (y)2 = + (5.15)
2R1 2R2
where R1 and R2 are the relative principal radii of curvature at the origin.

Mathematical Derivation Because the axes of the coordinate lines xi yi and xj yj are not,
in general, parallel, their orientations can differ by an angle 𝛼 ij , as shown in Figure 3. The
orientation of coordinates xi yi and xj yj also differ from coordinates xy by angles 𝛼 i and 𝛼 j ,
respectively. The convention used in Figure 3 for measuring the angles differs slightly from
the one used in the literature (Johnson 1985). Due to this difference, the formulas obtained
in this section can have sign differences when compared with the formulas presented in
some publications. Using the coordinate systems and angles shown in Figure 3, one can
write the following coordinate transformation:
[ ] [ ][ ]
xk cos 𝛼 k sin 𝛼 k x
= , k = i, j (5.16)
yk − sin 𝛼 cos 𝛼
k k y
Contact Problem 191

yi
yj

xj

αij

xi

Figure 5.3 Relative orientation of the coordinate system.

It follows that (xk )2 = (x)2 cos2 𝛼 k + 2xy sin 2𝛼 k + (y)2 sin2 𝛼 k and (yk )2 = (x)2 sin2 𝛼 k − 2xy
sin 2𝛼 k + (y)2 cos2 𝛼 k . Therefore, if the quadratic form of Eq. 12 is used, one has
j ( )2 ( )2 j ( )2 j ( )2
hij (x, y) = r3i − r3 = ai1 xi + ai2 yi − a1 xj − a2 yj
( ) ( )2
= ai1 (x)2 cos2 𝛼 i + xy sin 2𝛼 i + (y)2 sin2 𝛼 i + ai2 (x)2 sin2 𝛼 i − xy sin 2𝛼 i + (y)2 cos2 𝛼 i
j ( ) j( )2
−a1 (x)2 cos2 𝛼 j + xy sin 2𝛼 j + (y)2 sin2 𝛼 j − a2 (x)2 sin2 𝛼 j − xy sin 2𝛼 j + (y)2 cos2 𝛼 j
(5.17)
This equation can be written as
( )
j j
hij (x, y) = ai1 cos2 𝛼 i + ai2 sin2 𝛼 i − a1 cos2 𝛼 j − a2 sin2 𝛼 j (x)2
( )
j j
+ ai1 sin2 𝛼 i + ai2 cos2 𝛼 i − a1 sin2 𝛼 j − a2 cos2 𝛼 j (y)2
(( ) ( ) )
j j
+ ai1 − ai2 sin 2𝛼 i − a1 − a2 sin 2𝛼 j (xy) (5.18)

Using this equation, coefficients a1 , a2 , and a3 of Eq. 12 in the case of a quadratic polynomial
can be written as
a1 = ai1 cos2 𝛼 i + ai2 sin2 𝛼 i − a1 cos2 𝛼 j − a2 sin2 𝛼 j ⎫
j j
j j ⎪
a2 = a1 sin 𝛼 + a2 cos 𝛼 (− a1 sin )𝛼 − a2 cos2 𝛼 j ⎬
i 2 i i 2 i 2 j
(5.19)
( i ) j j
a3 = a1 − ai2 sin 2𝛼 i − a1 − a2 sin 2𝛼 j ⎪

Using Eq. 13, coefficient a3 can be written as
( ) ( )
1 1 1 i 1 1 1
a3 = − sin 2𝛼 − − sin 2𝛼 j (5.20)
2 Ri1 Ri2 2 Rj1 Rj2

The condition that a3 = 0, which is necessary to write Eq. 15, is then given by
( ) ( )
1 1 i 1 1
− sin 2𝛼 = − j sin 2𝛼 j (5.21)
Ri1 Ri2 j
R1 R2
It is also clear from Eq. 19 that
( ) ( )
a2 − a1 = − ai1 cos2 𝛼 i − sin2 𝛼 i − ai2 sin2 𝛼 i − cos2 𝛼 i
j ( ) j ( )
+ a1 cos2 𝛼 j − sin2 𝛼 j + a2 sin2 𝛼 j − cos2 𝛼 j
( ) ( )
j j
= − ai1 − ai2 cos 2𝛼 i + a1 − a2 cos 2𝛼 j (5.22)
192 Mathematical Foundation of Railroad Vehicle Systems

This equation can be written using Eqs. 13 and 21 as


( ) ( )
1 1 1 i 1 1 1
a2 − a1 = − − cos 2𝛼 + − cos 2𝛼 j
2 Ri1 Ri2 2 Rj1 Rj2

√(
√ )2 ( )2 ( )( )
1√ 1 1 1 1 1 1 1 1
= − + − j +2 − − j cos 2𝛼 ij
2 Ri1 Ri2 j
R1 R2 Ri1 Ri2 j
R1 R2
(5.23)
Using Eq. 19 and similar manipulations, one can show that a1 + a2 can be written as
( ) ( )
1 1 1 1 1 1
a1 + a2 = + − + (5.24)
2 Ri1 Ri2 2 Rj1 Rj2

Example 5.1 The parametric equation of a surface can be written in terms of two
[ ]T
parameters x and y as r = x y f (x, y) . The tangents to the surface at x and y can be
[ ]T [ ]T
written as rx = 𝜕r∕𝜕x = 1 0 𝜕f ∕𝜕x and ry = 𝜕r∕𝜕y = 0 1 𝜕f ∕𝜕y . The normal
[ ]T
to the surface is defined as n = rx × ry = −fx −fy 1 , where f x = 𝜕f /𝜕x and f y = 𝜕f /𝜕y.
Using the tangent vectors, the coefficients of the first fundamental form can be written
as
( )2 ( )2
E = rTx rx = 1 + fx , F = rTx ry = fx fy , G = rTy ry = 1 + fy
[ ]T
The curvature vectors are defined as rxx = 𝜕 2 r∕𝜕x2 = 0 0 fxx , ryy = 𝜕 2 r∕𝜕y2 =
[ ]T [ ] T
0 0 fyy , and rxy = 𝜕 2 r∕𝜕x𝜕y = 0 0 fxy , where f xx = 𝜕 2 f /𝜕x2 , f yy = 𝜕 2 f /𝜕y2 , and
f xy = 𝜕 f /𝜕x𝜕y. Using the curvature vectors, the coefficients of the second fundamental
2

form can be written as


f fxy fyy
̂ = xx , M = rTxy n
L = rTxx n ̂= ̂=
, N = rTyy n
|n| |n| |n|
where
√( ) n ̂ = n∕ |n| is a unit vector along the normal to the surface, and |n| =
2 ( )2
fx + fy + 1.
If function f is defined using the quadratic polynomial f = a1 (x)2 + a2 (y)2 , one has
fx = 2a1 x, fy = 2a2 y, fxx = 2a1 , fyy = 2a2 , fxy = 0,
√ ( )2 ( )2 √ ( )2 ( )2
and |n| = 1 + fx + fy = 1 + 2a1 x + 2a2 y . Using these derivatives, the
curvature vectors can be written as
[ ]T [ ]T
rxx = 0 0 2a1 , ryy = 0 0 2a2 , rxy = 𝟎
The coefficients of the first and second fundamental forms can be written in this case as
( )2 ( )2
E = 1 + 2a1 x , F = 4a1 a2 xy, G = 1 + 2a2 y
L = 2a1 ∕ |n| , M = 0, N = 2a2 ∕ |n|
Contact Problem 193

As discussed in Chapter 2, the principal curvatures can be defined using the character-
istic equation
| L − 𝜅n E M − 𝜅n F |
| |
| |=
|M − 𝜅 F N − 𝜅 G |
| n n |
( ) ( )2
EG − (F)2 𝜅n − (EN + GL − 2FM) 𝜅n + LN − (M)2 = 0
where 𝜅 n is the normal curvature. This characteristic equation has the following two
roots: √( )
2
−b ± b − 4a c
𝜅1,2 =
2a
where in this example
( )2 ( )2
a = EG − (F)2 = 1 + 2a1 x + 2a2 y = (|n|)2
[ ( ( )2 ) ( ( )2 )]
b = − (EN + GL − 2FM) = − 2a2 1 + 2a1 x + 2a1 1 + 2a2 y ∕ |n|
c = LN − (M)2 = 4a1 a2 ∕(|n|)2
At the origin, x = 0 and y = 0. In this case, the characteristic equation reduces to
(𝜅 n )2 − 2(a1 + a2 )𝜅 n + 4a1 a2 = 0, which can be written as (𝜅 n − 2a1 )(𝜅 n − 2a2 ) = 0. The
roots of this equation define the principal curvatures as 𝜅 1 = 2a1 and 𝜅 2 = 2a2 .

Contact Ellipse The analysis presented in this section shows that the gap hij between
the√two solids has the geometry of an ellipse. The ratio between the ellipse axes is defined
( )
as a1 ∕a2 . The actual size of the contact ellipse depends on the load applied to the two
solids. Due to the application of the load, the two surfaces can be displaced vertically with
respect to each other. Let ui and uj be, respectively, the displacements of the solids i and j,
as shown in Figure 4. In this figure, 𝛿 i and 𝛿 j are the deformation of the two points Pi and

Undeformed body i Fn

Deformed body i P
j

zj
δ δ
j uj
X-Y plane
i
δ ui
zi
Deformed body j P
i

Undeformed body j

Figure 5.4 Relative displacement between the solids in contact.


194 Mathematical Foundation of Railroad Vehicle Systems

Pj on two solids i and j, respectively. If two points Pi and Pj are assumed to coincide, one
can write the total deformation as 𝛿 = 𝛿 i + 𝛿 j = ui + uj + hij . Using the definition of the gap
function hij given by Eq. 15, this equation can be written as
ui + uj = 𝛿 − a1 (x)2 − a2 (y)2 (5.25)
This equation is used in the static linear-elastic force analysis in Hertz theory.

5.6 CONTACT ELLIPSE AND NORMAL CONTACT


FORCE
In both the constraint and elastic contact formulations, the dimensions of the contact ellipse
are determined in order to evaluate the tangential creep forces and spin moment. In addi-
tion to the dimensions of the contact ellipse, which depend on the contact pressure, a
compliant normal force model in the case of the elastic contact formulation can be devel-
oped based on Hertz theory, which provides an expression for the stiffness coefficient that
can be used to formulate the elastic force in such a compliant force model. As previously
mentioned, the assumptions of half-space are used because the contact region is assumed
small in comparison to the dimensions of the two solids. Furthermore, the displacement
and normal stresses as the result of the contact pressure are assumed negligible outside the
contact area, and the shear stresses 𝜏 xz and 𝜏 yz on the contact surfaces are assumed to be
zero (Johnson 1985; Goldsmith 1960).
Using these assumptions, the elastic component F ns of the applied normal force F n can
be considered equal to the restoring force resulting from the normal component of pressure
p in the contact region: that is, F ns = ∫ ∫ pdxdy. In Hertz theory, pressure p is assumed to
be a quadratic function of x and y, which can be written in the form
√ ( )2 ( )2
x y
p = p0 1 − − (5.26)
a b
where p0 is a constant pressure value, and a and b are the lengths of the semi axes of the
contact ellipse. Following the derivation presented in Appendix A, contact pressure p leads
to displacement u given by (Love 1944; Johnson 1985; Goldsmith 1960)
1 − 𝜈2 ( )
u= Le − Me (x)2 − Ne (y)2 (5.27)
𝜋E
In this equation, E is the modulus of elasticity, 𝜈 is the Poisson ratio, and
𝜋p0 ab ∞ dw 𝜋p0 b ( ⎫ )
Me = √( ) ( )
= Ke − E e⎪
2 ∫0 3 e2 (a)2
(a)2 + w (b)2 + w w ⎪
[( ) ]⎪
𝜋p0 ab ∞ dw 𝜋p0 b a 2
Ne = √( = E e − Ke ⎪
2 ∫0 ) ( )3 e2 (a)2 b ⎬ (5.28)
(a)2 + w (b)2 + w) w ⎪
𝜋p0 ab ∞ dw ⎪
Le =
2 ∫0
√( )( ) = 𝜋p0 bK e ⎪
(a) + w (b)2 + w w
2 ⎪


where Ee and K e are the elliptical integrals of argument e = 1 − (b∕a)2 , b < a. These
integrals are given in Appendix A. Equation 27 can be used to define the total displacement
Contact Problem 195

ui + uj as
( )
ui + uj = Le − Me (x)2 − Ne (y)2 ∕𝜋Eij (5.29)
where
( )2 ( )2
1 1 − 𝜈i 1 − 𝜈j
= + (5.30)
Eij Ei Ej
Using this equation and Eq. 25, one has
Le − Me (x)2 − Ne (y)2
𝛿 − a1 (x)2 − a2 (y)2 = (5.31)
𝜋Eij
It is clear from this equation and the definitions of Eq. 28 that
Le pb ⎫
𝛿= = 0ij Ke ⎪
𝜋Eij E ⎪
Me p0 b ( ) ⎪
a1 = = K − E ⎬ (5.32)
𝜋Eij Eij (e)2 (a)2
e e
[( ) ]⎪
N p0 b a 2 ⎪
a2 = eij = E e − Ke ⎪
𝜋E 2
Eij (e) (a) 2 b ⎭
It follows that
( ) ⎫
a2 R1 (a∕b)2 Ee − Ke ⎪
= =
a1 R2 Ke − E e ⎪
√ √( )⎬
(5.33)
√ 1 p b )( ⎪
a1 a2 = 12 = 0ij (a∕b)2
E e − K e K e − E e ⎪
R1 R2 E (a)2 (e)2

Because the pressure distribution is assumed semi-ellipsoidal, the elastic component of the
normal force F ns can be written as
2
Fns = pdxdy = p 𝜋ab (5.34)
∫∫ 3 0
This equation and Eq. 26 lead to (Hertz 1882; Love 1944; Goldsmith 1960)
√ ( )2 ( )2
3Fns x y
p= 1− + (5.35)
2𝜋 a b a b
In Hertz contact theory, friction or damping is not considered, and in this case, the total
force F n is equal to the elastic force F ns : that is F n = F ns . In this case, the contact ellipse
semi axes a and b are defined as
( ( ) )1∕3 }
a = m 3𝜋Fn K1 + K2 ∕4K3
( ( ) )1∕3 (5.36)
b = n 3𝜋Fn K1 + K2 ∕4K3
where K 1 and K 2 are constants that depend on the material properties of the two solids,
while K 3 is a constant that depends on the solid geometries. These constants are given as
( )2 ( )2 ⎫
1 − 𝜈i 1 − 𝜈j
K1 = , K = , ⎪
𝜋Ei 2
𝜋Ej ( ) ⎪
( ) ⎬ (5.37)
( ) 1 1 1 1 1 1 ⎪
K3 = a 1 + a 2 = + − +
2 Ri1 Ri2 2 Rj1 Rj2 ⎪

196 Mathematical Foundation of Railroad Vehicle Systems

Table 5.1 Coefficients m and n in terms of angle 𝜃 defined in degrees. Source: Courtesy of Hertz, H.

𝜽 m n 𝜽 m n 𝜽 m n

0.5 61.4 0.1018 10 6.604 0.3112 60 1.486 0.717


1 36.89 0.1314 20 3.813 1.4123 65 1.378 0.759
1.5 27.48 0.1522 30 2.731 0.493 70 1.284 0.802
2 22.26 0.1691 35 2.397 0.530 75 1.202 0.846
3 16.5 0.1964 40 2.136 0.567 80 1.128 0.893
4 13.31 0.2188 45 1.926 0.604 85 1.061 0.944
6 9.79 0.2552 50 1.754 0.641 90 1.0 1.0
8 7.86 0.285 55 1.611 0.678

Table 1 shows the coefficients m and n in Eq. 36 as functions of the angle 𝜃 (0 ∘ ≤ 𝜃 ≤ 90∘ )
(Hertz 1882), where 𝜃 is defined as
( )
𝜃 = cos−1 K4 ∕K3 (5.38)
and

√( )2 ( )2 ( )( )
( ) 1√ √ 1 1 1 1 1 1 1 1
K4 = a 2 − a 1 = − + − j +2 − − j cos 2𝛼 ij
2 Ri2 Ri1 j
R2 R1 Ri2 Ri1 j
R2 R1
(5.39)
As previously explained in this chapter, in the constraint and elastic contact formulations,
a set of algebraic equations is solved at the position analysis step to determine the surface
parameters of a potential contact point. Using these surface parameters, the principal curva-
tures and principal directions can be determined and used to compute the constants defined
in this section. An interpolation scheme can be employed to determine coefficients m and
n using the data presented in Table 1. These coefficients and constants are substituted in
Eq. 36 to determine the dimensions of the contact ellipse. An alternative to using interpo-
lation based on the data given in Table 1 is to obtain closed-form expressions for constants
m and n in terms of angle 𝜃. The closed-form expressions given here for coefficients m and
n were proposed by Berzeri (Shabana et al. 2001)
Bm ⎫
m = Am tan (𝜃 − 𝜋∕2) + + Dm , ⎪
(𝜃)Cm ⎪
⎬ (5.40)
1 ⎪
n= + Bn (𝜃)Cn + Dn sin 𝜃 ⎪
An tan (𝜃 − 𝜋∕2) + 1 ⎭
Contact Problem 197

Table 5.2 Coefficients in the closed-form expressions of m and n. Source:


Courtesy of Shabana, A.A., Zaazaa, K.E., and Sugiyama, H.

Coefficient Value Coefficient Value

Am −1.086 419 052 477 An −0.773 444 080 706


Bm −0.106 496 432 832 Bn 0.256 695 354 565
Cm 1.350 000 000 000 Cn 0.200 000 000 00
Dm 1.057 885 958 251 Dn −0.280 958 376 499

where, in this equation, the value of 𝜃 is given in radians and the coefficients Al , Bl , Cl , and
Dl , l = m, n, are defined in Table 2 (Shabana et al. 2001). The preceding equation results
in a good approximation of coefficients m and n and captures the asymptotic behavior of
coefficient m as 𝜃 approaches zero.

Compliant Force Model Using the first equation in Eq. 32 and the second equation in
Eq. 36, a compliant force model based on Hertz theory (Eq. 34) can be defined as

4cns
Fns = K(𝛿)1.5 = ( )√ (𝛿)1.5 (5.41)
3 K1 + K2 a1 + a2
( ( )√ )
where K = 4cns ∕ 3 K1 + K2 a1 + a2 is a stiffness coefficient, and cns is a constant given
in Table 3 (Goldsmith 1960). The compliant force F ns can be used to define the elastic com-
ponent of the normal force F n used in the elastic contact formulation.

Table 5.3 Coefficient cns . Source: Courtesy of Goldsmith, W.

a1 /a2 1.0 0.7041 0.4903 0.3333 0.2174 0.1325 0.0718 0.0311 0.00765

cns 0.3215 0.3180 0.3322 0.3505 0.3819 0.4300 0.5132 0.6662 1.1450

Example 5.2
In this example, a wheelset that consists of two wheels connected by an axle is assumed
to travel on a tangent track. At a given instant of time, the right wheel is assumed to be
in contact with a point on the rail that has a transverse radius of curvature of 0.26 m. The
wheel is assumed to have a conical profile with a rolling radius of 0.46 m at the instant of
time considered. The normal load applied to the right wheel is assumed to be 50 000 N.

(Continued)
198 Mathematical Foundation of Railroad Vehicle Systems

The wheel and rail are assumed to be made of steel with a modulus of elasticity equal to
2.1 × 1011 Pa and a Poisson ratio of 0.28. If the yaw angle of the wheelset is assumed to be
zero, determine the dimensions of the semi axes of the contact ellipse and the maximum
contact pressure.
Solution To determine the dimensions of the semi axes of the contact ellipse, the con-
stants K 1 , K 2 , and K 3 of Eq. 37 must be determined first. The superscripts w and r are
used in this example to refer to the wheel and rail, respectively. Because the wheel tread
is assumed to be conical, Rw1 = ∞; and because the rolling radius is given, Rw2 = 0.46 m.
For the tangent track assumed in this example, one has Rr1 = ∞ and Rr2 = −0.26 m. A
negative sign is used for Rr2 , to be consistent with the notations used in this book. There-
fore, Eq. 37 leads to
1 − (𝜈 w )2 1 − (0.28)2
K1 = = ( ) = 1.396 × 10−12 m2 ∕N,
𝜋Ew 𝜋 2.1 × 1011

1 − (𝜈 r )2 1 − (0.28)2
K2 = = ( ) = 1.396 × 10−12 m2 ∕N,
𝜋E r 𝜋 2.1 × 1011
( ) ( ) [ ]
1 1 1 1 1 1 1 1 1
K3 = + − + = 0 + − 0 −
2 Rw1 Rw2 2 Rr1 Rr2 2 0.46 (−0.26)

= 3.01 m−1
The coefficient K 4 can be evaluated using Eq. 39. Because in this example, the yaw angle
is assumed to be zero, the angle 𝛼 ij in Eq. 39 is zero; therefore,

( )2 ( )2 ( )( )
1 1 1 1 1 1 1 1 1
K4 = − + − + 2 − − cos 2𝛼 ij
2 Rw2 Rw1 Rr2 Rr1 Rw2 Rw1 Rr2 Rr1

( )2 ( )2 ( )( )
1 1 1 1 1
= −0 + −0 +2 −0 − 0 cos (0)
2 0.46 (−0.26) 0.46 (−0.26)

= 0.8361 m−1
Using constants K 3 and K 4 , the angle 𝜃 of Eq. 38 can be evaluated as
( )
𝜃 = cos−1 K4 ∕K3 = cos−1 (0.8361∕3.01) = 1.2893 rad
Using the data of Table 1, one can show that m and n can be approximated as m = 1.23
and n = 0.83. The closed-form expressions of Eq. 40 can also be used to determine the
values of m and n instead of the data of Table 1. Using the values of m and n and the
coefficients K 1 , K 2 , and K 3 ; the dimensions of the semi axes of the contact ellipse can be
evaluated as
[ ( ) ]1∕3
a = m 3𝜋Fn K1 + K2 ∕4K3 = 5.8805 × 10−3 m
[ ( ) ]1∕3
b = n 3𝜋Fn K1 + K2 ∕4K3 = 3.9681 × 10−3 m
Contact Problem 199

where in this equation, F n is assumed to be equal to the normal force given in this
example: that is, F n = 50 000 N. Recall that in Hertz’ theory, F n = F ns because this the-
( )
ory does not account for the energy dissipation in the contact area. Because 1∕Rw1 −
( ) ( ) ( )
1∕Rr1 < 1∕Rw2 − 1∕Rr2 , a and b are, respectively, the longitudinal and transverse
dimensions of the semi axes of the contact ellipse. Using the dimensions of the contact
ellipse, the contact ellipse area Ace can be evaluated as
Ace = 𝜋ab = 7.331 × 10−5 m2
The contact pressure at the origin, which represents the maximum contact pressure, can
be evaluated using Eq. 35 as
3 3
p0 = F = F = 102.3092 Mpa
2𝜋ab n 2Ace n
This simple example demonstrates the small dimensions of the contact ellipse and the
large value of the contact pressure for the given 50 000 normal force, material properties,
and assumed geometries of the wheel and rail surfaces.

5.7 CREEPAGE DEFINITIONS

The relative motion between two solids in contact can be a combination of rolling and slid-
ing. In the case of pure rolling, there is no relative sliding between the two points in the
contact region; therefore, there are no friction forces as a result of the relative displacement
between the two solids in the contact area. In the case of pure sliding, referred to as full sat-
uration, the dominant tangential contact force is the friction force, which is often evaluated
using Coulomb’s dry friction law in which the tangential friction force is written in terms
of the normal force multiplied by the coefficient of friction as 𝜇F n , where 𝜇 is the coeffi-
cient of friction and F n is the normal contact force. Predominantly sliding motion in the
case of railroad vehicle dynamics can occur during traction and braking scenarios. Under
normal operating conditions, the relative motion between the wheel and rail is predomi-
nantly rolling with some points on the wheel and rail sticking together or slipping relative
to each other, giving rise to creep forces. The small relative slip is the result of the relative
deformation displacement between points on the wheel and rail. Furthermore, the creep
phenomenon can result in a relative rotation about the normal to the contact surfaces. This
relative rotation, called spin, is produced by the relative deformation of the two surfaces in
the contact area and results in a creep spin moment that can be important in some scenar-
ios, including flange contacts. Both the slip and spin as a result of the creep phenomenon
have been found to play a significant role in the dynamics and stability of railroad vehicle
systems; therefore, in general, their effect cannot be ignored.
In railroad vehicle dynamics, creep forces are developed in terms of velocity parame-
ters called creepages (Carter 1926). Three components of the relative velocities are found
to be significant in the study of railroad vehicle system dynamics: the longitudinal and lat-
eral creepages 𝜁 x and 𝜁 y , respectively, and the spin creepage 𝜁 s . The longitudinal and lateral
200 Mathematical Foundation of Railroad Vehicle Systems

Z
ωi
Y

Figure 5.5 Tangential creep and friction forces.

creepages 𝜁 x and 𝜁 y , respectively, are dimensionless velocities defined by dividing the rel-
ative velocity between the wheel and rail at the contact point by the forward velocity of
the wheel center relative to the rail; this forward relative velocity is denoted as V. The spin
creepage 𝜁 s is defined by dividing the component of the relative angular velocity between
the wheel and rail along the normal to the contact surface by the relative velocity V: that
is, spin creepage is not dimensionless. The longitudinal and lateral creepages lead to tan-
gential creep forces, while the spin creepage leads to a spin moment. In railroad vehicle
system dynamics, slip is considered the result of the creep phenomenon if the relative veloc-
ity between the wheel and rail, shown in Figure 5, is smaller than a certain value vs . If the
relative velocity is higher than vs , the case of relative sliding is considered, and tangential dry
friction forces, defined by the equation 𝜇F n , are used instead of creep forces. The velocity
creepages are defined as
( w )T ( w )T
vP − vPr ̂tr1 vP − vPr ̂tr2 ̂r
(𝛚w − 𝛚r )T n
𝜁x = , 𝜁y = , 𝜁s = (5.42)
V V V
In this equation, vPw and vPr are, respectively, the absolute velocities of the contact points on
the wheel and rail; ̂tr1 , ̂tr2 , and n
̂ r are, respectively, the unit longitudinal and lateral tangents
and unit normal to the surface at the contact point; and 𝝎w and 𝝎r are the absolute angular
velocities of the wheel and rail, respectively. It is clear that in the case of pure rolling about
the wheel lateral axis only, all the creepage components are equal to zero. Nonetheless,
because the case of pure rolling cannot be sustained for a long duration due to the variations
in wheel displacements, creepages do not normally vanish, giving rise to creep forces and
moments.
Using one of the contact formulations discussed in this chapter, the global positions of
the contact points on the wheel and rail can be determined as rkP = Rk + Ak uPk , k = w, r,
as previously discussed. Using this equation, the absolute velocity vectors at the contact
( )
points on the wheel and rail can be defined as vPk = ṙ kP = Ṙ k + 𝛚k × Ak uPk , k = w, r; and
the unit tangent and normal vectors to the rail surface can be defined at the contact point,
respectively, as
( )
̂tr1 = Ar 𝜕uPr∕𝜕sr1 ∕ ||𝜕uPr∕𝜕sr1 || ,⎫
( ) ⎪
̂tr2 = Ar 𝜕uPr∕𝜕sr2 ∕ ||𝜕uPr∕𝜕sr2 || ,⎬ (5.43)
( ) ⎪
̂ r = tr1 × tr2 ∕ ||tr1 × tr2 ||
n ⎭
Contact Problem 201

The absolute velocity of the center of the wheel with respect to the rail vwr = Ṙ w − vPr can
be used to define the forward velocity V as V = vwr ̂tr1 . Because the generalized coordinates
T

and velocities are assumed to be known from the numerical integration of the equations
of motion and because the locations of the contact points on the wheels and rails can be
determined using the constraint or elastic contact formulation, the absolute velocities can
be computed during the dynamic simulations and used to determine the velocity V and
creepages 𝜁 x , 𝜁 y , and 𝜁 s .

Example 5.3
A wheelset w is assumed to travel on a tangent track with a forward velocity equal
to 15 m/s. At a given instant of time, the yaw and roll angles of the wheelset coordi-
nate system are assumed to be equal to zero: that is, 𝜓 w = 𝜙w = 0. The pitch angle at
this instant of time is assumed to be equal to multiple of 2𝜋: that is, 𝜃 w = 2k𝜋 for a
given integer k. In the initial configuration, the local position vector of the contact point
on the right wheel defined in the wheelset coordinate system X w Y w Z w is assumed to
[ ]T [ ]T
be uPw = 0 −la ∕2 −rP = 0 −0.73 −0.46 , where la is the distance between the two
wheel centers, and r P is the rolling radius of the contact of the right wheel. The angle
between the lateral tangent to the right rail tr2 at the contact point and the wheelset lat-
eral axis Y w is defined by contact angle 𝛿 c = 0.027 rad. Assuming component 𝜔wz of the
[ ]T
absolute angular velocity 𝛚w = 𝜔wx 𝜔wy 𝜔wz of the wheelset is equal to zero – that is,
𝜔wz = 0 – determine the values of the creepages for the right wheel contact at the given
configuration.
Solution The longitudinal, lateral, and spin creepages are defined by Eq. 42, respec-
tively, as
( w )T ( w )T
vP − vPr t̂ r1 vP − vPr t̂ r2 (𝛚w − 𝛚r )T n̂ r
𝜁x = , 𝜁y = , 𝜁s =
V V V
w r
where vP and vP are, respectively, the absolute velocities of the contact points on the
wheel and rail; t̂ r1 , t̂ r2 , and n̂ r are, respectively, the unit longitudinal and lateral tangents
and unit normal to the surface at the contact point; V is the wheel forward velocity; and
𝛚w and 𝛚r are the absolute angular velocity vectors of the wheel and rail, respectively.
In the case of fixed tangent rail, one has vPr = 𝟎 and 𝛚r = 0; therefore, the definitions of
the creepages reduce to
vPw T̂tr1 vw T̂tr T
𝛚w n̂r
𝜁x = , 𝜁y = P 2 , 𝜁s =
V V V
The absolute velocity vector of the contact point on the right wheel can be written as
( )
vPw = ṙ wP = Ṙ w + 𝛚w × Aw uPw , where Rw is the global position vector of the origin of
[ ]T
the wheelset coordinate system, 𝛚w = 𝜔wx 𝜔wy 𝜔wz is the absolute angular velocity
vector of the wheelset, uPw is the local position vector of the contact point, and Aw
is the orthogonal transformation matrix that defines the orientation of the wheelset
coordinate system. This transformation matrix, as discussed in Chapter 3, can be
formulated using the three Euler angles 𝜓 w , 𝜙w , and 𝜃 w , which are the result of

(Continued)
202 Mathematical Foundation of Railroad Vehicle Systems

rotations about the three axes Z w , X w , and Y w , respectively. For the values of Euler
angles given in this example, 𝜓 w = 0, 𝜙w = 0, and 𝜃 w = 2k𝜋, where k is an integer,
the transformation matrix Aw can be written as
⎡cos 𝜓 w cos 𝜃 w − sin 𝜓 w sin 𝜑w sin 𝜃 w − sin 𝜓 w cos 𝜑w cos 𝜓 w sin 𝜃 w + sin 𝜓 w sin 𝜑w cos 𝜃 w ⎤
⎢ ⎥
A = ⎢sin 𝜓 w cos 𝜃 w + cos 𝜓 w sin 𝜑w sin 𝜃 w
w
cos 𝜓 w cos 𝜑w sin 𝜓 w sin 𝜃 w − cos 𝜓 w sin 𝜑w cos 𝜃 w ⎥
⎢ ⎥
⎣ − cos 𝜑w sin 𝜃 w sin 𝜑w cos 𝜑w cos 𝜃 w ⎦

⎡1 0 0⎤
⎢ ⎥
= ⎢0 1 0⎥
⎢ ⎥
⎣0 0 1⎦

Vector Aw uPw that defines the global position vector of the contact point, where in this
[ ]T [ ]T
example uPw = 0 −0.73 −0.46 , can be written as Aw uPw = 0 −0.73 −0.46 . The
absolute velocity vector of the origin (reference point) of the wheelset is defined as Ṙ w =
[ ]T
15 0 0 m/s. Because the absolute velocity vector of the contact point in the case of
pure rolling is equal to zero – that is, vPw = 𝟎 – this vector can be written in terms of the
components of the angular velocity vector 𝛚w as
( )
vPw = ṙ wP = Ṙ w + 𝛚w × Aw uPw

⎡15⎤ ⎡−0.46𝜔wy + 0.73𝜔wz ⎤ ⎡0⎤


⎢ ⎥ ⎢ ⎥ ⎢ ⎥
=⎢0⎥+⎢ 0.46𝜔wx ⎥ = ⎢0⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣0⎦ ⎣ −0.73𝜔x w
⎦ ⎣0⎦
Because the yaw angular velocity 𝜔wz is assumed in this example to be zero, the nonholo-
nomic pure rolling contact condition defines 𝜔wx = 0 and 𝜔wy = 15∕0.46 = 32.6086 rad/s;
therefore, the absolute angular velocity vector of the wheelset can be written as
[ ]T
𝛚w = 0 32.6086 0
Using this angular velocity vector, the absolute velocity vector of the geomet-
( )
ric center of the right wheel can be written as ṙ wC = Ṙ w + 𝛚w × Aw uCw , where
[ ]T [ ]T
uCw = 0 −la ∕2 0 = 0 −0.73 0 is the local position vector of the geometric
center of the right wheel defined in the wheelset coordinate system X w Y w Z w . Therefore,
one has
⎡15⎤ ⎡0⎤ ⎡15⎤
( w w) ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
ṙ wC ̇
= R + 𝛚 × A uC = ⎢ 0 ⎥ + ⎢0⎥ = ⎢ 0 ⎥
w w

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦ ⎣0⎦ ⎣ 0 ⎦
Using the value of the contact angle 𝛿 c = 0.027 rad and the wheelset transformation
matrix Aw = I, where I is the 3 × 3 identity matrix, the unit tangent and normal vectors
are defined in the XYZ global coordinate system as
⎡1 0 0 ⎤ ⎡1 0 0 ⎤
[ ] ⎢ ⎥ ⎢ ⎥
̂tr1 t̂ r2 n̂ r = Aw ⎢0 cos 𝛿c sin 𝛿c ⎥ = ⎢0 0.9996 0.027 ⎥
⎢ ⎥ ⎢ ⎥
⎣0 − sin 𝛿c cos 𝛿c ⎦ ⎣0 −0.027 0.9996⎦
Contact Problem 203

Using vector ṙ wC previously evaluated in this example and the longitudinal tangent to
the rail t̂ r1 , the forward velocity of the geometric center of the right wheel with respect
to the rail can be written as V = ṙ wC t̂ r1 = 15 m/s. Recall that due to the condition of pure
T

rolling vPw = 𝟎, the creepages can then be evaluated as


vPw T t̂ r1 vw T t̂ r 𝛚w n̂ r
T

𝜁x = = 0, 𝜁y = P 2 = 0, 𝜁s = = 0.0587
V V V
The results of this example demonstrate that while the longitudinal and lateral creep-
ages are equal to zero, the spin creepage is not equal to zero even in the case of pure
rolling. Some of the creep force formulations discussed in this chapter include coupling
between the tangential creep force and spin moment; therefore, a nonzero value of the
spin creepage can give rise to lateral tangential creep force as well as spin moment when
these formulations are used even in the case of zero longitudinal and lateral creepages.

5.8 CREEP FORCE FORMULATIONS

The constraint and elastic contact formulations can be used to determine the locations of
contact points on the wheel and rail by solving a set of nonlinear algebraic equations for
the wheel and rail surface parameters. The surface parameters can be used to determine
the locations of the contact points on the wheel and rail surfaces, as previously explained.
Knowing the locations of the contact points, the geometric properties of the surfaces, such
as principal curvatures and directions at these points, can be determined. The normal con-
tact forces and creepages can also be computed using the coordinates of the contact points
and the generalized velocities. Knowing the normal contact forces, the geometry of the two
surfaces in contact, and the material properties of the wheel and rail, the dimensions of the
contact ellipse can be evaluated using Hertz theory. Computing tangential creep forces at
a wheel/rail contact point requires knowledge of the normal contact force, dimensions of
the contact ellipse, normalized creepages, and material properties.
Several creep force formulations have been developed and used to model wheel/rail
interaction forces. These formulations have different degrees of generality and are based on
different simplifying assumptions. Understanding the assumptions used in some of these
creep force models is necessary for their proper use in railroad vehicle system applications.
For example, some of the creep force formulations are based on three-dimensional theory,
while others are based on one- or two-dimensional theory. Some formulations account
for spin creepage and spin moment, while others neglect the effect of spin moment. In
most formulations used in railroad vehicle system dynamics, the creep force model is
assumed to depend on the relative velocities; therefore, the influence of inertia on creep
forces is not taken into account in formulating these force models. The inertia effect was
found to be significant at speeds that exceed 500 km/h (Kalker 1986). This speed is much
higher than the operating speeds of trains currently being used, and therefore, using a
quasi-static creep force model is justified. Furthermore, accounting for the inertia effect in
the small contact region can significantly complicate the creep force formulations, which
204 Mathematical Foundation of Railroad Vehicle Systems

are governed in this case by second-order ordinary differential equations. It is also doubtful
that such an inertia effect in most railroad vehicle motion scenarios will have a significant
effect on overall train dynamics.
More comprehensive creep force models can be developed using the general
three-dimensional theory of elasticity and the computational FE method. Nonetheless, as
more accuracy is demanded, the computational cost rises; therefore, a trade-off must be
made in order to efficiently solve and perform computer simulations of complex railroad
vehicle systems that consist of a large number of components and joints. An example of a
creep force formulation is the two-dimensional simplified theory of rolling contact in which
the relative displacement between the wheel and the rail at the contact point is defined in
a coordinate system X c Y c Z c in which the components of the traction force are defined. If
Ac is the orthogonal transformation matrix that defines the orientation of this coordinate
system with respect to the global coordinate system, and if the components of tangential
force F tx and F ty are defined along the X c and Y c axes of the coordinate system X c Y c Z c ,
respectively, the displacement-force relationship of the simplified theory of rolling contact
T cT
( w ) [ ]T
is assumed to take the simple form Ac rwr P =A rP − rrP = dx Ftx dy Fty 0 (Kalker
1973), where dx and dy are, respectively, longitudinal and lateral flexibility coefficients
that depend on the material properties and geometry of the wheel/rail contact surfaces,
P = rP − rP , and rP = R + A uP , k = w, r is the global position vector of the contact
rwr w r k k k k

point on body k. The normal contact force, which depends on the relative displacement
between the wheel and rail along the normal to the contact surfaces, can still be determined
using one of the formulations previously discussed in this chapter.
With the advances made in computational methods and computer technology, many of
the simplified wheel/rail creep force formulations are no longer being used, because more
sophisticated railroad vehicle algorithms are being developed. Nonetheless, knowledge
of the history of developing such creep force models is important for understanding the
assumptions originally made for the analysis of complex wheel/rail interaction forces
and for explaining the need to relax some of these assumptions as more accuracy is
demanded.

Carter’s Theory One of the early creep force theories, based on an analytical rela-
tionship between longitudinal tangential force and longitudinal creepage, was proposed
by Carter (1926). Carter’s theory makes the following assumptions: (i) simplified wheel
and rail surface profile geometries are considered; (ii) the wheel is represented by a
cylinder with a radius that is much larger than the dimensions of the contact area;
(iii) the rail is represented by a flat plate; and (iv) the shape of the contact area is
assumed to be a rectangular strip with small dimensions compared to the wheel and
rail dimensions such that the assumptions of an infinite elastic medium with a localized
pressure distribution can be made. Carter’s creep force model that defines the relationship
between longitudinal tangential force and longitudinal creepage is given by (Kalker
1991)
⎧ ( ( )2 )
⎪𝜇Fnl −kc 𝜁x + 0.25 kc 𝜁x ||𝜁x || if kc ||𝜁x || ≤ 2
Ftl = ⎨ ( ) (5.44)
⎪−𝜇Fnl sign 𝜁x if kc ||𝜁x || > 2

Contact Problem 205

where F tl is the tangential force per unit lateral length in the longitudinal direction; 𝜇 is the
coefficient of friction; F nl is the total normal force per unit lateral length; 𝜁 x is the longitu-
dinal creepage defined in Carter’s theory as 2(V t − V c )/(V t + V c ), where subscripts t and c
refer, respectively, to tangential forward and circumferential velocities; and kc = 4r w /𝜇a is
Carter’s creepage coefficient, where r w is the wheel rolling radius and a is the length of the
contact ellipse semi axis in the rolling direction. In Carter’s theory, the equivalent length lt
of the contact area in the transverse direction of the rail can be written as lt = (4b/3), where b
is the length of the contact ellipse semi axis in the lateral direction. The preceding equation
defines a nonlinear relationship between tangential creep force and the longitudinal creep-
age. The difference between the results obtained using this nonlinear relationship and the
results of the linear force-creepage relationship is shown in Figure 6.

1.2
Coulomb’s maximum value
1.0

0.8 Linear law


Ftl /μFnl

0.6
Carter’s law

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
k│ζx│

Figure 5.6 Carter’s creep force model.

It is clear that Carter’s theory makes several simplifying assumptions, including neglect-
ing the effect of the lateral and spin creepages 𝜁 y and 𝜁 s , simplified geometries of the
wheel and rail, a simplified expression for the forward wheel velocity, and the restriction
to one-dimensional analysis. Therefore, such a creep contact theory is not suited for the
analysis of complex railroad vehicle systems that can have general motion and in which
wheel/rail interaction creep forces cannot be predicted using a one-dimensional theory.

Johnson and Vermeulen’s Theory Carter’s theory was generalized by Johnson to the
three-dimensional analysis of two spheres without consideration of spin creepage (Johnson
1958a,b) and was further extended by Vermeulen and Johnson to smooth half-space without
consideration of spin creepage (Vermeulen and Johnson 1964). In Johnson and Vermeulen’s
theory, the contact area is assumed elliptical with uneven distribution of the slip and stick
regions. Furthermore, the adhesion area is also assumed elliptical, as shown in Figure 7. As
illustrated in this figure, the elliptical adhesion area and the contact ellipse are assumed to
206 Mathematical Foundation of Railroad Vehicle Systems

b′

a
a′
Rolling direction
X
Slip area

Adhesion area
b

Figure 5.7 Contact area in Johnson and Vermeulen’s theory.

share the leading edge. In Johnson and Vermeulen’s contact model, tangential force vector
Ft is defined as
⎧ [( ( )) ] (( 𝜁 ) ( ) )
𝜁yn
[ ] ⎪𝜇Fn 1 − l𝜁 ∕3 3 − 1 | |
xn
i+ j for |l𝜁 | ≤ 3,
Ftx ⎪ l𝜁 l𝜁 | |
Ft = =⎨ (( ) ( ) ) (5.45)
Fty ⎪−𝜇F 𝜁xn 𝜁yn | |
i + j for | l | > 3,
⎪ n
l𝜁 l𝜁 | 𝜁|

where 𝜁 xn = 𝜋abG𝜁 x /𝜇F n hx is a normalized longitudinal creepage; 𝜁 yn = 𝜋abG𝜁 y /𝜇F n hy
√( ) ( )2 [ ]T [ ]T
2
is a normalized lateral creepage; l𝜁 = 𝜁xn + 𝜁yn ; i = 1 0 ; j = 0 1 ; F n is the
normal contact force; G is the modulus of rigidity; 𝜁 x and 𝜁 y are, respectively, the longi-
tudinal and lateral creepages; a and b are, respectively, the lengths of the semi axes of the
contact ellipse in the rolling and lateral directions; 𝜇 is the coefficient of friction; and

⎧ √ ⎫
( )
⎪Be − 𝜈 De − Ce for a ≤ b, e = 1 − (a∕b)2 ⎪
hx = ⎨ √ ⎪
[ ( )]
⎪(b∕a) De − 𝜈 De − Ce for a ≥ b, e = 1 − (b∕a)2 ⎪⎪

⎧ √ ⎬ (5.46)
2 2⎪
⎪Be − 𝜈(a∕b) Ce for a ≤ b, e = 1 − (a∕b) ⎪
hy = ⎨ [ ] √

⎪(b∕a) De − 𝜈Ce for a ≥ b, e = 1 − (b∕a)2 ⎪
⎩ ⎭
In this equation, 𝜈 is Poisson’s ratio, and Be , De and Ce are the elliptical integrals of argument
e. The differences in the results predicted by Johnson and Vermeulen’s creep force model
and the experimental measurements conducted were attributed to the assumption of an
elliptical adhesion region. Furthermore, because Johnson and Vermeulen’s theory does not
account for spin creepage, which has been found to be significant, such theory is rarely used
in the development of the new computational railroad vehicle system algorithms.

Kalker’s Linear Theory In Kalker’s linear theory, creepages are assumed to be small to
justify using a linear force-creepage relationship (Kalker 1967). Because of this assumption,
the contact area is assumed to be predominantly an adhesion area. In formulating the linear
theory for wheel/rail contact, Kalker introduced the constants G = (Gw + Gr )/2Gw Gr and
𝜈 = G(𝜈 w Gr + 𝜈 r Gw )/2Gw Gr , where Gk and 𝜈 k , k = w, r are, respectively, the modulus of
Contact Problem 207

rigidity and Poisson ratio of the wheel and rail, respectively. In Kalker’s linear theory, which
accounts for the effect of spin creepage, tangential creep forces and spin moments can be
written, respectively, in terms of the creepages as

Ftx = −Gabd(11 𝜁x ) ⎫

Fty = −Gab d22 𝜁y + abd23 𝜁s ⎪
( √ )⎬ (5.47)
Ms = −Gab − abd23 𝜁y + abd33 𝜁s ⎪

where 𝜁 x , 𝜁 y , and 𝜁 s are, respectively, the longitudinal, lateral, and spin creepages; a and b
are, respectively, the lengths of the semi axes of the contact ellipse in the rolling and lateral
directions; and d11 , d22 , d23 , and d33 are creepage coefficients that are functions of the Pois-
son ratio and the a/b ratio and are given in Table 4 for different values of the a/b ratio and
Poisson ratio (Kalker 1990).

Table 5.4 Coefficients in Kalker’s linear theory. Source: Courtesy of Kalker, J.J.

d 11 d 22 d 23 = − d 32 d 33

g 𝝂=0 0.25 0.5 𝝂=0 0.25 0.5 𝝂=0 0.25 0.5 𝝂=0 0.25 0.5

(a/b)
0.1 2.51 3.31 4.85 2.51 2.52 2.53 0.334 0.473 0.731 6.42 8.28 11.7
0.2 2.59 3.37 4.81 2.59 2.63 2.66 0.483 0.603 0.809 3.46 4.27 5.66
0.3 2.68 3.44 4.8 2.68 2.75 2.81 0.607 0.715 0.889 2.49 2.96 3.72
0.4 2.78 3.53 4.82 2.78 2.88 2.98 0.720 0.823 0.977 2.02 2.32 2.77
0.5 2.88 3.62 4.83 2.88 3.01 3.14 0.827 0.929 1.07 1.74 1.93 2.22
0.6 2.98 6.72 4.91 2.98 3.14 33.1 0.930 1.03 1.18 1.56 1.68 1.86
0.7 3.09 3.81 4.97 3.09 3.28 3.48 1.03 1.14 1.29 1.43 1.50 1.60
0.8 3.19 3.91 5.05 3.19 3.41 3.65 1.13 1.25 1.40 1.34 1.37 1.42
0.9 3.29 4.01 5.12 3.29 3.54 3.82 1.23 1.36 1.51 1.27 1.27 1.27

(b/a)
1.0 3.4 4.12 5.2 3.40 3.67 3.98 1.33 1.47 1.63 1.21 1.19 1.16
0.9 3.51 4.22 5.3 3.51 3.81 4.16 1.44 1.57 1.77 1.16 1.11 1.06
0.8 3.65 4.36 5.42 3.65 3.99 4.39 1.58 1.75 1.94 1.10 1.04 0.954
0.7 3.82 4.54 5.58 3.82 4.21 4.67 1.76 1.95 2.18 1.05 0.965 0.852
0.6 4.06 4.78 5.8 4.06 4.50 5.04 2.01 2.23 2.50 1.01 0.892 0.751
0.5 4.37 5.10 6.11 4.37 4.90 5.56 2.35 2.62 2.96 0.958 0.819 0.650
0.4 4.84 5.57 6.57 4.84 5.48 6.31 2.88 3.24 3.70 0.912 0.747 0.549
0.3 5.57 6.34 7.34 5.57 6.40 7.51 3.79 4.32 5.01 0.868 0.674 0.446
0.2 6.96 7.78 8.82 6.96 8.14 9.79 5.72 6.63 7.89 0.828 0.601 0.341
0.1 10.7 11.7 12.9 10.7 12.8 16.0 12.2 14.6 18.0 0.795 0.526 0.228

g = 0, C11 = 𝜋 2 ∕4 (1 − 𝜈) ; C22 = 𝜋 2 ∕4; C23 = −C32 = 𝜋 g∕3; and C33 = 𝜋 2 ∕16 (1 − 𝜈) g.
208 Mathematical Foundation of Railroad Vehicle Systems

The preceding equation shows that the lateral creep force depends on the spin creepage,
and the spin moment depends on the lateral creepage. That is, the lateral creep force and
spin moments are coupled in this linear theory. The creepage coefficients in the preced-
ing equation were calculated using the program CONTACT with the assumption of small
relative slip and without assuming an elliptical contact area. The results showed that the
error is less than 5%, confirming the accuracy of the coefficients presented in Table 4 in the
case of small creepages. The linear theory has been used in railroad vehicle dynamics lit-
erature where interpolations based on the data presented in Table 4 are used to determine
the creepage coefficients for different values of the contact ellipse ratio (a/b) (Haque et al.
1979). Due to the limitations of the linear theory presented in this section, more general
and nonlinear formulations that relax the assumptions of the linear theory were proposed
by Kalker. Some of these formulations, which are used in computational railroad vehicle
algorithms, are discussed before concluding this section.

Example 5.4
A wheel and a rail are assumed to be in contact and are made of steel with a Poisson
ratio equal to 0.25 and modulus of rigidity equal to 8 × 1010 N/m2 . At a given configu-
ration, the dimensions of the contact ellipse are found to be a = 5.8805 × 10−3 m and
b = 2.94025 × 10−3 m; and the longitudinal, lateral, and spin creepages are, respectively,
given as 𝜁 x = 0, 𝜁 y = 0.008, and 𝜁 s = 0.0587 m−1 . Use Kalker’s linear theory to compute
tangential creep forces and spin moment.
Solution Using the data of Table 4, the creepage coefficients in Kalker’s lin-
ear theory can be determined for the ratio of the semi axis of the contact ellipse
a/b = 5.8805/2.94025 = 2 and for the Poisson ratio 𝜈 = 0.25 as d11 = 5.10, d22 = 4.9,
d23 = 2.62, and d33 = 0.819. Substituting these coefficients into Eq. 47, the longitudinal
and lateral creep forces and the spin moment can be evaluated, respectively, as
Ftx = −Gabd11 𝜁x = 0
( √ )
Fty = −Gab d22 𝜁y + abd23 𝜁s = −551.06 N
( √ )
Ms = −Gab − abd23 𝜁y + abd33 𝜁s = −1.193 N.m

Heuristic Nonlinear Creep Force Model Johnson and Vermeulen’s theory neglects
the effect of spin creepage and considers only longitudinal and lateral creepages. The effect
of spin creepage, however, cannot be ignored in many motion scenarios, including flange
contacts, as previously mentioned. Kalker’s linear theory also has its own limitations due
to the assumption of small creepages. Therefore, Johnson and Vermeulen’s theory and
Kalker’s linear theory are not suitable for developing general computational algorithms for
solving the highly nonlinear wheel/rail contact problem characterized by geometric and
force nonlinearities. Geometric nonlinearities are the result of the nonlinear geometry of
the contact surfaces, and force nonlinearities are the result of the stick–slip phenomenon.
Contact Problem 209

To address the limitations of the creep force theories mentioned earlier, some researchers
proposed nonlinear contact force formulations that take into account the effect of spin
creepage. An example of these formulations is the work of Shen et al. (1983), who proposed
to combine features of Johnson and Vermeulen’s nonlinear theory, which does not account
for spin creepage, and Kalker’s linear theory, which accounts for spin creepage, with the
goal of developing a heuristic nonlinear theory. In the heuristic nonlinear theory, the creep
forces of Kalker’s linear theory are used in the evaluation of the creep force in Johnson
and Vermeulen’s theory to obtain a correction factor that takes into account the effect of
spin creepage. This correction factor is a nonlinear function of the creepages; therefore,
the resulting creep force formulation is nonlinear and takes into account the effect of spin
creepage. To develop the heuristic nonlinear theory, the first two equations of Eq. 47 are
used. These equations can be written as
( ) }
Ftx K = −Gabd(11 𝜁x )
( ) √ (5.48)
Fty K = −Gab d22 𝜁y + abd23 𝜁s

where subscript K refers to Kalker’s linear theory. The magnitude of the resultant creep
force in the preceding equation can be written as
( ) √( ) ( )2
2
Ft K = Ftx K + Fty K

( )2 ( √ )2
= Gab d11 𝜁x + d22 𝜁y + abd23 𝜁s (5.49)

While the magnitude of this resultant creep force cannot exceed the friction force 𝜇F n , John-
son and Vermeulen’s nonlinear theory can be used to determine a force limit (F t )J . To this
( ) ( )
end, l𝜁 in Johnson and Vermeulen’ theory is approximated as l𝜁 ≈ l𝜁 = Ft K ∕𝜇 Ft J . Sub-
stituting this equation into the resultant creep force of Johnson and Vermeulen’s theory,
one obtains (Shen et al. 1983)
⎧ [( ) ( )2 ( )3 ]
( ) ⎪𝜇F 1 1
l𝜁 − 3 l𝜁 + 27 l𝜁 , l𝜁 ≤ 3
Ft J = ⎨ n (5.50)
⎪𝜇Fn , l𝜁 > 3

Using this value, one can define a correction factor kt = (F t )J /(F t )K and use it to define the
tangential creep force as
[ ] [( ) ]
Ftx Ftx
Ft = = kt ( ) K (5.51)
Fty Fty K
This expression for tangential creep force accounts for the effect of spin creepage. As
reported in the literature, while the heuristic nonlinear theory produces more realistic
creep force results outside the linear range as compared to Kalker’s linear theory (Shen
et al. 1983), this theory produces unsatisfactory results in the case of high values of spin
creepage (Kalker 1991). Nonetheless, Shen et al. (1983) demonstrated that the results of the
heuristic theory are in good agreement with the results obtained using Kalker’s simplified
theory in some scenarios. Kalker’s simplified theory is discussed later in this section.
210 Mathematical Foundation of Railroad Vehicle Systems

Polach’s Nonlinear Creep Force Model If the maximum stress distribution in Hertz
contact theory is denoted as 𝜎 m , the maximum tangential stress can be written as 𝜏 m = 𝜇𝜎 m ,
where 𝜇 is the friction coefficient. In Polach’s nonlinear creep force model, the contact area
is assumed to be elliptic, and the relative displacement between the two solids and tangen-
tial stress in the adhesion region are assumed to increase linearly from the first to the second
edge of the contact area. Sliding is assumed to take place when tangential stress reaches its
maximum value 𝜏 m = 𝜇𝜎 m . Tangential force is defined in this nonlinear model as
( )
2𝜇Fn 𝜒P
( )2 + tan 𝜒P
−1
FPtx = − (5.52)
𝜋 1 + 𝜒P
( )
where F n is the normal contact force, 𝜒P = G𝜋ab𝜒 P ∕4𝜇Fn l𝜁 s , G is the modulus of rigid-
ity, 𝜒 P√is a constant that depends on Kalker’s
√( ) coefficients 11 and d22 and is defined as
d√
( )2 ( )2 2 ( )2 ( )2 ( )2
𝜒P = d11 𝜁x ∕l𝜁 + d22 𝜁y ∕l𝜁 , l𝜁 = 𝜁x + 𝜁y , l𝜁 s = 𝜁x + 𝜁ys , and 𝜁 ys is a
modified lateral creepage used to account for the effect of spin creepage. The modified lat-
eral creepage 𝜁 ys is evaluated according to the following definition:

⎧ | | | |
⎪𝜁y , |𝜁y + 𝜁s a| ≤ |𝜁y |
| | | |
𝜁ys = ⎨( ) (5.53)
| | | |
⎪ 𝜁y + 𝜁s a , |𝜁y + 𝜁s a| > |𝜁y |
| | | |

where a is the dimension of the semi axis of the contact ellipse in the rolling direction. Mak-
ing the assumption that the moment resulting from spin creepage and/or lateral creepage
is small compared to other external and inertia moments applied to the wheel or wheelset,
the lateral tangential creep force that takes into consideration the effect of spin creepage
can be written as
9 [ ( )]
FPty = − a𝜇Fn KP 1 + 6.3 1 − e−a∕b (5.54)
16

| |(( ) ( ) ) ( )3
where K P is a constant defined as KP = |𝜀y | (𝛿)3 ∕3 − (𝛿)2 ∕2 + 1∕6 − (1∕3) 1 − 𝛿2 ,
√| | ( ( ))
𝛿 = ((cP )2 − 1)/((cP )2 + 1), cP = 8Gb abd23 𝜁ys ∕3𝜇Fn 1 + 6.3 1 − e−a∕b , b is the dimen-
sion of the semi axis of the contact ellipse in the lateral direction, and d23 is Kalker’s creep
coefficient previously introduced in this chapter. Using these definitions, the tangential
creep forces in Polach’s nonlinear model are written as
( ) ( ) ( )
𝜁x 𝜁y 𝜁s
Ftx = FPtx , Fty = FPtx + FPty (5.55)
l𝜁 s l𝜁 s l𝜁 s
Because numerical investigations have shown that Polach’s nonlinear tangential creep force
model defined by the preceding equation is accurate, this model has been implemented in
several computational algorithms developed for the nonlinear dynamic analysis of railroad
vehicle systems (Polach 1999).

Simplified Theory of Contact To develop expressions for tangential forces using the
simplified theory of contact, a coordinate system X c Y c Z c is defined in the neighborhood of
the contact point. In this coordinate system, the relative displacements between the wheel
and rail can be defined using the transpose of the orthogonal transformation matrix Ac that
Contact Problem 211

defines the orientation of the coordinate system X c Y c Z c in the global coordinate system
XYZ. If the tangential creep forces at a point in the X c Y c Z c coordinate system are given by
[ ]T
the vector Ft = F tx F ty , one can introduce flexibility coefficients dx and dy and write the
relative displacements between the wheel and rail in the neighborhood of the contact point
as
( wr )c T cT
( w ) [ ]T
rP = Ac rwr
P =A rP − rrP = dx F tx dy F ty 0 (5.56)

If x and y are assumed to be the longitudinal and lateral surface parameters in the coor-
dinate system X c Y c Z c at contact point P, one can define the displacement vector uwr
t =
T ( wr ( wr T r ) r )
wr c ̂ ̂
ut (x, y) = A rP − rP n n , which has a zero component along the normal to the
contact surfaces. Using the longitudinal surface parameter x, the wheel forward velocity V
can be written as V = dx/dt. In the case of steady-state rolling, the slip is assumed to be
small in the contact area, and the components of the relative velocity along the normal can
be neglected; therefore, one has (Kalker 1979)

⎡𝜁x − 𝜁s y⎤
( )c T ⎢ ⎥
vPwr = Ac ṙ wr
P = V ⎢𝜁y + 𝜁s x⎥ (5.57)
⎢ ⎥
⎣ 0 ⎦

This velocity vector can also be defined as


( )c 𝜕utwr dx 𝜕u wr
vPwr = =V t (5.58)
𝜕x dt 𝜕x
Using the preceding two equations, one has

⎡𝜁x − 𝜁s y⎤
⎢ ⎥ 𝜕u wr
V ⎢𝜁y + 𝜁s x⎥ = V t (5.59)
⎢ ⎥ 𝜕x
⎣ 0 ⎦

which upon using Eq. 56, leads to


[ ] [ ( )]
𝜁x − 𝜁s y dx 𝜕F tx ∕𝜕x
V =V ( ) (5.60)
𝜁y + 𝜁s x dy 𝜕F ty ∕𝜕x

Assuming that the longitudinal and lateral creepages are constant in the small contact area,
integrating the preceding equation leads to the definition of tangential creep forces in the
simplified theory of contact as
( )
⎡ x 𝜁x − 𝜁s y + c1 (y) ⎤
[ ] ⎢ ⎥
F tx ⎢ dx ⎥
Ft = =⎢ ( 1
)
⎥ (5.61)
F ty ⎢ x 𝜁 y + 𝜁
2 s
x + c2 (y) ⎥
⎢ ⎥
⎣ dy ⎦

where c1 (y) and c2 (y) are constants of integration, which can be evaluated using the condi-
tion that the tangential creep force is zero at the leading edge of the contact ellipse. Using
212 Mathematical Foundation of Railroad Vehicle Systems

this condition, one obtains


( )( )
⎡ x − xb 𝜁x − 𝜁s y ⎤
[ ] ⎢ ⎥
F tx ⎢ d x ⎥
Ft = =⎢ ( ) 1 ( ( )2 ⎥) (5.62)
⎢ 𝜁y x − xb + 2 𝜁s (x) − xb
2
F ty ⎥
⎢ ⎥
⎣ dy ⎦

where xb = a 1 − (y∕b)2 , and a and b are the dimensions of the semi axes of the con-
tact ellipse in the rolling and lateral directions, respectively. Using the preceding equation,
neglecting the spin moment, and integrating over the contact area, one obtains
⎡ 8a2 b𝜁x ⎤
[ ] −
Ftx F tx dx dy⎤ ⎢
⎡∫ b ∫ xb (y) 3dx ⎥
= ⎢ b x (y) ⎥=⎢ ⎥
−b −xb (y)
Ft = (5.63)
Fty ⎢∫ ∫ b F dx dy⎥ ⎢ 8a b𝜁y 𝜋a b𝜁s ⎥
2 3
⎣ −b −xb (y) ty ⎦ ⎢− − ⎥
⎣ 3dy 4dy ⎦
To determine the flexibility coefficients dx and dy , the expressions for tangential creep forces
previously obtained in Kalker’s linear theory ( can be used.
√ These ) expressions were given in
Eq. 47 as F tx = − Gabd11 𝜁 x and Fty = −Gab d22 𝜁y + abd23 𝜁s . By equating these tangen-
tial creep forces with the forces in Eq. 63, the flexibility coefficients dx and dy are defined in
terms of the coefficients of the Kalker’s linear theory as
8a 8a 𝜋a2
dx = , dy1 = , dy2 = √ (5.64)
3Gd11 3Gd22 4G abd23
Two values of dy are provided in this equation, because the second equation in Eq. 63 has
two coefficients associated with creepages 𝜁 y and 𝜁 s . Jacobson and Kalker (2001) suggested
combining the three flexibility coefficients in one coefficient.
It is clear that by neglecting the effect of the spin moment, an analytical solution for the
tangential creep forces can be obtained using the simplified theory of contact. This theory
was used as the basis for developing the computer program FASTSim, which is widely used
in railroad vehicle system applications (Kalker 1982). Using numerical techniques allows
for using FASTSim in predicting the tangential creep forces in more general wheel/rail con-
figurations. In FASTSim, a strip discretization of the wheel/rail contact surface is made, and
such a discretization is used to evaluate the resultant tangential creep forces.

More General Wheel/Rail Contact Theories The assumptions of the simplified the-
ory of contact can be relaxed to obtain more general nonlinear wheel/rail contact theories.
This can be accomplished at the expense of increasing the computational cost. As previ-
ously mentioned, a general approach can be based on the FE discretization in which FE
meshes are developed for the two solids in contact. The motion of the solid surfaces can be
monitored to determine the distances between the points on the surfaces. The search for the
contact points can be based on formulating algebraic equations and using these equations
to minimize the distances between the two surfaces, or using a numerical approach and
selected discrete points to determine the points of contact between the two solids. Regard-
less of the contact-search method used, the FE approach is more general since it can handle
Contact Problem 213

both conformal and non-conformal contacts and allows for predicting stresses at the contact
points using a more general continuum mechanics approach without making simplifying
assumptions. These stresses can be used to determine the normal forces at each pair of
contact points. The contact stresses can be predicted using linear or nonlinear constitutive
models based on the theory of elasticity. However, using the FE approach can be compu-
tationally prohibitive in the case of railroad vehicle systems where dynamic simulations
over an extended time may be required, as previously mentioned in this chapter. More-
over, such a computationally intensive approach may not lead to significant improvements
in the accuracy of the results in most railroad vehicle system applications. With this fact in
mind, Kalker developed an approach that is more general than the simplified theory of con-
tact. This approach is implemented in a computer program called USETAB (Kalker 1990),
which is commercially available. The USETAB program is designed for a general wheel/rail
contact problem; therefore, it requires the evaluation of more coefficients as compared to
the simplified theory of contact. The evaluation of these coefficients is necessary in order
to account for all possible wheel/rail contact configurations. USETAB produces two sets
of coefficient tables: one contains the creepage coefficients that enter into formulating the
tangential creep forces, while the second contains the constants used in Hertz theory. To
account for all possible wheel/rail contact configurations, the Hertz constants included in
the second table are produced using a large number of runs of another computer program
called CON93, which was also developed by Kalker.
To use the USETAB program in a simulation of wheel/rail contact, one must provide the
normal contact force, dimensions of the contact ellipse semi axes, creepages, and material
properties. The normal force and creepages must be defined in a contact frame that has
one of its axes along the normal to the contact surfaces at the contact point. The longitu-
dinal, lateral, and spin creepages must also be defined in this contact frame. Based on this
input, USETAB computes the longitudinal and lateral components of the tangential creep
force as well as the spin moment. These force and moment components are defined in the
directions of the axes of the contact frame in which the creepages are defined. The results
of the USETAB computer program have been found to be accurate when compared with
the results of the fully nonlinear contact theory. The difference between the creep force
results predicted using USETAB and the fully nonlinear theory was found to be approxi-
mately 1.5%, while the difference between the results of FASTSim and the fully nonlinear
theory was found to be 15%. Despite the fact that the USETAB algorithm is not based on a
fully nonlinear theory, it has proven to give accurate results in a large number of railroad
vehicle system applications without sacrificing computational efficiency.

5.9 CREEP FORCE AND WHEEL/RAIL CONTACT


FORMULATIONS
It is clear that the accuracy of the tangential contact force results obtained using the
creep force formulations presented in the preceding section depends on the accuracy
of predicting the location of the wheel/rail contact points. The contact points can be
determined online using one of the constraint and elastic contact formulations previously
214 Mathematical Foundation of Railroad Vehicle Systems

discussed in this chapter. In these two formulations, a set of nonlinear algebraic equations
[ ]T
for each wheel/rail contact is solved for the four surface parameters s = sw1 sw2 sr1 sr2 .
These surface parameters, which allow for modeling general three-dimensional contact,
can be used to determine the locations of the contact points on the wheel and rail surfaces,
as previously explained. Using the locations of the contact points and the associated surface
parameters, surface geometric properties such as the principal curvatures and principal
directions can be computed and used in Hertz contact theory with the normal force and
material properties of the wheel and rail to determine the dimensions of the contact
ellipse.
While the constraint and elastic contact formulations previously discussed in this chapter
are three-dimensional contact formulations and are recommended for computer imple-
mentation because of their generality and ability to capture spatial motion scenarios, other
approaches can be used to determine the locations of the contact points. Some of these
approaches employ only two surface parameters, one for the wheel profile and the second
for the rail profile, and make the assumption that contact occurs in a plane defined by the
lateral tangent and the normal to the surface. Other formulations are based on nodal search
and do not require a certain degree of surface smoothness. However, these approaches may
require some form of interpolation to determine the geometric properties that enter in the
definition of the contact area. Using lookup tables to determine the location of the contact
points on the wheel and rail using precomputed data is not discussed in this section because
of the serious limitations of such an approach, the difficulties that might be encountered,
and the amount of data storage required for capturing general three-dimensional wheel/rail
contact scenarios. On the other hand, examples of the nodal search and planar contact
approaches are briefly discussed in this section for completeness.

Nodal Search Method The nodal search method is an elastic contact formulation that
can be used as an alternative to algebraic equations for solving for the surface parame-
ters that define the contact point locations. In this method, point meshes are created on
the wheel and rail surfaces to define the profile of the wheel and rail (Shabana et al. 2004,
2005). Using discrete mesh points, the distance between these points on the wheel and rail
profiles can be computed and used to determine whether two points are in contact. While
such an approach has the advantage of not requiring a certain degree of smoothness of
the wheel and rail surfaces during the search process, as previously mentioned, it has the
drawback that the change in the wheel and rail lateral surface parameters is not smooth
since the contact is assumed to occur at discrete mesh points. As the contact point changes
from one mesh point to another, there can be a jump discontinuity in the wheel/rail relative
velocity at the contact point, and this velocity jump discontinuity leads to creepage discon-
tinuity, which in turn leads to creep force discontinuity. Numerical experimentation has
shown that the force discontinuity can be significant because creepage coefficients used in
formulating creep forces are normally very large. The impulsive force resulting from such a
discontinuity can lead to numerical problems and deterioration of the solution accuracy if
interpolation schemes are not used to achieve a certain degree of smoothness. The coordi-
nates of the contact points determined using the nodal search can be used with the absolute
nodal coordinate formulation (ANCF) geometry to determine the principal curvatures and
principal directions that are required for computing the dimensions of the contact ellipse.
Contact Problem 215

s2w w
Z

w
X

s1r

Figure 5.8 Nodal search method.

The main steps used in the nodal search for determining the locations of the contact points
on the wheel and rail surfaces can be summarized as follows (Shabana et al. 2004, 2008):

1. The rail arc length sr1 traveled by the wheel is computed and used to determine the rail
cross section in which potential contact points may lie. The distance traveled by the
wheel can be determined by selecting a reference point Q on the wheel, as shown in
Figure 8, and solving the first-order differential equation ṡ r1 = ṙ wQ T t̂ r1 simultaneously with
the system equations of motion, where ṙ wQ is the absolute velocity of point Q and t̂ r1 is a
unit vector along the longitudinal tangent to the rail surface. As an alternative to using
the first-order ordinary differential equation ṡ r1 = ṙ wQ T t̂ r1 , one can use the Cartesian coor-
dinates of the wheel or wheelset and solve a set of nonlinear algebraic equations for the
trajectory coordinates. By varying the rail lateral surface parameter sr2 for the given rail
longitudinal surface parameter sr1 , the locations of the mesh points that define the rail
profile can be determined, as previously explained in this chapter.
2. Wheel angular parameter sw2 is computed and used to define the wheel diametric section
in which potential contact points may lie. Knowing the component of the angular veloc-
ity of the wheel about its own axis, a simple procedure can be used to determine angular
surface parameter sw2 . Knowing angular surface parameter sw2 , the location of any point
on the wheel profile can be determined if wheel lateral surface parameter sw1 is given. By
varying wheel lateral surface parameter sw1 for given angular surface parameter sw2 , the
locations of the mesh points that define the wheel profile can be determined, as previ-
ously explained in this chapter.
3. The distances between the discrete mesh points on the wheel and rail are computed
to determine which points come in contact. If two points come into contact, rail pro-
file parameter sr2 and wheel profile parameter sw1 associated with the contact points are
[ ]T
determined. Knowing the four surface parameters s = sw1 sw2 sr1 sr2 , the geometric
properties of the surfaces can be approximated and used with the normal force and the
material properties to determine the dimensions of the contact area using Hertz theory.
This approach allows for using the same procedure previously described to determine
the tangential creep forces when the constraint and elastic contact formulations are used
to determine the contact points. Because the nodal search method is not a constraint
216 Mathematical Foundation of Railroad Vehicle Systems

formulation and allows for wheel/rail separations, the normal force can be determined
using a compliant force model with assumed stiffness and damping coefficients.

It is clear that one of the disadvantages of the nodal search method is the need to introduce
a number of first-order ordinary differential equations equal to the number of wheel/rail
contacts in order to solve for the distance traveled by each wheel. Furthermore, this method
can lead to a large number of contact points for each wheel/rail contact if fine point meshes
are used. On the other hand, using a coarse mesh may lead to missing potential contact
points and to a higher degree of discontinuities. Therefore, when fine point meshes are used
with the nodal search method, an efficient search algorithm is needed. The mesh points
can be grouped in batches, and a limit can be introduced on the number of batches that
have potential contact points. The number of contact points is assumed not to exceed the
number of contact batches. For a given batch, one pair of points that leads to the maxi-
mum wheel/rail penetration can be selected to define the contact point. This nodal search
algorithm allows for the possibility of having several wheel/rail contact points.

Planar Contact In some wheel/rail contact formulations used in existing research and
commercial software, the wheel/rail contact problem is simplified. Instead of using two
surface parameters for each contact surface, only one profile parameter sw1 is used for the
wheel, and one profile parameter sr2 is used for the rail. The other two surface parameters
sw2 and sr1 are assumed to be known from the wheel and rail configurations and do not enter
into the definition of the algebraic equations used in the search algorithm. Because of this
assumption, the contact between the wheel and rail is reduced to a planar curve repre-
sentation instead of the three-dimensional surface representation. This simplified planar
contact approach neglects kinematic couplings between the geometric surface parameters
and leads to approximation in predicting the location of the contact points. Since one needs
to solve for only two surface parameters, the planar constraint contact conditions can be for-
mulated using three algebraic equations instead of five; two of these equations can be used
to determine the two surface parameters sw1 and sr2 , and the third equation is used to elim-
inate the freedom of the wheel to translate with respect to the rail along the normal to the
contact surfaces. Similar algebraic equations can be used in the elastic contact formulation,
as previously discussed in this chapter.
To formulate the planar contact conditions, an intermediate wheel coordinate system,
which does not share the wheel pitch rotation, is introduced. This coordinate system can be
conveniently defined using the wheel trajectory coordinates introduced in Chapter 3. The
[ ]T
six trajectory coordinates are defined by the vector pw = sw ywr zwr 𝜓 wr 𝜑wr 𝜃 wr ,
where sw is the rail arc length that defines the location of the origin of the trajectory wheel
coordinate system X tw Y tw Z tw ; ywr and zwr are the coordinates of the wheel center of mass
with respect to the origin of the trajectory wheel coordinate system X tw Y tw Z tw ; and 𝜓 wr ,
𝜑wr , and 𝜃 wr are the three Euler angles that define the orientation of the wheel coordinate
system X w Y w Z w with respect to the trajectory wheel coordinate system X tw Y tw Z tw . As
described in Chapter 3, the sequence of Euler angle rotations used is Z w − X w − Y w . Regard-
less of the coordinates used (Cartesian or trajectory), the vector of trajectory coordinates
pw is assumed to be known from the solutions of the dynamic equations of motion since
this vector can be systematically determined using absolute Cartesian coordinates.
Contact Problem 217

To define the intermediate wheel coordinate system, which does not have the pitch rotation
𝜃 wr of the wheel about its Y w axis, two Euler angles 𝜓 wr and 𝜑wr are used to first write the
transformation matrix Atw of this coordinate system in the trajectory wheel coordinate sys-
tem. Using this transformation matrix, the orthogonal transformation matrix that defines
the orientation of the intermediate wheel coordinate system in the global system can be
written as (Shabana and Rathod 2007; Shabana et al. 2008)
[ ]
Awi = Ar Atw Az Ax = a1wi a2wi a3wi (5.65)
where Ar is the transformation matrix that defines the orientation of the rail body coordi-
nate system in the global coordinate system, Atw = Atw (sw ) is the transformation matrix that
defines the orientation of the trajectory wheel coordinate system X tw Y tw Z tw with respect to
the rail body coordinate system X r Y r Z r , and
⎡cos 𝜓 wr − sin 𝜓 wr 0⎤ ⎡1 0 0 ⎤
⎢ ⎥ ⎢ ⎥
Az = ⎢ sin 𝜓 wr cos 𝜓 wr 0⎥ , Ax = ⎢0 cos 𝜑wr − sin 𝜑 ⎥
wr (5.66)
⎢ ⎥ ⎢ ⎥
⎣ 0 0 1⎦ ⎣0 sin 𝜑wr cos 𝜑wr ⎦
Using the preceding two equations, one can write
⎡cos 𝜓 wr − sin 𝜓 wr cos 𝜑wr sin 𝜓 wr sin 𝜑wr ⎤
⎢ ⎥
Awi = Ar Atw ⎢ sin 𝜓 wr cos 𝜓 wr cos 𝜑wr − cos 𝜓 wr sin 𝜑wr ⎥ (5.67)
⎢ ⎥
⎣ 0 sin 𝜑wr cos 𝜑wr ⎦
If absolute Cartesian coordinates are used as the generalized coordinates, transformation
matrix Awi can be obtained using the following simple matrix multiplication
Awi = Aw ATy (5.68)
where Aw is the transformation matrix that defines the orientation of the wheel coordinate
system X w Y w Z w in the global coordinate system XYZ, and Ay is the matrix that accounts
for the pitch rotation about the wheel Y w axis and is defined as
⎡ cos 𝜃 wr 0 sin 𝜃 wr ⎤
⎢ ⎥
Ay = ⎢ 0 1 0 ⎥ (5.69)
⎢ ⎥
⎣− sin 𝜃 wr 0 cos 𝜃 wr ⎦
The location of an arbitrary point on the rail space curve at arc length sr1 can be written as
( )
rro = Rr + Ar uro sr1 (5.70)
where Rr is the global position vector of the origin of the rail body coordinate system X r Y r Z r
( )
and uro sr1 is the position vector of the arbitrary point with respect to the rail body coor-
dinate system, as shown in Figure 9. The arc length sr1 at which vector rro of Eq. 70 has
a zero component along the X wi axis of the wheel can be determined using the equation
T ( )
a1wi rro − Rw = 0, where Rw is the global position vector of the origin of the wheel body
coordinate system X w Y w Z w and a1wi , which is the first column of transformation matrix Awi ,
represents a unit vector along the X wi axis of the intermediate wheel coordinate system. One
can therefore write
T ( )
a1wi Rr + Ar uro − Rw = 0 (5.71)
218 Mathematical Foundation of Railroad Vehicle Systems

r
Z
r
Y
r
X
O uro

r
s1

Figure 5.9 Planar contact.

In the case of arbitrary track geometry, this equation is a nonlinear function of rail
arc length sr1 . Because the wheel and rail generalized coordinates are assumed known,
Eq. 71 can be solved iteratively using a Newton–Raphson
( T ) algorithm to determine sr1 .
T ( )
In this algorithm, the algebraic equation a1wi Ar tr1 Δsr1 = −a1wi Rr + Ar uro − Rw
is iteratively solved for the Newton difference Δsr1 , where tr1 = 𝜕uro∕𝜕sr1 is the longi-
tudinal tangent to the rail space curve defined in the rail body coordinate system
r
X r Y r Z r . Therefore, the Newton(difference) Δs1 can be determined using the equation
wiT
( r ) wiT r r
Δs1 = −a1 R + A uo − R ∕ a1 A t1 . The Newton difference is used to update sr1
r r r w

until convergence is achieved. This Newton–Raphson iterative procedure is an alternate to


using the first-order ordinary differential equation ṡ r1 = ṙ wQ tr1 , which was introduced previ-
T

ously in this section to determine sr1 in the nodal search method. Knowing sr1 and assuming
sw2 = 𝜃 wr , the following contact conditions can be applied for the wheel/rail contact
T ( r )
tr2 R + Ar ur − Rw − Awi uwi = 0 ⎫
T ( ) ⎪
nr Rr + Ar ur − Rw − Awi uwi = 0⎬ (5.72)
T ⎪
nr tw1 = 0 ⎭

where the vectors and matrices that appear in this equation are as previously defined.
Knowing the generalized coordinates of the wheel and rail, and knowing sw2 and sr1 , the
first and third equations in the preceding equation can be used to solve for the two profile
parameters sw1 and sr2 . Enforcing the second equation in the preceding equation eliminates
the freedom of the wheel to translate with respect to the rail along the normal to the contact
surfaces. In the elastic contact formulation, the second equation can be used to determine
the penetration that enters into calculating the compliant normal contact force model.
In the planar contact formulation discussed in this section, it is assumed that the surface
parameters sw2 = 𝜃 wr and sr1 are known and fixed when the preceding equations are solved
iteratively to determine sw1 and sr2 . Consequently, the planar contact method described in
this section does not ensure that the two points on the wheel and the rail coincide, and
these two points may differ by a small longitudinal shift. This is mainly because Eq. 71,
T ( )
a1wi Rr + Ar uro − Rw = 0, is used to determine sr1 separately. To ensure that there is no
longitudinal shift between the two points on the wheel and the rail, the following three
equations can be used to iteratively solve for the three parameters sw1 , sr1 , and sr2 and account
Contact Problem 219

for the kinematic coupling between these surface parameters (Shabana and Rathod 2007):
T ( )
tr1 Rr + Ar ur − Rw − Awi uwi = 0⎫
T ( ) ⎪
tr2 Rr + Ar ur − Rw − Awi uwi = 0⎬ (5.73)
T
nr tw1 = 0 ⎪

In the constraint contact formulation, the second equation in Eq. 72,
T ( )
nr Rr + Ar ur − Rw − Awi uwi = 0, can be enforced to eliminate the freedom of the
wheel to travel with respect to the rail along the normal to the contact surfaces. In this
case, the normal contact force is determined as a constraint (reaction) force, as discussed
in Chapter 6. In the elastic contact formulation, on the other hand, the second equation in
T ( )
Eq. 72 can be used to compute the penetration as 𝛿 = nr Rr + Ar ur − Rw − Awi uwi . This
penetration can be used to compute the normal contact force. In both the constraint and
elastic contact formulations, using Eq. 73 to solve iteratively for sw1 , sr1 , and sr2 ensures that
there is no longitudinal shift and the two contact points on the wheel and rail coincide.

5.10 MAGLEV FORCES

The focus of this chapter has mainly been on the analysis of the wheel/rail contact prob-
lem. While wheels and rails remain the primary means of generating the traction forces that
move most freight and passenger trains, wheel/rail contact is the source of many problems,
including the following: (i) train speeds are limited due to wheel/rail contact; (ii) because
of wheel and rail contact, maintenance costs resulting from wear are very high – measured
in billions of dollars annually for a single freight company in North America; (iii) vibra-
tion and noise levels can be very high; (iv) the suspension system and bogie structures of
wheel/rail trains are not simple and require many components including axles, bearings,
springs, absorbers, etc.; (v) using wheels and rails means the possibility of derailments and
other accidents is high; (vi) the pantograph/catenary contact that provides the electricity
required for operating high-speed passenger trains can also limit speed, be a source of wear,
and lead to high maintenance costs and frequent inspections; and (vii) wheel/rail contact
gives rise to creep contact forces that make it difficult to develop virtual models and com-
puter programs that perform efficient simulations and virtual prototyping. As previously
discussed in this book, wheel/rail creep forces have a direct effect on critical vehicle speeds.

Maglev and Wheel/Rail Trains To avoid the problems that arise as the result of using
wheel/rail contact to move trains, magnetically levitated (maglev) trains have received atten-
tion over the past half-century. These trains, which do not normally rely on wheels, are
floated by magnetic forces; therefore, the vehicles are not in contact with the guideway.
Maglev trains were first proposed in the early 1930s but were first implemented in 2003
in Shanghai, China. They have several desirable features including higher speeds, lower
maintenance costs, less vibration and noise, and light weight. But despite these features,
maglev trains, while operating in some countries, are still, for the most part, at the experi-
mental stage and are currently used mainly for short distances rather than travel between
cities hundreds of miles apart. Their use for freight transportation is also still doubtful,
220 Mathematical Foundation of Railroad Vehicle Systems

particularly in the case of very long trains that consist of many cars with very high axle
loads.
Using maglev trains for long travel distances is debated due to safety concerns and the
economic advantages of air transportation, which does not require building an expensive
infrastructure on the ground. Although Japan approved a maglev train line for inner-city
transportation to be completed by 2027, for the foreseeable future, using maglev trains as a
primary freight and passenger transportation method for long distances remains in doubt;
it is likely that such maglev technology will be used, for the most part, for relatively short
distances.
Nonetheless, the maglev non-contact concept eliminates many of the problems resulting
from wheel/rail contact and allows for a significant increase in train speed. For example,
it is expected that train speed will exceed 600 km/h, and the noise level is reduced from
75–80 dB for wheel/rail trains to 60–65 dB for maglev trains (Lee et al. 2006). Furthermore,
because there is no wheel/rail contact or friction in the case of a maglev train and no
need for a pantograph/catenary system, many of the wear problems that result from more
concentrated contact loads can be eliminated or reduced, leading to lower maintenance
costs, environmentally superior transportation, and improved ride comfort (Dukkipati
2000; Lee et al. 2006).

Maglev Train Suspensions Unlike conventional trains, which are driven by wheel/rail
interaction forces, in maglev trains, a magnetic levitation force separates the vehicle from
the guideway. This magnetic levitation force serves as the vehicle’s suspension system. The
train is driven by propulsion forces generated by a linear motor instead of the rotary motor
used in conventional train systems. There are three types of maglev suspension systems:
electrodynamic suspension (EDS), electromagnetic suspension (EMS), and hybrid electromag-
netic suspension (HEMS). These three maglev suspension systems are briefly described next
(Lee et al. 2006):
● Electrodynamic suspension (EDS): In an EDS system, shown in Figure 10a, repulsive mag-
netic forces are used for levitation. Magnets are attached to the vehicle, and inducing coils
or conducting sheets are placed on the guideway. As the magnets attached to the vehicle

Vehicle Vehicle

Secondary
suspension

Electrodynamic
suspension
Electromagnetic
(a) (b) suspension

Figure 5.10 Maglev suspension.


Contact Problem 221

move forward, current flows in the guideway induced coils or conducting sheets, gener-
ating the magnetic field required to produce the repulsive force that lifts the vehicle. This
type of suspension is stable and does not normally require an active control system to con-
trol the air gap between the vehicle and guideway; this air gap is approximately 100 mm.
Because this maglev suspension system has good stability characteristics, it is suited for
high-speed operation and heavy loads. However, because the repulsive magnetic levita-
tion forces are the result of vehicle motion above the inducing coils or conducting sheets,
the vehicle must reach a certain speed (approximately 100 km/h) in order to generate
the electric current required to produce the levitation forces. For this reason, when the
electrodynamic suspension is used, rubber tires are used at lower speeds; therefore, this
suspension system is not suitable for urban transportation that requires frequent stops
and typically a low operation speed.
● Electromagnetic suspension (EMS): In the EMS system, shown in Figure 10b, attraction
forces instead of EDS repulsive forces produce magnetic levitation. The attraction forces
are produced using electromagnets attached to the vehicle and ferromagnets placed on
the guideway. The iron core of a magnetic circuit is excited by an electric current-carrying
coil, and as a result, the core on the vehicle is attracted to the ferromagnetic rail. As dis-
cussed in this section, the electromagnetic suspension has stability problems because
the magnetic levitation attraction force increases as the small air gap (approximately
±10 mm) between the pole face of the electromagnet on the vehicle and the ferromagnet
on the guideway decreases. Therefore, a feedback control system is required to ensure
the stability of such a suspension system. Nonetheless, unlike the EDS force, the EMS
levitation magnetic force does not depend on vehicle speed, and the vehicle can be lifted
at low or zero speed. For this reason, this suspension system can be used for urban mass
transportation that requires frequent stops and a relatively low operation speed.
● Hybrid electromagnet suspension (HEMS): In the HEMS systems, permanent magnets are
used with the electromagnets in some EMS systems to reduce the electric power required
for magnetic levitation. For a given nearly constant air gap, permanent magnets can be
designed to provide the required levitation forces without the need for electromagnets
that produce a force-gap relationship that can be a source of instability. Using permanent
magnets ensures that the gap remains nearly constant, reduces reliance on an active con-
trol system to control the gap, and leads to a significant reduction in the electric power
required for magnetic levitation. Nonetheless, the HEMS system requires a greater varia-
tion in the amplitude of the electric current as compared to EMS because the permanent
magnets have the same permeability as air (Lee et al. 2006).

Electromagnetic Suspension Forces This section formulates the magnetic levita-


tion force produced by the EMS system in order to illustrate the force-gap relationship
when such a suspension system is used in maglev trains. As previously mentioned, when
EMS is used, attraction forces lift the vehicle instead of the repulsive forces produced by
EDS. The force of attraction between the pole surface and the ferromagnet can be written
as (Sinha 1987)
( )2
Φp Ap
Fz = (5.74)
Pp
222 Mathematical Foundation of Railroad Vehicle Systems

where Ap = lp wp ; lp and wp are, respectively, the length and width of the pole face, as shown
in Figure 11; Pp is the permeability of the free space; and Φp is the flux density across the
air gap. The flux density can be written in terms of the reluctance of the mutual flux RM
and the magneto-motive force F m as Φp = F m /Ap RM . The magneto-motive force F m can
be written in terms of the number of turns of coil nco and coil electric current i = i(t) as
F m = nco i(t). Therefore, the flux density Φp can be written as Φp = nco i(t)/Ap RM . Substituting
this equation into Eq. 74, one obtains
( )2
1 nco i (t)
Fz = (5.75)
Pp Ap RM
The reluctance of mutual flux RM consists of the reluctance of the air gap 2dz /Pp lp wp , elec-
tromagnet (wl + 2h + 2lp )/Pe lp wp , and ferromagnet (wl + 2lp )/Pf h1 wp , where dz = dz (t) is the
distance between the pole faces and the ferromagnet, h1 is the thickness of the ferromagnet;
wl and h are the dimensions shown in Figure 11; and Pe and Pf are, respectively, the per-
meability of the electromagnet and the ferromagnet. Therefore, one can write the equation
for the reluctance of mutual flux RM as (Dukkipati 2000)
( ) ( ) ( )
( ) 2dz wl + 2h + 2lp wl + 2lp
R M dz = + + (5.76)
Pp lp wp Pe lp wp Pf h1 wp

If assumptions are made that Pe = Pf and h1 ≃ lp , the preceding equation reduces to


( ) 2 ( )
R M dz ≃ dz (t) + cp (5.77)
Pp Ap
where cp = (Pp /Pe )(2lp + wl + h). Substituting Eq. 77 into Eq. 75 leads to
( )2 ( )2
( ) Pp nco Ap i (t) ( )
Fz dz , i = ≤ Fz max (5.78)
4 dz (t) + cp

Force (F z )max , which is associated with the saturation state of the flux density Φp across the
air gap, can be written as (F z )max = ((Φp )max Ap )/Pp , where (Φp )max is the maximum flux den-
( ) ( ( ))
sity defined as Φp max = PM ∕ PM + PL Φp , PM = 1/RM is the mutual flux permeance,

Ferromagnet

Fz(t)
tp
dz(t) wl
lp lp
h

wp
i(t)
v(t)
Electromagnet

Figure 5.11 Electromagnetic suspension forces.


Contact Problem 223

PL = Pp h((wp /wl ) + ln(1 + (𝜋lp /2wl ))) is the permeance of the leakage flux, and Φp is in the
range 1.5–2 Wb/m2 (Dukkipati 2000).
The form of the magnetic levitation force of Eq. 78 sheds light on some of the dynamics
and stability issues that need to be considered in the design of maglev trains. It is clear that
force F z is a quadratic function of the electric current; therefore, it increases as the electric
current increases, as shown in Figure 12a. force F z , on the other hand, is inversely pro-
portional to (dz )2 ; therefore, magnetic levitation force F z decreases as air gap dz increases.
This behavior, which is shown in Figure 12b, leads to instability. To avoid this instability,
an active feedback control system or, alternatively, HEMS system may be required in order
to ensure safe operation of a maglev train.

dz1 d
z2 dz3

i1 < i2 < i3
Fz

Fz
i3
dz1 <dz2 <dz3 i2
i1
i dz

Figure 5.12 Relationship between levitation force, current, and air gap.

Having determined the magnetic levitation force F z in the guide coordinate system, this
force can be used to define the magnetic levitation force vector in any other coordinate
system, including the global coordinate system, by using the proper coordinate transforma-
tion. If Fml is the magnetic levitation force defined in the global coordinate system, one can
systematically use this force vector, as described in Chapter 6, to determine the generalized
forces associated with the generalized coordinates Rk and 𝜽k , k = v, g, where subscripts
v and g refer, respectively, to the vehicle and guideway. These generalized forces can be
introduced to the constrained multibody system (MBS) dynamic equations developed in the
following chapter. The procedure for developing the expressions of the generalized forces
and equations of motion is described in detail in Chapter 6.

Summary of Maglev Technology Concerns To place more reliance on maglev


transportation, more research and technology advances are needed. Several issues and
concerns need to be addressed, as follows:
● Using non-contact elements between the vehicle and guideway eliminates the options of
using passive elements such as springs and dampers to control the dynamics of maglev
systems. Therefore, such systems may require implementing more expensive feedback
control systems to ensure safety.
● As the weight and number of rail cars increase, more electric power is needed to pro-
duce the required levitation forces to support the increased loads. Therefore, the existing
maglev technology is not suited for long freight trains, which often have up to 200 cars
with high axle loads.
224 Mathematical Foundation of Railroad Vehicle Systems

● Changing train directions, as in the case of switching, can be difficult due to the guideway
and magnetic levitation.
● The effect on passengers of the magnetic field generated by strong magnets needs to be
evaluated. As the levitation force increases, strong magnets are required, and this in turn
leads to a stronger magnetic field whose effect cannot be ignored.
● As a result of not using wheel/rail contact in maglev trains, traction motors must be
designed to provide both propulsion and braking forces using the electromagnetic forces.
Some of these concerns are currently being researched to determine the feasibility of devel-
oping a maglev train system that can be efficiently used for mass transportation, ensures
safe operation, is economically feasible, and is suited for long-distance travel. But because of
the high axle load of freight trains, the focus of this research is on passenger transportation,
while freight transportation will continue to rely on wheel/rail contact for the foreseeable
future.
225

Chapter 6

EQUATIONS OF MOTION

While different approaches can be used to formulate the dynamic equations of motion
of constrained dynamical systems, the Lagrangian approach lends itself to easily devel-
oping general algorithms for the computer-aided analysis of railroad vehicle systems. In
the Lagrangian approach, the concepts of virtual displacement, virtual work, generalized
coordinates, and generalized forces are fundamental. As explained in Chapter 3, the general
unconstrained spatial motion of a rigid body is described using six independent coordi-
nates. Three of these coordinates describe the translation of the body reference point, and
three rotation parameters define the body orientation. The orientation coordinates can be
three Euler angles or four Euler parameters, as discussed in Chapter 3. In this chapter, the
Cartesian translation and orientation coordinates are referred to as the absolute generalized
coordinates, which define, respectively, the body translation and orientation with respect to
a selected global coordinate system. Another set of coordinates that is also used in formu-
lating the dynamic equations of railroad vehicle systems is the set of trajectory coordinates,
which includes coordinates that define the body translation and orientation with respect
to a body-track coordinate system that follows the body motion. This body-track coordinate
system is referred to in this chapter as the trajectory body coordinate system.
In this chapter, the generalized coordinates introduced in Chapter 3 are used to define
the global position of the origin and the orientation of the body coordinate system. Because
the formulations presented in this chapter are based on the Newton–Euler equations of
motion, the body reference point, which defines the origin of the body coordinate system,
is assumed to be attached to the body center of mass in order to eliminate the inertia cou-
pling between the body translation and rotation. The virtual work principle used in the
Lagrangian formulation is introduced and used with the expressions for the absolute veloc-
ity and acceleration vectors of an arbitrary point on the body, obtained in Chapter 3, to
formulate the body equations of motion. This chapter also explains using contact condi-
tions to formulate the wheel/rail interaction forces and discusses hunting oscillations using
a constrained wheelset model that accounts for coupling between the lateral and yaw dis-
placements.

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
226 Mathematical Foundation of Railroad Vehicle Systems

6.1 NEWTONIAN AND LAGRANGIAN APPROACHES


Two approaches are commonly used to formulate the constrained dynamic equations of
motion of multibody systems (MBS) whose components are connected by mechanical
joints: the Newtonian approach and the Lagrangian approach. The Newtonian approach,
which is based on vector mechanics, requires the use of free-body diagrams by making
cuts at the joints. These free-body diagrams show the joint reaction forces as well as the
inertia and applied forces. This Newtonian approach is not well-suited for the development
of general MBS algorithms and can be used to develop the dynamic equations of motion of
relatively simple systems. For the analysis of complex systems, such as railroad vehicles,
the Lagrangian approach, which has its roots in D’Alembert’s principle and employs scalar
quantities such as the virtual work and kinetic and potential energies, is often used to
develop general-purpose MBS algorithms. When the Lagrangian approach is used, there is
no need to make free-body diagrams or study the equilibrium of the bodies separately. This
is mainly because the Lagrangian approach is based on the connectivity conditions. The
connectivity conditions, which describe mechanical joints in the system, can be formulated
using a set of nonlinear algebraic constraint equations. In the Lagrangian approach, the
joint forces take a standard form expressed in terms of the constraint Jacobian matrix
and multipliers, called Lagrange multipliers. The fact that algebraic constraint equations
can be used to systematically define joint reaction forces allows for developing general
computation procedures for the computer-aided analysis of a wide class of physics and
engineering systems.
When the Lagrangian approach is used, the constraint forces can be kept in the formu-
lation or can be systematically eliminated. Writing the equations of motion in terms of
constraint forces leads to large, sparse-matrix equations of motion. The solution for the
accelerations and constraint forces can be obtained by combining the equations of motion
with the constraint equations at the acceleration level. This approach, in which redundant
coordinates are used to formulate the dynamic equations, is called the augmented formula-
tion in the MBS literature.
Alternatively, the Lagrangian approach, which is based on the Lagrange–D’Alembert
principle, can also be used with an embedding technique to eliminate the constraint forces.
In this case, constraint equations are used to write dependent accelerations in terms of
independent accelerations using a velocity transformation matrix. Using this approach, the
constraint forces and dependent coordinates can be systematically eliminated, leading to
a number of equations of motion equal to the number of independent coordinates, which
are called the degrees of freedom of the system. Since the independent coordinates are
not related by constraint equations, the embedding technique leads to a minimum set of
equations from which the constraint forces are eliminated.
In the Lagrangian approach, the concept of generalized coordinates is fundamental. Gen-
eralized coordinates in the MBS literature are the coordinates used to define the system
configuration and formulate the generalized forces. Therefore, generalized coordinates do
not have to be independent coordinates, as was assumed in classical mechanics literature
(Greenwood 1988). The independent coordinates are called the system degrees of freedom.
Given a constrained dynamical system, the configuration of the system can be defined
using the generalized coordinates q = [q1 q2 … qn ]T , where n is the total number of
Equations of Motion 227

generalized coordinates. The vector of generalized coordinates can, in general, include both
independent and dependent coordinates, and therefore, one can write this vector in a par-
titioned form as
[ ]T
q = qTi qTd (6.1)

In this equation, qi is the vector of independent coordinates, and qd is the vector of depen-
dent coordinates, with the dependency arising from motion constraints due to mechanical
joints and specified motion trajectories, as explained in this chapter. If no constraints are
imposed on the motion of the system, all the coordinates become independent because they
are not related by algebraic equations.

6.2 VIRTUAL WORK PRINCIPLE AND CONSTRAINED


DYNAMICS

The virtual work principle allows for systematic derivation of the dynamic equations of
motion of complex railroad vehicle systems. In this section, the principle of virtual work is
derived using a system of particles. If a rigid body is considered to consist of a large number
of particles, the same principle can also be applied to rigid body systems.

Virtual Displacement and Generalized Forces The virtual work principle is based
on the concept of virtual displacement, which refers to a change in the configuration of the
system as the result of an arbitrary infinitesimal change of the coordinates, consistent with
the forces and constraints imposed on the system at a given time instant t. Consider a system
of np particles. The position vector of a particle i, denoted as ri , can be written in terms of
[ ]T
the system generalized coordinates q = q1 q2 … qn as
( )
ri = ri (q, t) = ri q1 , q2 , … , qn , t , i = 1, 2, … , np (6.2)

At the position level, it is not always possible to write position coordinates in terms of gen-
eralized coordinates using closed-form linear relationships. The virtual displacement, on
the other hand, allows writing the virtual change in position vectors in terms of the virtual
change in generalized coordinates using linear equations. A virtual change in the position
vector of Eq. 2 can be written as
𝜕ri 𝜕ri 𝜕ri
𝛿ri = 𝛿ri (q) = 𝛿q1 + 𝛿q2 … + 𝛿q
𝜕q1 𝜕q2 𝜕qn n
∑n
𝜕ri
= 𝛿q , i = 1, 2, … , np (6.3)
j=1
𝜕qj j

In this expression of virtual displacement, no differentiation is made for time t because the
virtual change is assumed to occur at a given time instant.
Using the definition of virtual displacement, the virtual work 𝛿W i of a force Fi acting on
particle i is defined as the dot product of the force vector with the virtual change in the
position vector of the point of application of the force. Therefore, the virtual work of force
228 Mathematical Foundation of Railroad Vehicle Systems

vector Fi can be written as


( i )
T T 𝜕r 𝜕ri 𝜕ri
𝛿W i = Fi 𝛿ri = Fi 𝛿q1 + 𝛿q2 … + 𝛿qn
𝜕q 𝜕q2 𝜕qn
( ) 1
∑ n
T 𝜕r
i ∑n
= Fi 𝛿qj = Qj 𝛿qj (6.4)
j=1
𝜕qj j=1

where Qj = (𝜕ri /𝜕qj ) T Fi is called the generalized force associated with generalized coordi-
nate qj . It is clear from the preceding equation that the Cartesian space used in the New-
tonian approach is replaced in the Lagrangian approach by a generalized space that has
dimension n.

Generalized Inertia and Applied Forces Newton’s second law is used as the start-
ing point in developing the principle of virtual work for a system of particles. This law of
motion states that the result of forces acting on a particle is equal to the rate of change of
momentum of this particle. This statement can be written mathematically as Fi = Ṗ i or,
equivalently, Fi − Ṗ i = 𝟎, where Pi = mi ṙ i is the linear momentum of particle i and Fi is the
vector of the resultant forces applied to the particle. Force vector Fi can be written as the
sum of two vectors Fi = Fie + Fic , where Fie is the vector of external forces and Fic is the vector
of constraint forces. It follows that mi r̈ i − Fie − Fic = 𝟎, with the assumption that the mass of
the particle mi remains constant. Multiplying the equation mi r̈ i − Fie − Fic = 𝟎 by the virtual
displacement 𝛿ri , one obtains
( i i )T
m r̈ − Fie − Fic 𝛿ri = 0, i = 1, 2, … , np (6.5)

This equation leads to


n
∑p
( i i )T
m r̈ − Fie − Fic 𝛿ri = 0 (6.6)
i=1

This equation, called the Lagrange–D’Alembert equation, can be written as 𝛿W i − 𝛿W e −


𝛿W c = 0, where
np np np
∑ ( ) ∑ ∑
i i T T T
𝛿Wi = m r̈ 𝛿r ,
i
𝛿We = Fie 𝛿ri , 𝛿Wc = Fic 𝛿ri (6.7)
i=1 i=1 i=1

In this equation, 𝛿W i is the virtual work of the inertia forces, 𝛿W e is the virtual work of the
external forces, and 𝛿W c is the virtual work of the constraint forces. Because joint reaction
forces between two particles are equal in magnitude and opposite in direction, and because
the virtual change in a specified coordinate is zero, one has 𝛿W c = 0, and the principle of
virtual work in dynamics can be written as 𝛿W i − 𝛿W e = 0 or

𝛿Wi = 𝛿We (6.8)


(
∑n ∑np ( i i )T )
Substituting Eqs. 3 and 7 into Eq. 8, one obtains j=1 ̈ − Fie 𝜕ri ∕𝜕qj 𝛿qj = 0,
i=1 m r
which can be written as
n (
∑ ( ) ( ))
Qi j − Qe j 𝛿qj = 0 (6.9)
j=1
Equations of Motion 229

( ) ∑np ( i i )T i ( ) ∑np i T i
where Qi j = i=1 m r̈ 𝜕r ∕𝜕qj and Qe j = i=1 Fe 𝜕r ∕𝜕qj are, respectively, the gen-
eralized inertia and generalized external forces associated with the generalized coordinate
qj . If the coordinates q1 , q2 , …, qn are independent, Eq. 9 defines the second-order ordinary
differential equations of the system as
( ) ( )
Qi j = Qe j , j = 1, 2, … , n (6.10)
[( ) ( )
This equation can also be written in vector form as Qi = Qe , where Qi = Qi 1 Qi 2 …
( ) ]T [( ) ( ) ( ) ]T
Qi n and Qe = Qe 1 Qe 2 … Qe n .

Constrained Dynamics In railroad vehicle system dynamics, the coordinates q1 , q2 , …,


qn may not be independent because of mechanical joints and specified motion trajectories.
Examples of mechanical joints that impose restrictions on the system motion were provided
in Chapter 3. In this case of constrained dynamics, the system coordinates are related by a
set of algebraic constraint equations, which can be nonlinear functions of the coordinates
and can be written in a vector form as
[ ]T
C (q, t) = C1 C2 … Cnc = 𝟎 (6.11)

where nc is the number of constraint functions, which must be less than or equal to
the number of coordinates n: that is, nc ≤ n. If nc = n, the system is referred to as a
kinematically driven system. For a virtual change in the coordinates, the preceding equation
leads to
⎡ 𝜕C1 ∕𝜕q ⎤
⎢ ⎥
( ) ⎢ 𝜕C2 ∕𝜕q ⎥
𝜕C
𝛿C = 𝛿q = Cq (q, t) 𝛿q = ⎢ ⎥ 𝛿q = 𝟎 (6.12)
𝜕q ⎢ ⋮ ⎥
⎢ ⎥
⎢𝜕C ∕𝜕q⎥
⎣ nc ⎦
In this equation, Cq = 𝜕C/𝜕q is the constraint Jacobian matrix defined as

⎡ 𝜕C1 ∕𝜕q ⎤ ⎡ 𝜕C1 ∕𝜕q1 𝜕C1 ∕𝜕q2 … 𝜕C1 ∕𝜕qn ⎤


⎢ ⎥ ⎢ ⎥
⎢ 𝜕C2 ∕𝜕q ⎥ ⎢ 𝜕C2 ∕𝜕q1 𝜕C2 ∕𝜕q2 … 𝜕C2 ∕𝜕qn ⎥
Cq (q, t) = ⎢ ⎥=⎢ ⎥ (6.13)
⎢ ⋮ ⎥ ⎢ ⋮ ⋮ ⋱ ⋮ ⎥
⎢ ⎥ ⎢ ⎥
⎢𝜕C ∕𝜕q⎥ ⎢𝜕C ∕𝜕q 𝜕C ∕𝜕q … 𝜕Cnc ∕𝜕qn ⎥⎦
⎣ nc ⎦ ⎣ nc 1 nc 2

and 𝜕Ck /𝜕q, k = 1, 2, …nc , is a row vector. The number of rows in the constraint Jacobian
matrix Cq is equal to the number of constraint functions nc , and the number of columns is
equal to the number of coordinates n. The constraint functions must be linearly indepen-
dent, and therefore, the constraint Jacobian matrix is assumed to have a full row rank. That
is, one must be able to identify an nc × nc non-singular square matrix formed by nc columns
of the constraint Jacobian matrix Cq . This nc × nc non-singular square matrix allows, in the
Lagrangian dynamics, writing the dependent variables in terms of the independent vari-
ables or degrees of freedom.
230 Mathematical Foundation of Railroad Vehicle Systems

Example 6.1 As discussed in Chapter 3, a spherical (ball) joint eliminates all the
degrees of freedom of the relative translation between two bodies connected by
this joint. Because this joint eliminates three translational degrees of freedom, it is
formulated using three algebraic constraint equations. It was shown in Chapter 3 that
the constraint equations of the spherical joint between two arbitrary bodies i and j
connected by the joint at points Pi and Pj , respectively, can be written as
j
riP − rP = Ri + Ai uPi − Rj − Aj uPj = 𝟎
where rkP , Rk , Ak and uPk , k = i, j, are, respectively, the global position vector of the joint
definition point Pk , the global position vector of the reference point of body k, the trans-
formation matrix that defines the body orientation, and the local position vector of point
Pk with respect to the body coordinate system. In the formulation of the spherical joint,
[ T T ]T
the absolute Cartesian coordinates qk = Rk 𝛉k introduced in Chapter 3 are used.
As shown in Chapter 3, the derivatives of vector rkP with respect to the orientation
parameters 𝛉k of body k can always be written as 𝜕rkP ∕𝜕𝛉k = −Ak ũ Pk Gk , where ũ Pk is the
skew-symmetric matrix associated with vector uPk , and Gk is the matrix that relates
the angular velocity vector 𝛚k to the time derivatives of the orientation parameters:
that is, 𝛚k = Gk 𝛉̇ k . Therefore, the constraint Jacobian matrix of the spherical joint can
be written as
[ ] [ ]
𝜕C 𝜕C 𝜕C 𝜕C 𝜕C 𝜕C
Cq = =
𝜕qi 𝜕qj 𝜕Ri 𝜕𝛉i 𝜕Rj 𝜕𝛉j
[ ]
= I −Ai ũ PiGi −I Aj ũ PjGj
This matrix has three rows and six columns if Euler angles are used, and it has three
rows and seven columns if Euler parameters are used.

Example 6.2 As discussed in Chapter 3, a cylindrical joint is a two-degree-of-freedom


joint that allows relative translation and rotation between two bodies only along the joint
axis. Therefore, a cylindrical joint is formulated mathematically using four algebraic
constraint equations. The four constraint equations of the cylindrical joint between two
bodies i and j can be written in terms of the coordinates of the two bodies, as shown in
Chapter 3, as
( ) [ T ]T
C qi , qj = v1i vj v2i vj v1i rijP v2i rijP = 𝟎
T T T

ij j j
where rP = riP − rP ; riP and rP are, respectively, the global position vectors of two points Pi
and Pj on bodies i and j defined on the joint axis; vj = Aj vj is a vector on body j along the
joint axis; and v1i = Ai vi1 and v2i = Ai vi2 are two vectors defined on body i perpendicular to
the joint axis. In making these definitions, Ak is the matrix that defines the orientation of
the body coordinate system, and vk is a constant vector defined in the coordinate system
of body k, k = i, j. The constraint Jacobian matrix of the cylindrical joint can be written
Equations of Motion 231

as
[ ] [ ]
𝜕C 𝜕C 𝜕C 𝜕C 𝜕C 𝜕C
Cq = =
𝜕qi 𝜕qj 𝜕Ri 𝜕𝛉i 𝜕Rj 𝜕𝛉j
−vj Ai ṽ i1Gi −v1i Aj ṽ j Gj ⎤
T T
⎡ 𝟎 𝟎
⎢ ⎥
−vj Ai ṽ i Gi −v2i Aj ṽ j Gj ⎥
T T
⎢ 𝟎 2 𝟎
=⎢ T ij T ⎥
−v1i Ai ũ PiGi rP Ai ṽ i1Gi v1i Aj ũ PjGj ⎥
i T T T
⎢v1 − −v1i
⎢ iT ij T ⎥
−v2i Ai ũ PiGi − rP Ai ṽ i2Gi vi Aj ũ j Gj ⎦
T T T
⎣v2 −v2i 2 P

This is a 4 × 6 matrix when Euler angles are used and a 4 × 7 matrix when Euler param-
eters are used.

Lagrange Multipliers Because Cq 𝛿q = 0, multiplying this nc -dimensional vector by any


other arbitrary nc -dimensional vector 𝛌 leads to 𝛌T Cq 𝛿q = 0. This equation can be added to
Eq. 9, which can be written as (Qi − Qe )T 𝛿q = 0 to yield
( )T
Qi − Qe + CTq 𝛌 𝛿q = 0 (6.14)
Because the generalized coordinates q1 , q2 , …, qn are not independent due to the constraint
( )
equations of Eq. 11, the coefficients Qi − Qe + CTq 𝛌 k of 𝛿qk , k = 1, 2, …, n, in the pre-
ceding equation cannot be assumed zero unless additional conditions are imposed on the
meaning of the arbitrary vector 𝛌. Using the coordinate partitioning of Eq. 1, one can write
[ ]T
𝛿q = 𝛿qTi 𝛿qTd , where qd is an nc -dimensional vector, while qi is an (n − nc )-dimensional
vector. According to this coordinate partitioning, Eq. 14 can be written as
( )T ( )T
Qi − Qe + CTq 𝛌 d 𝛿qd + Qi − Qe + CTq 𝛌 i 𝛿qi = 0 (6.15)
In this equation, subscripts d and i refer to vector elements associated with the depen-
( ) ( )
dent and independent coordinates, respectively: that is, Qi − Qe + CTq 𝛌 𝛼 = Qi 𝛼 −
( )
Qe 𝛼 + CTq𝛼 𝛌, 𝛼 = i, d. Because the constraint Jacobian matrix Cq always has a full row
rank, the dependent coordinates qd can be selected such that the nc × nc constraint Jacobian
matrix Cqd associated with the dependent coordinates is non-singular. One can then select
( ) ( )
vector 𝛌 to be the solution of the system of algebraic equations Qi d − Qe d + CTqd 𝛌 = 𝟎:
that is,
( )−1 (( ) ( ) )
𝛌 = − CTqd Qi d − Qe d (6.16)
( )T
Using this condition, Eq. 15 reduces to Qi − Qe + CTq 𝛌 i 𝛿qi = 0. Since the elements of vec-
( ) ( ) ( )
tor qi are independent, one also has Qi − Qe + CTq 𝛌 i = Qi i − Qe i + CTqi 𝛌 = 𝟎, which
( ) ( )
when combined with Qi d − Qe d + CTqd 𝛌 = 𝟎 leads to
Qi − Qe + CTq 𝛌 = 𝟎 (6.17)
This vector equation, which has n scalar equations, is the constrained dynamical equation
of motion of the system formulated in terms of redundant coordinates that are not
totally independent. For a given system of forces and moments, the unknowns in this
equation are the n accelerations and the nc elements of vector 𝛌, which are called Lagrange
232 Mathematical Foundation of Railroad Vehicle Systems

multipliers. Since no coordinates are eliminated from Eq. 17, vector CTq 𝛌 represents the
generalized constraint forces. It is clear that generalized constraint forces are formulated
using connectivity conditions without the need for making cuts to form free-body dia-
grams. Equation 17, however, can be viewed as the generalized form of the Newton–Euler
equations obtained using free-body diagrams that include external and reaction forces.
Therefore, the Cartesian-based Newton–Euler equations can be considered a special
form of Eq. 17, which is more general because it is not restricted to the Cartesian-space
description.
By using redundant coordinates, the number of unknowns is n + nc in Eq. 17 when the
forces are given. These unknowns, which include n accelerations and nc Lagrange multi-
pliers, can be determined using Eqs. 11 and 17, which can be written as
}
Qi − Qe + CTq 𝛌 = 𝟎
[ ]T (6.18)
C (q, t) = C1 C2 … Cnc = 𝟎
This equation is a system of n + nc differential/algebraic equations (DAEs) with the differ-
ential equations representing equations of motion and the algebraic equations representing
constraint equations. In MBS dynamics, two different techniques can be used to solve Eq. 18
for the n accelerations and nc Lagrange multipliers or, equivalently, the generalized con-
straint forces. These two techniques, which are discussed in later sections of this chapter,
are the augmented formulation and the embedding technique.

6.3 SUMMARY OF RIGID-BODY KINEMATICS

The Newton–Euler equations used in rigid body dynamics can be systematically derived
using Newton’s law for particles and the kinematic equations developed in Chapter 3. By
assuming that a rigid body consists of a large number of particles, one can obtain an expres-
sion for the inertia forces of the rigid body. By equating the virtual work of the inertia forces
to the virtual work of the applied forces and using the assumption that the body reference
point is located at the body center of mass, the Newton–Euler equations can be developed
as demonstrated in the following section. Because of the assumption of a centroidal body
coordinate system, the Newton–Euler equations do not include inertia coupling between
the rigid body translation and rotation. Before deriving the Newton–Euler equations for
rigid bodies, a summary of the rigid body kinematic equations developed in Chapter 3 is
first presented.

Position Equations The basic rigid-body kinematic equations presented in Chapter 3


are used in this chapter to derive the Newton–Euler equations. As explained in Chapter 3,
the unconstrained spatial motion of a rigid body i in a railroad vehicle system is described
[ ]T
using six independent coordinates: three coordinates Ri = Rix Riy Riz define the transla-
tion of a selected reference point Oi on the body, and the other three coordinates 𝛉i define the
Equations of Motion 233

Zi
Yi

Oi
ui

Xi Pi
Z
Y Ri

ri

Figure 6.1 Rigid-body kinematics.

orientation of the body coordinate system X i Y i Z i with respect to the global coordinate sys-
tem XYZ. Using these coordinates, the global position of an arbitrary point on the rigid body
can be written as shown in Figure 1 as ri = Ri + Ai ui , where Ai is a 3 × 3 transformation
matrix, expressed in terms of the orientation parameters 𝛉i , that defines the orientation of
[ ]T
the body coordinate system X i Y i Z i in the global coordinate system; and ui = uix uiy uiz =
[ i i i ]T
x y z is the constant position vector of the arbitrary point with respect to the refer-
ence point Oi . As previously mentioned, in the Newton–Euler formulation, reference point
Oi is assumed to be the body center of mass. If Euler angles are used to define the body
[ ]T
orientation, one has 𝛉i = 𝜓 i 𝜙i 𝜃 i . By using the Euler-angle sequence Z i − X i − Y i , the
orthogonal transformation matrix Ai can be written as

⎡cos 𝜓 i cos 𝜃 i − sin 𝜓 i sin 𝜙i sin 𝜃 i − sin 𝜓 i cos 𝜙i cos 𝜓 i sin 𝜃 i + sin 𝜓 i sin 𝜙i cos 𝜃 i ⎤
⎢ ⎥
A = ⎢sin 𝜓 i cos 𝜃 i + cos 𝜓 i sin 𝜙i sin 𝜃 i cos 𝜓 i cos 𝜙i sin 𝜓 i sin 𝜃 i − cos 𝜓 i sin 𝜙i cos 𝜃 i ⎥
i

⎢ − cos 𝜙i sin 𝜃 i sin 𝜙i cos 𝜙i cos 𝜃 i ⎥


⎣ ⎦
(6.19)
Using this sequence of Euler rotations in railroad vehicle dynamics, the first rotation 𝜓 i
about the Z i axis is the yaw angle, the second rotation 𝜙i about the X i axis is the roll angle,
and the third rotation 𝜃 i about the Y i axis is the pitch angle.
As discussed in Chapter 3, using three independent parameters such as Euler angles
leads to singular configurations (Roberson and Schwertassek 1988; Shabana 2010). For
this reason, Euler parameters are often used in general MBS algorithms. In the case of
[ ]T
Euler parameters, the orientation coordinates are defined by vector 𝛉i = 𝜃0i 𝜃1i 𝜃2i 𝜃3i ,
∑3 ( )2
related by the algebraic relationship k=0 𝜃ki = 1. Because of this algebraic relationship,
the transformation matrix in terms of Euler parameters can assume different forms. One of
234 Mathematical Foundation of Railroad Vehicle Systems

these forms is
( )2 ( )2 ( ) ( )
⎡1 − 2 𝜃2i − 2 𝜃3i 2 𝜃1i 𝜃2i − 𝜃0i 𝜃3i 2 𝜃1i 𝜃3i + 𝜃0i 𝜃2i ⎤
i ⎢ ( i i ) ( ) 2 ( )2 ( ) ⎥
A = ⎢ 2 𝜃1 𝜃2 + 𝜃0i 𝜃3i 1 − 2 𝜃1i − 2 𝜃3i 2 𝜃2i 𝜃3i − 𝜃0i 𝜃1i ⎥ (6.20)
⎢ ( i i ) ( i i ) ( i )2 ( i )2 ⎥
⎣ 2 𝜃1 𝜃3 − 𝜃0 𝜃2
i i
2 𝜃2 𝜃3 + 𝜃0 𝜃1
i i
1 − 2 𝜃1 − 2 𝜃2 ⎦

Euler parameters have many identities that can be used to simplify kinematic and dynamic
equations. Some of these identities are provided in Chapter 3.

Velocity and Acceleration Equations The absolute velocity of an arbitrary point on


a rigid body can be written using one of the two forms ṙ i = Ṙ i + 𝛚i × ui and ṙ i = Ṙ i +
( )
Ai 𝛚i × ui , where Ṙ i is the absolute velocity vector of reference point Oi , ui = Ai ui , 𝛚i =
Ai 𝛚i is the absolute angular velocity vector defined in the global coordinate system, and
𝛚i is the absolute angular velocity vector defined in the body X i Y i Z i coordinate system.
As discussed in Chapter 3, the angular velocity vectors can be written in terms of the time
derivatives of the orientation coordinates as 𝛚i = Gi 𝛉̇ i and 𝛚i = Gi 𝛉̇ i . In the case of the Euler
[ ]T
angles 𝛉i = 𝜓 i 𝜙i 𝜃 i with the Z i , X i , Y i rotation sequence, matrices Gi and Gi are given,
respectively, by

⎡0 cos 𝜓 i − sin 𝜓 i cos 𝜙i ⎤ ⎡− cos 𝜙i sin 𝜃 i cos 𝜃 i 0⎤


⎢ ⎥ ⎢ ⎥
Gi = ⎢0 sin 𝜓 i cos 𝜓 i cos 𝜙i ⎥ , Gi = ⎢ sin 𝜙i 0 1⎥ (6.21)
⎢ ⎥ ⎢ ⎥
⎣1 0 sin 𝜙i ⎦ ⎣ cos 𝜙 cos 𝜃 sin 𝜃 0⎦
i i i

When the four Euler parameters are used, the two matrices Gi and Gi are given, respec-
tively, by

⎡−𝜃1i 𝜃0i −𝜃3i 𝜃2i ⎤ ⎡−𝜃1i 𝜃0i 𝜃3i −𝜃2i ⎤


⎢ ⎥ ⎢ ⎥
Gi = 2 ⎢−𝜃2i 𝜃3i 𝜃0i −𝜃1i ⎥ , Gi = 2 ⎢−𝜃2i −𝜃3i 𝜃0i 𝜃1i ⎥ (6.22)
⎢ i i ⎥ ⎢ i i ⎥
⎣−𝜃3 −𝜃2 𝜃1 𝜃0 ⎦ ⎣−𝜃3 𝜃2 −𝜃1 𝜃0 ⎦
i i i i

Using the definition of matrices Gi and Gi , the velocity vector and the virtual change in the
position vector of the arbitrary point can be written, respectively, as

ṙ i = Li q̇ i , 𝛿ri = Li 𝛿qi (6.23)


[ ] [ ]
where Li = I −̃ ui Gi = I −Ai ũ i Gi , and I is the 3 × 3 identity matrix. The acceleration
vector can be written as

r̈ i = Li q̈ i + L̇ i q̇ i (6.24)
( )
Other alternate forms of the acceleration vector are r̈ i = R̈ i + 𝛂i × ui + 𝛚i × 𝛚i × ui and
( ) ( ( ))
̈ i + Ai 𝛂i × ui + Ai 𝛚i × 𝛚i × ui , where 𝛂i = Ai 𝛂i and 𝛂i are the angular accel-
r̈ i = R
eration vectors defined, respectively, in the global and body coordinate systems. The form
of the acceleration vector of Eq. 24 is used in the following section to derive Newton–Euler
equations for rigid bodies.
Equations of Motion 235

6.4 INERTIA FORCES


Inertia forces play a significant role in the dynamics and stability of railroad vehicle systems.
Because of inertia forces, limits on the speed of the railroad vehicle in motion scenarios must
be observed and are specified in the operation and safety guidelines. In general, inertia
forces can be developed using two types of forces: one is linear in accelerations, and the
second is quadratic in velocities. The quadratic velocity forces give rise to centrifugal forces
and gyroscopic moments that must be accounted for, particularly during curve negotiations.
The dynamics of rigid bodies are governed by the well-known Newton–Euler equations,
which will be derived in this chapter after developing expressions for the inertia and
applied forces. Two forms of Newton–Euler equations can be developed. In the first form,
the equations are expressed in terms of the angular acceleration vector and the Cartesian
definition of the moments. In the second form, the equations are expressed in terms of
the second derivatives of the orientation coordinates, which are not directly associated
with the actual Cartesian moments. Recall that the angular acceleration vector 𝛂i can be
written in terms of the time derivatives of the orientation parameters 𝛉̈ i as 𝛂i = Gi 𝛉̈ i + Ġ i 𝛉̇ i .
Similarly, one can write 𝛂i = Gi 𝛉̈ i + Ġ i 𝛉̇ i . These kinematic relationships can be used to
define alternate forms of the inertia forces of rigid bodies.

Mass Matrix A rigid body can be assumed to consist of an infinite number of particles.
The mass of an infinitesimal volume of body i can be written as dmi = 𝜌i dV i , where 𝜌i is
the mass density and V i is the body volume. Therefore, ( the ) inertia
( force) of this infinitesi-
mal volume can be written, as shown in Figure 2, as dm r̈ = 𝜌 dV r̈ i . Therefore, the
i i i i

virtual work of the inertia force of body i can be written, upon using Eq. 24, as
T
𝛿Wii = 𝜌i r̈ i 𝛿ri dV i
∫V i
( )T
= 𝜌i Li q̈ i + L̇ i q̇ i 𝛿ri dV i (6.25)
∫V i
Using the definition of 𝛿ri = Li 𝛿qi of Eq. 23, one can show that the virtual work of the inertia
forces can be written as
( )T
𝛿Wii = Mi q̈ i − Qiv 𝛿qi (6.26)

(ρidV i)ri
i i
Z Y

dmi
i
O
i
X

Figure 6.2 Inertia forces.


236 Mathematical Foundation of Railroad Vehicle Systems

where

𝜌i Li L̇ i q̇ i dV i
T T
Mi = 𝜌i Li Li dV i , Qiv = − (6.27)
∫V i ∫V i
In these equations, Mi is the symmetric mass matrix of the body and Qiv is a vector of inertia
[ ]
forces that is quadratic in the velocities. Using the definition Li = I −Ai ũ i Gi in which
[ ] T
ui = xi yi zi is the position of the arbitrary point with respect to the body coordinate
system, the mass matrix can be written as
[ ]
I [ ] i
( i i i )T I −Ai ũ i Gi dV
T i
Mi = 𝜌i Li Li dV = 𝜌i
∫V i ∫V i − AuG ̃
[ ] [ ]
I −A ̃ i Gi
iu
m i
m i
i RR R𝜃
= 𝜌i ( )T iT iT i i dV = (6.28)
∫V i − Ai ũ i Gi G ũ ũ G mi𝜃R mi𝜃𝜃
where

miRR = ∫V i 𝜌i IdV i = mi I, ⎫
( ) ⎪
miR𝜃 = mi𝜃R = −Ai ∫V i 𝜌i ũ i dV i Gi ,
T ⎪
⎬ (6.29)
( ) ⎪
m𝜃𝜃 = G ∫V i 𝜌 ũ ũ dV G = G I𝜃𝜃G ⎪
i
i iT i iT i i i T i i

̃ iT ũ i dV i is the constant and
i = ∫ 𝜌i u
In this equation, I is the 3 × 3 identity matrix, and I𝜃𝜃 Vi
symmetric inertia tensor of the body, which can be written more explicitly as
⎡iixx iixy iixz ⎤
⎢ ⎥
i
I𝜃𝜃 = 𝜌i ũ ũ i dV i = ⎢iixy
iT
iiyy iiyz ⎥
∫V i ⎢i ⎥
⎣ixz iiyz iizz ⎦
(( ) )
(
⎡ yi 2 + zi )2 −xi yi ⎤
−xi zi
⎢ ⎥
⎢ (( ) ( )2 ) ⎥ i
2
= 𝜌i ⎢ −xi yi x i + zi −yi zi ⎥ dV (6.30)
∫V i ⎢ (( ) )
2 ( )2 ⎥
⎢ −xi zi −yi zi xi + yi ⎥
⎣ ⎦
The diagonal elements iixx , iiyy , and iizz are called the mass moments of inertia, while the
off-diagonal elements iixy , iixz , and iiyz are called the products of inertia.

Principal Mass Moments of Inertia It is clear from Eq. 30 that the definitions of the
mass moments and products of inertia depend on the choice of the body coordinate system
in which vector ui is defined. The orientation of the body coordinate system X i Y i Z i can
be selected such that the products of inertia iixy , iixz , and iiyz are all zeros, and the result-
ing mass moments of inertia include the maximum and minimum values. These mass
moments of inertia are called the principal moments of inertia, and the axes of the body
Equations of Motion 237

coordinate system in this case are called the principal axes or principal directions. To deter-
mine the principal moments of inertia and principal axes, one can solve the eigenvalue
[ i ]
problem I𝜃𝜃 − 𝜇 i I di = 𝟎, where 𝜇 i is the eigenvalue that defines the principal moments of
inertia and di is the eigenvector that defines the principal directions. To obtain a nontrivial
[ i ] | i |
solution, the matrix I𝜃𝜃 − 𝜇 i I is assumed to be singular: that is, |I𝜃𝜃 − 𝜇 i I| = 0. The cubic
| |
| i |
polynomial resulting from the condition |I𝜃𝜃 − 𝜇 i I| = 0 is called the characteristic equation.
| |
The roots of the characteristic equation define the three eigenvalues 𝜇ki , k = 1, 2, 3, which
are the three principal moments of inertia. Associated with these three eigenvalues are
three eigenvectors dik , k = 1, 2, 3, which can be determined to within an arbitrary con-
[ i ]
stant using the equation I𝜃𝜃 − 𝜇ki I dik = 𝟎, k = 1, 2, 3. These three eigenvectors define the
i
three principal axes or directions. Because the inertia tensor I𝜃𝜃 is symmetric, the eigen-
values 𝜇k are real, and the eigenvectors dk are orthogonal vectors. Because the eigenvec-
i i

tors are determined to within an arbitrary constant, these eigenvectors can be scaled to be
orthogonal unit ]vectors d ̂i , k = 1, 2, 3, which can be used to form the orthogonal matrix
[ k
i
D = d ̂i d̂i d ̂i . Using this orthogonal matrix and the symmetry of the inertia tensor,
1 2 3
( i ) T i
one can show that I𝜃𝜃 pr
= Di I𝜃𝜃 Di is a diagonal matrix whose diagonal elements are the
T
i di , k = 1, 2, 3.
eigenvalues 𝜇ki = dik I𝜃𝜃 k

Quadratic Velocity Inertia Forces To develop an expression for the quadratic


vector Qiv = −∫V i 𝜌i Li L̇ i q̇ i dV i of [Eq. 27, L̇ i =
T
velocity
[ ( inertia force )] ( one can write )]
i i i i i ̇
𝟎 − Ȧ ũ G + A ũ G , which can be written as L̇ = 𝟎 − A 𝛚
i i i i i i i i ̇
̃ ũ G + A ũ G . It
i

follows that
( )
̃ i ũ i 𝛚i + Ai ũ i Ġ i 𝛉̇ i
L̇ i q̇ i = − Ai 𝛚
( ( )) ( )
= Ai 𝛚i × 𝛚i × ui + Ai Ġ i 𝛉̇ i × ui (6.31)

Therefore, one can write

𝜌i Li L̇ i q̇ i dV i
T
Qiv = −
∫V i
( ( )) ( )
⎡ Ai 𝛚i × 𝛚i × ui + Ai Ġ i 𝛉̇ i × ui ⎤
=− 𝜌i ⎢ [ ( ) ]⎥ dV i (6.32)
∫V i ⎢−GiT ũ iT (𝛚i × (𝛚i × ui )) + Ġ i 𝛉̇ i × ui ⎥
⎣ ⎦
It can be shown that this vector can be written as
[( ) ] (( ( )) ( ) )
Q i ⎡Ai 𝛚i × 𝛚i × hi + Ġ i 𝛉̇ i × hi ⎤
Qiv = ( i )R = − ⎢ )) ⎥
v
( ( (6.33)
Qv 𝜃 ⎢ GiT 𝛚i × (I i 𝛚i ) + I i Ġ i 𝛉̇ i ⎥
⎣ 𝜃𝜃 𝜃𝜃 ⎦
238 Mathematical Foundation of Railroad Vehicle Systems

In this equation, hi = ∫V i 𝜌i ui dV i is the moment of mass of the body. Vector Qiv of Eq. 33 can
also be written using vectors defined in the global coordinate system as
[( ) ] [ ( ) ( ) ]
Qiv R 𝛚i × 𝛚i × hi + Ġ i 𝛉̇ i × hi
Qv = ( i ) = −
i
T ( ( ) ( )) (6.34)
Q v 𝜃 Gi 𝛚i × Ii 𝛚i + Ii Ġ i 𝛉̇ i 𝜃𝜃 𝜃𝜃

where Ii𝜃𝜃 = i AiT ,


Ai I𝜃𝜃 𝛚i = A i 𝛚i , Gi = Ai G i , and hi = Ai hi .
( )
Centrifugal Forces and Gyroscopic Moments In Eq. 33, vector Qiv R associated
with the translation coordinate of the body is the centrifugal force vector, while vector
𝛚i × I𝜃𝜃 i 𝛚i is the gyroscopic moment. Centrifugal forces and gyroscopic moments, which

are, respectively, inertia forces and moments, play an important role in the nonlinear
dynamics and stability of railroad vehicle systems. The source of the gyroscopic moment
can be better understood by examining the expression of the absolute acceleration vector
( )
̈ i + 𝛂i × ui + 𝛚i × 𝛚i × ui . The last term in this acceleration expression is the
r̈ i = R
normal component, which gives rise to the centrifugal force vector 𝛚i × (𝛚i × hi ) in vector
( i)
Qv R of Eq. 34. From Eq. 32, it is clear that the gyroscopic moment can be written
( ( ))
in terms of ∫V i 𝜌i ui × 𝛚i × 𝛚i × ui dV i , which can be interpreted as the moment of
the centrifugal force. Because centrifugal forces can assume large values during curve
negotiations, a limit is imposed on the speed of the vehicle to avoid derailment. This
speed, called the balance speed, is determined using the superelevation, which is designed
to ensure that the lateral component of the centrifugal force will not exceed the lateral
component of the gravity force. Gyroscopic moments also have a significant effect on
vehicle dynamics and stability and must be taken into consideration to ensure stability
during curve negotiations, as discussed in the literature (Shabana et al. 2013).

Newton–Euler Assumption The Newton–Euler equations for a rigid body are derived
using the assumption that the body reference point Oi is selected to be the body center
of mass. This assumption leads to simplifications in the form of the equations of motion,
including eliminating the inertia coupling between the body translation and rotation. If the
reference point is selected to be the body center of mass, the moment of mass about the cen-
ter of mass (reference point) must be equal to zero. In this case, one has(hi = ∫V i 𝜌i ui)dV i = 𝟎.
Using this condition, it is clear from Eq. 29 that mi = mi = −Ai ∫ i 𝜌i ũ i dV i Gi = 𝟎,
T
R𝜃 𝜃R V
and the mass matrix of the body of Eq. 28 reduces to
[ ] [ ]
i
miRR 𝟎 mi I 𝟎
M = = T i (6.35)
𝟎 mi𝜃𝜃 𝟎 Gi I𝜃𝜃 Gi
In this form of the mass matrix, there is no inertia coupling between the body translation
and rotation. Furthermore, using the condition hi = ∫V i 𝜌i ui dV i = 𝟎, the centrifugal force
( ) ( )
vector Qiv R is identically zero: that is, Qiv R = 𝟎. In this case, vector Qiv reduces to
[( ) ]
Qiv ⎡ 𝟎 ⎤
Qv = ( i )R = − ⎢ iT ( i ( i i )
i ( ))⎥
Qv 𝜃 ⎢G 𝛚 × I𝜃𝜃 𝛚 + I𝜃𝜃 i Ġ i 𝛉̇ i ⎥
⎣ ⎦
[ ]
𝟎
=− T ( i ( i i) ( )) (6.36)
G 𝛚 × I 𝛚 + Ii Ġ i 𝛉̇ i
i
𝜃𝜃 𝜃𝜃
Equations of Motion 239

Therefore, the virtual work of the inertia forces of Eq. 26 can be written as
( )T
𝛿Wii = Mi q̈ i − Qiv 𝛿qi
( ( ( i i) ( )))T
̈ iT 𝛿Ri + GiT I𝜃𝜃 Gi 𝛉̈ i + Gi 𝛚i × I𝜃𝜃 Ġ i 𝛉̇ i
T
= mi R i
𝛚 + I𝜃𝜃
i
𝛿𝛉i
( )T ( )T
= Qii R 𝛿Ri + Qii 𝜃 𝛿𝛉i (6.37)

or, alternatively, in terms of vectors defined in the global system as


( )T
𝛿Wii = Mi q̈ i − Qiv 𝛿qi
( ( ( ) ( )))T i
̈ iT 𝛿Ri + GiT Ii𝜃𝜃 Gi 𝛉̈ i + GiT 𝛚i × Ii𝜃𝜃 𝛚i + Ii𝜃𝜃 Ġ i 𝛉̇ i
= mi R 𝛿𝛉
( )T ( )T
= Qii R 𝛿Ri + Qii 𝜃 𝛿𝛉i (6.38)
( ) ( )
In both Eqs. 37 and 38, Qii R = mi R ̈ i ; and in Eq. 37, Qi = GiT I𝜃𝜃 ̈i +
i Gi 𝛉
( ( ) ( )) i 𝜃
T
Gi 𝛚i × I𝜃𝜃 i 𝛚i + I i ̇ i̇i
𝜃𝜃 G 𝛉 are, respectively, the generalized inertia forces asso-
ciated with the translational and orientation coordinates of the body. The generalized
inertia force associated with the orientation coordinates can also be written using Eq. 38 as
( i) T ( ( ))
Qi 𝜃 = Gi Ii𝜃𝜃 Gi 𝛉̈ i + Gi 𝛚i × Ii𝜃𝜃 𝛚i + Ii𝜃𝜃 Ġ i 𝛉̇ i . These expressions for the inertia forces
T

will be used to derive the Newton–Euler equations for rigid bodies.

6.5 APPLIED FORCES

The principle of virtual work of Eq. 8 states that the virtual work of inertia forces obtained
in the preceding section is equal to the virtual work of applied forces. This section discusses
the formulation of the virtual work of applied forces. It is shown that Cartesian moments are
not associated with angles, and generalized forces associated with orientation parameters
can be obtained from Cartesian moments using the two matrices Gi and Gi that are used to
define the angular velocity vectors in the global and body coordinate systems, respectively.
Figure 3 shows a force vector Fi acting at point Pi on a rigid body. The global position of
point Pi is defined by vector riP = Ri + Ai uPi, where uPi is the local position of the point of
application of the force. The virtual work of this force vector is defined as the dot product
of the force and the virtual change in the position vector of the point of application of the
force: that is,
T T
𝛿Wei = Fi 𝛿riP = Fi LiP 𝛿qi (6.39)
[ ]
where LiP = I −Ai ũ PiGi . Equation 39 can then be written as
( )T ( )T
𝛿Wei = Qie R 𝛿Ri + Qie 𝜃 𝛿𝛉i (6.40)
( ) ( ) T ( )
where Qie R = Fi and Qie 𝜃 = −Gi ũ Pi Ai Fi are, respectively, the generalized applied
T T

forces associated with the translational and orientation coordinates of the body. Equation 40
implies that force vector Fi acting at an arbitrary point Pi on the body is equipollent to
240 Mathematical Foundation of Railroad Vehicle Systems

Zi
Yi

Oi
ui
Pi Fi
Xi
Z
Ri
Y
ri

Figure 6.3 Applied force.

another system defined at point Oi , which consists of the same force vector Fi associated
( ) T ( )
with the body translation, and a generalized force vector Qie 𝜃 = −Gi ũ Pi Ai Fi associ-
T T

ated with the orientation coordinates. To understand the relationship between the gen-
( )
eralized force vector Qie 𝜃 associated with the orientation coordinates and the Cartesian
( i)
moment, vector Qe 𝜃 is written as
( i) ( ) T ( ) T ( )
Qe 𝜃 = Gi ũ PiAi Fi = Gi uPi × Fi = Gi uiP × Fi
T T
(6.41)
T
where Fi = Ai Fi is the force vector defined in the body coordinate system, and
uiP = Ai uPi. The preceding equation shows that the Cartesian moment of the force
( )
uiP × Fi = Ai uPi × Fi is not the moment associated with the orientation parameters 𝛉i .
The generalized forces associated with the orientation parameters 𝛉i can be obtained from
the applied moments using the transpose of the matrices Gi and Gi : that is, if Mi𝛼 = Ai M𝛼i
is a Cartesian moment that applies to rigid body i, then the generalized forces associated
with the orientation parameters are defined as
( i) T T
Qe 𝜃 = Gi Mi𝛼 = Gi M𝛼i (6.42)
where M𝛼i is the Cartesian moment defined in the body coordinate system. Equation 42 will
be used in converting the generalized Newton–Euler equations obtained using the virtual
work principle to the Newton–Euler equations expressed in terms of angular acceleration
vectors 𝛂i and 𝛂i . In general, for any system of forces and moments acting on a rigid body,
Equations of Motion 241

( ) ( )
an equation similar to Eq. 40 can be developed, where Qie R and Qie 𝜃 are vectors of the
resultant generalized forces associated with the generalized translational and orientation
coordinates, respectively.

6.6 NEWTON–EULER EQUATIONS

Two forms of Newton–Euler equations are presented in this section. The first is the general-
ized Newton–Euler equations obtained directly from using the principle of virtual work. This
form of the Newton–Euler equations is written in terms of generalized orientation parame-
ters. In the second form, the Newton–Euler equations are expressed in terms of the angular
acceleration vector. These two alternate forms of the Newton–Euler equations show that
the Cartesian moment Mi𝛼 = Ai M𝛼i is associated with the angular acceleration vector and
not with the second time derivative of the orientation coordinates.

Generalized Newton–Euler Equations Using the virtual work principle of Eq. 8,


( )T ( )T
𝛿W i = 𝛿W e , and using the virtual work of the inertia forces 𝛿Wii = Qii R 𝛿Ri + Qii 𝜃 𝛿𝛉i
( )T ( ) T
of Eq. 37 and the virtual work of the applied forces 𝛿Wei = Qie R 𝛿Ri + Qie 𝜃 𝛿𝛉i of Eq. 40,
one obtains
(( i ) ( ) )T (( ) ( ) )T
Qi R − Qie R 𝛿Ri + Qii 𝜃 − Qie 𝜃 𝛿𝛉i = 0 (6.43)
[ T T ] T
In the case of unconstrained motion, the elements of vector qi = Ri 𝛉i are indepen-
dent, and therefore, the coefficients of their virtual change in the preceding equation must
( ) ( ) ( ) ( )
be equal to zero: that is, Qii R = Qie R and Qii 𝜃 = Qie 𝜃 . Using the definition of the iner-
tia forces given in Eq. 37, one obtains the two equations
( ) ⎫
̈ i = Qie
mi R R
( ( ))⎪
( iT i i ) i ( i ) ( i i) ⎬ (6.44)
G I𝜃𝜃 G 𝛉̈ = Qe 𝜃 − Gi 𝛚i × I𝜃𝜃 Ġ i 𝛉̇ i ⎪
T
𝛚 + I𝜃𝜃
i

These are the generalized Newton–Euler equations, which can be written in a matrix
form as
[ ] [ ] [( ) ]
mi I 𝟎 R̈i Qie R ⎡ 𝟎 ⎤
( iT i i ) = ( ) + ⎢ ( ( ) ( ))⎥ (6.45)
𝟎 G I𝜃𝜃 G 𝛉̈ i Qie 𝜃 ⎢−G 𝛚 × I𝜃𝜃 𝛚 + I𝜃𝜃 G 𝛉̇
iT i i i i ̇ i i ⎥
⎣ ⎦
This equation can be written in terms of vectors defined in the global coordinate system as
[ ] [ ] [( ) ] [ ]
mi I 𝟎 R̈i Qie R 𝟎
( iT i i ) = ( i) + T ( ( ) ( )) (6.46)
𝟎 G I G𝜃𝜃 𝛉̈ i Qe 𝜃
−Gi 𝛚i × Ii 𝛚i + Ii Ġ i 𝛉̇ i
𝜃𝜃 𝜃𝜃

Equations 45 and 46 can be used to obtain the Newton–Euler equations written in terms of
the angular acceleration vectors 𝛂i and 𝛂i , respectively.
242 Mathematical Foundation of Railroad Vehicle Systems

Newton–Euler Equations Substituting the equation 𝛂i = Gi 𝛉̈ i + Ġ i 𝛉̇ i into Eq. 45 and


using Eq. 42, one obtains
[ ] [ ] [( ) ] [ ]
mi I 𝟎 ̈i
R Qie R 𝟎
= + ( i i) (6.47)
𝟎 I𝜃𝜃i 𝛂i M𝛼i −𝛚i × I𝜃𝜃 𝛚
In this form of the Newton–Euler equations, the coefficient matrix is constant. Alterna-
tively, one can use the relationship 𝛂i = Gi 𝛉̈ i + Ġ i 𝛉̇ i to obtain the Newton–Euler equations
written in terms of vectors defined in the global coordinate system as
[ ] [ ] [( ) ] [ ]
mi I 𝟎 R̈i Qie R 𝟎
= + ( ) (6.48)
𝟎 Ii𝜃𝜃 𝛂i Mi𝛼 −𝛚i × Ii𝜃𝜃 𝛚i
In this form of the Newton–Euler equations, the coefficient matrix is not constant since
i AiT . One can show that the Newton–Euler equations of Eq. 47 or 48 reduce to
Ii𝜃𝜃 = Ai I𝜃𝜃
the three Newton–Euler equations used in the planar analysis, as demonstrated by the fol-
lowing example.

Example 6.3 In the case of planar analysis, the global Z axis and body Z i axis coincide.
In this special case, one can write the six accelerations R ̈ i and 𝛂i in terms of only three
acceleration components R̈ x , R̈ y , and the angular acceleration 𝜃̈zi about the Z axis as
i i

⎡R̈ ix ⎤ ⎡1 0 0⎤
⎢̈i⎥ ⎢ ⎥
R 0 1 0⎥ i
[ ] ⎢ y⎥ ⎢ ⎡R̈ x ⎤
̈
R i ⎢ ̈
Rz i ⎥ ⎢0 0 0⎥ ⎢ i ⎥
= ⎢ i⎥ = ⎢ ⎥ ⎢R̈ ⎥ = B q̈
𝛂 i ⎢ 𝛼 ⎥ ⎢0 0 0⎥ ⎢ i ⎥
y di i
⎢ x⎥ ⎢ ⎥ ⎣ 𝜃̈z ⎦
⎢ 𝛼iy⎥ ⎢0 0 0⎥
⎢ i⎥ ⎢ ⎥
⎣ 𝛼 z⎦ ⎣0 0 1⎦
where
⎡1 0 0⎤
⎢ ⎥
⎢0 1 0⎥
⎡R̈ ix ⎤
⎢0 0 0⎥ ⎢ ⎥
Bdi = ⎢ ⎥, q̈ i = ⎢R̈ iy ⎥
⎢0 0 0⎥ ⎢ ̈i ⎥
⎢ ⎥ ⎣ 𝜃z ⎦
⎢0 0 0⎥
⎢ ⎥
⎣0 0 1⎦
One can show that, when Euler angles are used, the acceleration transformation con-
sidered in this example is the result of imposing the three motion constraints Riz = 0,
𝜙i = 0, and 𝜃 i = 0. In the case of planar motion, the angular velocity vector is given
[ ]T
by 𝛚i = 0 0 𝜃̇ zi . Substituting for R̈ i and 𝛂i into Eq. 47, and premultiplying by the
Equations of Motion 243

transpose of matrix Bdi , one obtains the planar Newton–Euler equations defined as
⎡mi 0 0 ⎤ ⎡R̈ ix ⎤ ⎡Qiex ⎤
⎢ ⎥⎢ i⎥ ⎢ i ⎥
⎢ 0 m 0 ⎥ ⎢R̈ y ⎥ = ⎢Qey ⎥
i

⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ 0 0 iizz ⎦ ⎣ 𝜃̈zi ⎦ ⎣ Mzi ⎦
In this equation, Qiex and Qiey are the components of the resultant force acting on the body,
and Mzi is the resultant moment about the Z axis. It is clear from the previous equation
that there is no gyroscopic moment in the case of planar motion since the rotation is
about a fixed axis.

Euler Parameters The generalized form of the Newton–Euler equations of motion is


used, for the most part, in this book. This form allows for using the same form of the con-
straint Jacobian matrix Cq for the position, velocity, and acceleration analyses. Equation 45
can be written in the case of unconstrained motion for rigid body i as
Mi q̈ i = Qie + Qiv , i = 1, 2, … , nb (6.49)
where nb is the total number of bodies in the system and
[ ] [ ] [( ) ] ⎫
mi I 𝟎 ̈i
R Qie ⎪
Mi = ( iT i i ) , q̈ i = , Qie = ( i )R ,⎪
𝟎 G I𝜃𝜃 G 𝛉̈ i Qe 𝜃 ⎪
⎡ 𝟎 ⎤ ⎬ (6.50)

Qiv =⎢ ( ( ) ( )) ⎥

⎢ −GiT 𝛚i × I𝜃𝜃 i 𝛚i + I𝜃𝜃

i Ġ i 𝛉̇ i ⎥
⎦ ⎪

It is clear from the preceding two equations that the system inertia force vector Qi
of Eq. 17 or 18 can be written in terms of the inertia forces of the bodies as Qi =
[ ]T
T T n T
Q1i Q2i … Qi b , where
[( ) ] ̈i
Qii R ⎡ mi R ⎤
Qii ⎢
= ( i ) = ( iT i i ) i ( ( ) ( ))⎥ ,
⎢ G I𝜃𝜃 G 𝛉̈ + Gi 𝛚i × I𝜃𝜃
T i 𝛚i + I i ̇ i̇i ⎥
Qi 𝜃
⎣ 𝜃𝜃 G 𝛉 ⎦
i = 1, 2, … , nb (6.51)

If Euler parameters are used as the orientation coordinates, Ġ i 𝛉̇ i = 𝟎, and the preceding
equation reduces to
[( i ) ] [ ]
Qi R ̈i
mi R
i
Qi = ( ) = ( T ) T ( ( i i )) , i = 1, 2, … , nb (6.52)
Qii 𝜃 Gi I𝜃𝜃
i
Gi 𝛉̈ i + Gi 𝛚i × I𝜃𝜃 𝛚
244 Mathematical Foundation of Railroad Vehicle Systems

The simplification resulting from using Euler parameters comes at the expense of adding
an algebraic constraint equation for each body in the system, as discussed in Chapter 3.
Nonetheless, using Euler parameters is recommended for developing general MBS railroad
vehicle algorithms because of the singularity problem associated with any three-parameter
representation. Equation 49 can be used with Eq. 17 as the basis for developing two different
procedures for formulating the equations of motion of complex railroad vehicle systems
subject to kinematic constraints, as discussed in the following section.

6.7 AUGMENTED FORMULATION AND EMBEDDING


TECHNIQUE

It was shown previously (Eq. 17) that the equations of motion of mechanical systems such as
railroad vehicle systems can be written as Qi − Qe + CTq 𝛌 = 𝟎, subject to the kinematic con-
[ ]T
straint equations C (q, t) = C1 C2 … Cnc = 𝟎. In this equation of motion, −CTq 𝛌 defines
the generalized constraint forces. The vector of the inertia forces can always be written as
Qi = Mq̈ − Qv , where M is the system mass matrix, q is the vector of the system coordi-
nates, and Qv is the vector of inertia forces that is quadratic in the velocities. Based on the
development presented in the preceding section, one can write the system mass matrix M
and the vectors q and Qv as
⎡M1 𝟎 ··· 𝟎 ⎤ ⎡ q1 ⎤ ⎡ Q1v ⎤
⎢ ⎥ ⎢ 2⎥ ⎢ 2⎥
⎢ 𝟎 M2 ··· 𝟎 ⎥ ⎢q ⎥ ⎢ Qv ⎥
M=⎢ ⎥, q = ⎢ ⎥, Qv = ⎢ ⎥ (6.53)
⎢ ⋮ ⋮ ⋱ ⋮ ⎥ ⎢ ⋮ ⎥ ⎢ ⋮ ⎥
⎢ 𝟎 𝟎 · · · Mnb ⎥⎦ ⎢qnb ⎥ ⎢Qnb ⎥
⎣ ⎣ ⎦ ⎣ v ⎦
Therefore, the system equations of motion Qi − Qe + CTq 𝛌 = 𝟎 can be written as
Mq̈ + CTq 𝛌 = Qe + Qv . For a given set of forces and moments that define vector Qe ,
the n scalar equations that form the matrix equation of motion Mq̈ + CTq 𝛌 = Qe + Qv
have n unknown accelerations that form vector q̈ plus nc unknown Lagrange multipliers
that form vector 𝛌. Therefore, nc additional equations are needed in order to determine
the n + nc unknowns. These additional equations are the kinematic constraint functions
[ ]T
defined by vector C (q, t) = C1 C2 … Cnc = 𝟎. Differentiating this vector twice with
respect to time, one obtains two types of vectors: one is linear in acceleration, and the
other absorbs all other terms that are not linear in acceleration, including terms that are
quadratic in velocities. Therefore, one can always define the constraint functions at the
acceleration level as Cq q̈ = Qd , where Cq = 𝜕C/𝜕q is the constraint Jacobian matrix and Qd
is the vector that absorbs terms that are not linear in accelerations.
The two equations Mq̈ + CTq 𝛌 = Qe + Qv and Cq q̈ = Qd represent the foundation for
developing two computational procedures that are discussed in this chapter. These two
equations are reproduced here for convenience:
}
Mq̈ + CTq 𝛌 = Qe + Qv
(6.54)
Cq q̈ = Qd
Equations of Motion 245

In the first approach, referred to as the augmented formulation, redundant coordinates are
used to obtain a system of equations written explicitly in terms of constraint forces. In the
second approach, referred to as the embedding technique, constraint equations are used to
systematically eliminate dependent variables, leading to a system of equations from which
the constraint forces are eliminated.

Example 6.4 In Example 1, the three constraint equations of a spherical joint between
two arbitrary bodies i and j at points Pi and Pj , respectively, were written as
( ) j j
C qi , qj = riP − rP = Ri + Ai uiP − Rj − Aj uP = 𝟎
where rkP , Rk , Ak and ukP , k = i, j, are, respectively, the global position vector of the joint
definition point Pk , the global position vector of the reference point of body k, the trans-
formation matrix that defines the body orientation, and the local position vector of point
Pk with respect to the body coordinate system. Vector qk is the vector of absolute Carte-
[ T T ]T
sian coordinates qk = Rk 𝛉k . Differentiating the constraint equations of the spheri-
cal joint with respect to time, one obtains the constraint equations at the velocity level as
( )
Ċ qi , qj = ṙ iP − ṙ P = Ṙ i + 𝛚i × uiP − Ṙ j − 𝛚j × uP = 𝟎
j j

where uiP = Ai uiP , and 𝛚i = Gi 𝛉̇ i is the angular velocity vector. Differentiating the
preceding equation with respect to time, one obtains the spherical-joint constraint
equations at the acceleration level as
( )
C̈ qi , qj = r̈ iP − r̈ P
j

( ) ( )
=R ̈ j − 𝛂j × ujP − 𝛚j × 𝛚j × ujP = 𝟎
̈ i + 𝛂i × uiP + 𝛚i × 𝛚i × uiP − R

In this equation, 𝛂k = 𝛚̇ k = Gk 𝛉̈ k + Ġ k 𝛉̇ k , k = i, j. Therefore, the preceding equation


can be written as
( ) ( ) ( )
̈i−u
C̈ qi , qj = R ̃ iP 𝛂i + 𝛚i × 𝛚i × uiP − R ̈j+u ̃ jP 𝛂j − 𝛚j × 𝛚j × ujP
( ) ( ) ( )
̈i−u
=R ̃ iP Gi 𝛉̈ i + Ġ i 𝛉̇ i + 𝛚i × 𝛚i × uiP − R ̈j+u ̃ jP Gj 𝛉̈ j + Ġ j 𝛉̇ j − 𝛚j
( )
j
× 𝛚j × u P = 𝟎

This equation can be rearranged and rewritten as


( )
̈i−u
C̈ qi , qj = R ̃ iP Gi 𝛉̈ i − R ̃ jP Gj 𝛉̈ i
̈j+u
( ) ( )
+ 𝛚i × 𝛚i × uiP − u ̃ iP Ġ i 𝛉̇ i − 𝛚j × 𝛚j × ujP + u
̃ PjĠ j 𝛉̇ j = 𝟎

It is clear that this equation has two types of terms: the first is linear in accelerations,
and the second absorbs all other terms that are not linear in accelerations. Therefore,
the preceding equation can be written as
( )
̈i−u
C̈ qi , qj = R ̃ iP Gi 𝛉̈ i − R ̃ jP Gj 𝛉̈ i − Qd = 𝟎
̈j+u

(Continued)
246 Mathematical Foundation of Railroad Vehicle Systems

where vector Qd that absorbs terms that are not linear in the acceleration is recognized as
( ( ) ( ) )
Q = − 𝛚i × 𝛚i × u i − u
d ̃ i Ġ i 𝛉̇ i − 𝛚j × 𝛚j × uj + u
P P P ̃ j Ġ j 𝛉̇ j
P

For mechanical joint constraints, which are not explicit functions of time, vector Qd is
quadratic in velocities. It is also important to recognize that the terms that are linear in
accelerations are the same as Cq q: ̈ that is,

Cq q̈ = R̈i−ũ iP Gi 𝛉̈ i − R
̈j+u ̃ jP Gj 𝛉̈ i
[ T T ]T [ T T ]T
where q = qi qj , qk = Rk 𝛉k , k = i, j, and the constraint Jacobian matrix Cq
was obtained in Example 1 as
[ ] [ ]
𝜕C 𝜕C 𝜕C 𝜕C 𝜕C 𝜕C
Cq = =
𝜕qi 𝜕qj 𝜕Ri 𝜕𝛉i 𝜕Rj 𝜕𝛉j
[ ]
= I −Ai ũ P G −I Aj ũ P G
i i j j

[ ]
= I −̃ uiP Gi −I u ̃ jP Gj

If Euler parameters, instead


( of Euler angles, are used, has Ġ i 𝛉̇ i = Ġ j 𝛉̇ j = 𝟎, and vec-
( one ))
( ) j
tor Qd reduces to Qd = − 𝛚i × 𝛚i × uiP − 𝛚j × 𝛚j × uP .

Augmented Formulation In the augmented formulation, the two equations in Eq. 54


are combined to form one matrix equation as
[ ][ ] [ ]
M CTq q̈ Qe + Q v
= (6.55)
Cq 𝟎 𝛌 Qd
For given initial coordinates and velocities, the coefficient matrix and the right-hand side of
this equation can be constructed. Therefore, Eq. 55 can be solved for the acceleration vector
q̈ and the vector of Lagrange multipliers 𝛌. The vector of Lagrange multipliers 𝛌 can be used
to determine the generalized constraint forces −CTq 𝛌, while the independent components of
the acceleration vector q̈ i can be integrated to determine the independent coordinates and
velocities. The dependent coordinates qd can be determined using the constraint equations
[ ]T
C (q, t) = C1 C2 … Cnc = 𝟎 with the assumption that the independent coordinates qi
are known from the results of the numerical integration. Therefore, the system of the nc
[ ]T
nonlinear constraint equations C (q, t) = C1 C2 … Cnc = 𝟎 becomes a function of only
nc unknown coordinates. This nonlinear system of equations can be solved using an itera-
tive Newton–Raphson procedure. ( In)this iterative procedure, one constructs at each itera-
( )
tion k the system of equations Cqd Δqd k = −(C)k , where Cqd = 𝜕C∕𝜕qd is the nc × nc
k
constraint Jacobian matrix associated with the dependent coordinates, and Δqd is the vec-
tor of Newton differences (Atkinson 1978). Because the constraint equations are assumed to
be linearly independent, the constraint Jacobian matrix Cqd = 𝜕C∕𝜕qd is assumed to have
( ) ( )
full rank, and therefore, a solution of the equation Cqd Δqd k = −(C)k for the Newton
k
differences Δqd can be obtained. At the end of each iteration, the coordinates are updated
Equations of Motion 247

according to the equation (qd )k + 1 = (qd )k + (Δqd )k . Convergence is achieved if the norm
of the Newton differences or the norm of the constraint equations becomes smaller than a
specified tolerance 𝜀: that is, |Δqd | ≤ 𝜀 or |C| ≤ 𝜀.
Having determined the coordinates using the iterative procedure and constraint
[ ]T
equations at the position level C (q, t) = C1 C2 … Cnc = 𝟎, the constraint equations at
the velocity level can be defined as Cq q̇ + Ct = 𝟎, where Ct = 𝜕C/𝜕t is the partial derivative of
the constraint equations with respect to time; this vector is the zero vector if the constraint
equations are not explicit functions of time. If the vector of independent velocities q̇ i is
known from the numerical integration, the constraint equations at the velocity level can
be written as Cqd q̇ d + Cqi q̇ i + Ct = 𝟎. This equation can be arranged and written as
( )
Cqd q̇ d = − Cqi q̇ i + Ct (6.56)
This is a linear system of equations in the dependent velocities q̇ d . Because the Jacobian
matrix Cqd associated with the dependent velocities is non-singular, and because of the
linearity of these equations, the system of Eq. 56 can be solved for the dependent velocities
q̇ d without the need to use an iterative procedure required the position analysis step. Once
all the coordinates and velocities (independent and dependent) are determined, Eq. 55 can
be constructed and solved for all the accelerations, which are used to advance the numerical
integration.

Example 6.5 To explain some of the concepts used in the augmented formulation, a
wheel/rail example is considered (Shabana et al. 2008). Figure 4 shows a wheel with
radius r, mass mw , and mass moment of inertia J w about its center of mass. The wheel
is assumed to roll without sliding on a circular rail, which is assumed to be fixed and to
have a constant radius of curvature R. The external forces acting on the wheel are defined
[ ]T
by the resultant force vector Fw = Fxw Fyw and the moment M w . Because of the rolling
assumption and because the rail is assumed to be fixed, there is no friction force between
the wheel and rail, and the system has one degree of freedom. Using Figure 4, the tangent
and normal vectors at the wheel/rail contact point can be written as
[ ] [ ]
− cos 𝜙 − sin 𝜙
r
t = , r
n =
sin 𝜙 − cos 𝜙
The absolute velocity of the contact point on the wheel is defined as ṙ wc = Ṙ w + 𝛚w × uwc ,
[ ]T [ ]T
where 𝛚w = 0 0 𝜃̇ w , uwc = −r sin 𝜙 −r cos 𝜙 0 , and 𝜃̇ w is the angular velocity of
the wheel. Because of the rolling and no-slipping conditions, the absolute velocity vector
at the contact point is equal to zero, leading to the equation
[ ]
Ṙ wx + r 𝜃̇ w cos 𝜙
ṙ c = Ṙ + 𝛚 × uc =
w w w w
=𝟎
Ṙ wy − r 𝜃̇ w sin 𝜙
These two scalar constraint equations, which ensure that the wheel rolls without slip-
ping, eliminate two dependent velocities; consequently, the system has only one inde-
pendent velocity. The rolling conditions at the acceleration level can be written as r̈ wc = 𝟎.

(Continued)
248 Mathematical Foundation of Railroad Vehicle Systems

O
X

ϕ Fyw
Mw
R
Fxw

r θw Nx
mwg
Ny
c

Figure 6.4 Wheel/rail example.

Using the geometry of the system shown in Figure 4, one has Rwx = − (R − r) sin 𝜙, Rwy =
− (R − r) cos 𝜙, which, upon substitution in the preceding equation that defines ṙ wc , gives
the algebraic relationship between 𝜃̇ w and 𝜙̇ as r 𝜃̇ w = (R − r) 𝜙.̇ That is, 𝜙̇ can be deter-
mined if 𝜃̇ w is known.
As previously mentioned, the Lagrangian approach does not require using free-body
diagrams since the connectivity (constraint) conditions are used to define the constraint
forces. To understand the relationship between Newtonian and Lagrangian mechanics,
the free-body diagram shown in Figure 4, which is not required in Lagrangian dynamics,
is first used. Using this free-body diagram and the Newtonian approach, one can write
the three Newton–Euler equations for the planar wheel as Mw q̈ w = Qwe + Qwc , where
[ ]T
qw = Rwx Rwy 𝜃 w ,
⎡ mw 0 0⎤ ⎡ Fxw ⎤ ⎡ Nx ⎤
⎢ ⎥ ⎢ w ⎥ ⎢ ⎥
w
M =⎢ 0 m w 0 ⎥ , Qe = ⎢Fy − m g⎥ , Qc = ⎢
w w w
Ny ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣Nx r cos 𝜙 − Ny r sin 𝜙⎦
w w
⎣ 0 0 J ⎦ ⎣ M ⎦
[ ]T
N = Nx Ny is the vector of reaction forces at the wheel/rail contact point, and g is
the gravity constant. The matrix equation of motion Mw q̈ w = Qwe + Qwc has three scalar
equations and the five unknowns R̈ wx , R̈ wy , 𝜃̈ w , Nx , and N y . Therefore, the two constraint
equations that define the rolling conditions are needed in order to have a number of
equations equal to the number of unknowns. These two sets of equations can be writ-
ten as
}
Mw q̈ w = Qwe + Qwc ,
r̈ wc = 𝟎
Therefore, when Newtonian mechanics is used, the algebraic constraint equations are
also needed to solve for the unknown accelerations and constraint forces. The rolling
Equations of Motion 249

constraint equations at the acceleration level can be written as


[ ]
R̈ wx + r 𝜃̈ w cos 𝜙 − r 𝜃̇ 𝜙̇ sin 𝜙
C̈ = r̈ wc = Cq q̈ − Qd = =𝟎
R̈ wy − r 𝜃̈ w sin 𝜙 − r 𝜃̇ 𝜙̇ cos 𝜙
Combining this equation with the equations of motion, one obtains the augmented
Lagrangian form of the equations of motion, which does not in general require using
the free-body diagram as (Shabana et al. 2008)

⎡m
w 0 0 −1 0 ⎤ ⎡R̈ wx ⎤ ⎡ Fxw ⎤
⎢ 0 mw 0 0 −1 ⎥ ⎢R̈ w ⎥ ⎢F w − mw g⎥
⎢ ⎥⎢ y ⎥ ⎢ y ⎥
⎢ 0 0 Jw −r cos 𝜙 r sin 𝜙⎥ ⎢ 𝜃̈ w ⎥ = ⎢ M w ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ 1 0 r cos 𝜙 0 0 ⎥ ⎢Nx ⎥ ⎢ r 𝜃̇ 𝜙̇ sin 𝜙 ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ 0 1 −r sin 𝜙 0 0 ⎦ ⎣ Ny ⎦ ⎣ r 𝜃̇ 𝜙̇ cos 𝜙 ⎦
This equation can be written in the form of Eq. 55 as
[ ][ ] [ ]
M CTq q̈ Qe + Qv
=
Cq 𝟎 𝛌 Qd
where, in this planar example, Qv = 0, M = Mw , q = qw , and Qe = Qwe (previously defined
in this example); and
[ ] [ ] [ ]
1 0 r cos 𝜙 r 𝜃̇ 𝜙̇ sin 𝜙 Nx
Cq = , Qd = , 𝛌=−
0 1 −r sin 𝜙 r 𝜃̇ 𝜙̇ cos 𝜙 Ny
It is clear from this simple example that the vector of Lagrange multipliers 𝛌 is the neg-
ative of the reaction forces.

Identification of the System Degrees of Freedom Because of the complex-


ity of railroad vehicle systems, independent and dependent coordinates are identified
numerically using the nc × n constraint Jacobian matrix Cq . This can be achieved using
a Gaussian elimination process with full pivoting (Atkinson 1978; Wehage 1980). Using
the Gaussian elimination procedure, matrix Cq can be converted to a matrix in the form
[ ]
Inc ×nc Inc ×nd , where Inc ×nc is the nc × nc identity matrix, Inc ×nd is an nc × nd matrix that
results from the Gaussian eliminate steps, and nd is the number of independent coordi-
nates. The coordinates associated with the columns of the square non-singular identity
matrix Inc ×nc are selected as the dependent coordinates qd , while the coordinates associated
with the non-square matrix Inc ×nd are assumed to be the independent coordinates qi . The
choice of the dependent and independent coordinates using this numerical procedure
ensures that the dependent variables can always be determined using the kinematic
constraint equations if the independent variables are known. In some mechanical system
applications, it is sufficient to identify the set of independent coordinates (degrees of
freedom) only once at the beginning of the simulation and use this set for the entire
250 Mathematical Foundation of Railroad Vehicle Systems

simulation time. In some other systems, particularly in the case of closed chains, the
set of independent coordinates must be continuously changed in order to avoid singular
configurations. While the number of degrees of freedom remains the same for a given
system topology, the set of degrees of freedom is not unique. This is also clear from the
fact that the constraint Jacobian matrix Cq (q, t) is a nonlinear function of the system
coordinates; and as the system configuration changes with time, the Gaussian procedure
can lead to different sets of dependent and independent coordinates.

Efficient Implementation of the Augmented Formulation In the computer imple-


mentation of the augmented formulation, partitioning the constraint Jacobian matrix as
[ ]
Cq = Cqd Cqi is not necessary. Because most joints and constraints involve no more than
two bodies in the system, the constraint Jacobian matrix Cq is a sparse matrix that has a
large number of zeros. This allows for exploiting sparse-matrix techniques to obtain effi-
cient solutions for the position coordinates, velocities, and accelerations (Duff et al. 1986).
Instead of partitioning the constraint Jacobian matrix, one can form a larger sparse matrix
whose entries ensure that the Newton differences and velocities associated with the inde-
pendent
( ) ( coordinates qi remain unchanged. For example, instead of using the equation
)
Cqd Δqd k = −(C)k in the Newton–Raphson iterations to solve for the coordinates, one
k
can use the following equation at iteration k:
[ ] [ ]
Cq −(C)k
(Δq)k = (6.57)
Id k 𝟎
In this equation, Id is an nd × n Boolean matrix with only zero and one entries, with the
ones in locations that ensure that the Newton differences associated with the independent
coordinates are equal to zero: that is, (Δqi )k = 0. Because the independent coordinates are
assumed to be known from the numerical integration, they are not allowed to vary dur-
ing the Newton–Raphson iterations. The coefficient matrix in Eq. 57 is square and sparse
with a large number of zeros. The structure of this matrix does not change unless the set
of independent coordinates is changed. This fact can be utilized to avoid repeated symbolic
factorization and scaling at every time step.
Similarly, instead of using Eq. 56 to solve for the velocities, one can combine the two
equations Cq q̇ = −Ct and q̇ i = q̇ i to form the following sparse-matrix equation:
[ ] [ ]
Cq −Ct
q̇ = (6.58)
Id q̇ i
In both the position and velocity analysis steps, the locations of the non-zero entries in
the constraint Jacobian matrix Cq do not change. Furthermore, if the set of independent
coordinates remains the same during dynamic simulation, the locations of the non-zero
entries of matrix Id do not change during dynamic simulations, as previously mentioned.
[ ]T
This allows for performing symbolic factorization of matrix CTq ITd for both the position
and velocity analysis steps. This symbolic factorization can also be performed only at points
in time when the set of independent coordinates is changed, to obtain an efficient solution
for both Eqs. 57 and 58 and avoid the need to partition the constraint Jacobian matrix as
[ ]
Cq = Cqd Cqi .
Equations of Motion 251

Similarly, the coefficient matrix in Eq. 55 is a sparse matrix in which the locations of
the non-zero entries do not change during dynamic simulations. Therefore, sparse-matrix
techniques can also be used to efficiently solve for the accelerations and vector of Lagrange
multipliers. It is important to emphasize again that only the independent accelerations
q̈ i must be integrated forward in time, and vectors qi and q̇ i should not be altered
outside the integrator when sophisticated integration methods for solving first-order
ordinary differential equations are used (Shampine and Gordon 1975). Vectors qi and
q̇ i should be used outside the integrator only to determine dependent coordinates and
velocities using the constraint equations at the position and velocity levels, and also
to evaluate accelerations. Therefore, the independent coordinates qi and independent
velocities q̇ i should not be altered outside the integrator, and the dependent accelera-
tions q̈ d should not be integrated if an explicit integration method with a well-designed
error-check criterion is used (Shampine and Gordon 1975). This numerical proce-
dure is consistent with the Lagrange–D’Alembert principle in which the equations
of motion are formulated in terms of and solved for the independent accelerations,
which can be integrated to determine the independent coordinates and velocities. The
dependent velocities and accelerations in the Lagrange–D’Alembert principle are deter-
mined using the velocity transformation, as discussed in the remainder of this section
(Shabana 2019).

Embedding Technique The technique of Lagrange multipliers, which is the basis


of the augmented formulation, was originally introduced to handle systems subjected to
non-holonomic constraints. However, this technique has been used in the MBS literature
for both holonomic and non-holonomic systems. Using this approach, the dynamic
equations of motion are formulated in terms of redundant coordinates, which are related
by algebraic constraint functions. Consequently, constraint forces must appear in the
equations of motion of the system.
Another approach that can be used to eliminate constraint forces and obtain a number
of equations equal to the number of the system degrees of freedom is the embedding
technique. In this alternate approach, algebraic constraint functions are used to write
dependent variables in terms of independent variables. This leads to the definition of the
velocity transformation matrix, which allows for systematically eliminating the constraint
forces and obtaining a minimum number of equations of motion. The embedding tech-
( )T
nique has its roots in the Lagrange–D’Alembert principle Mq̈ − Qe − Qv + CTq 𝛌 𝛿q = 0.
This equation, as previously discussed in this chapter, is a statement of the virtual work
( )T
principle 𝛿W i = 𝛿W e + 𝛿W c , where 𝛿Wi = Mq̈ − Qv 𝛿q is the virtual work of the
inertia forces of the system, 𝛿We = QTe 𝛿q is the virtual work of the applied forces, and
( )T
𝛿Wc = − CTq 𝛌 𝛿q is the virtual work of the constraint forces. As discussed before, when
( )T
the dynamic equilibrium of the system is considered, one has 𝛿Wc = − CTq 𝛌 𝛿q = 0, and
the Lagrange–D’Alembert principle can be written as
( )T
Mq̈ − Qe − Qv 𝛿q = 0 (6.59)

For a virtual change in coordinates, the constraint functions lead to

Cq 𝛿q = Cqi 𝛿qi + Cqd 𝛿qd = 𝟎 (6.60)


252 Mathematical Foundation of Railroad Vehicle Systems

This equation allows writing the virtual change in dependent coordinates in terms of the
( )−1
virtual change in independent coordinates as 𝛿qd = − Cqd Cqi 𝛿qi , which can be used to
write the virtual change in all coordinates in terms of the virtual change of independent
coordinates as
[ ]
𝛿qi ⎡ I ⎤
𝛿q = = ⎢ ( )−1 ⎥ 𝛿qi = Bdi 𝛿qi (6.61)
𝛿qd ⎢− C Cqi ⎥⎦
⎣ qd

where Bdi is the velocity transformation matrix defined as


⎡ I ⎤
Bdi = ⎢ ( )−1 ⎥ (6.62)
⎢− C Cqi ⎥⎦
⎣ qd

One also has


q̇ = Bdi q̇ i , q̈ = Bdi q̈ i + 𝛄i (6.63)
( ( )
where 𝛄i = Ḃ di q̇ i . Substituting Eqs. 61 and 63 into Eq. 59, one obtains M Bdi q̈ + 𝛄i −
)T
Qe − Qv Bdi 𝛿qi = 0. Because the elements of vector qi are independent, their coefficients
( ( ) )
can be set equal to zero, leading to BTdi M Bdi q̈ i + 𝛄i − Qe − Qv = 𝟎, which upon rear-
( T ) ( )
ranging the terms can be written as Bdi MBdi q̈ i = BTdi Qe + Qv − M𝛄i . This equation can
be written as
( )
Mii q̈ i = BTdi Qe + Qv − M𝛄i (6.64)
where Mii = BTdi MBdi is the generalized mass matrix associated with the independent coor-
dinates. The number of equations in the preceding matrix equation is equal to the number
of the system degrees of freedom nd . Furthermore, no constraint forces appear in Eq. 64
because this equation is formulated in terms of the independent coordinates. In fact, one
can show that BTdi CTq 𝛌 = 𝟎. This is clear from the definition of the velocity transformation
matrix in Eq. 62, which leads to

[ ]⎡ I ⎤
Cq Bdi = Cqi Cqd ⎢ ( )−1 ⎥ = 𝟎 (6.65)
⎢− C Cqi ⎥⎦
⎣ qd

That is, the product of the transpose of the velocity transformation matrix Bdi and the con-
straint forces is always equal to zero.
Equation 64 is an alternate formulation that leads to a smaller number of equations as
compared to the augmented formulation. However, Eq. 64 has a dense and highly nonlinear
mass matrix Mii , and therefore, using sparse-matrix techniques with algorithms based on
the embedding technique does not lead to significant computational savings. Nonetheless,
when the independent coordinates qi are selected to be joint variables for open-chain sys-
tems, one can write the system coordinates in terms of the independent coordinates using
closed-form expressions. In this case of open-chain systems, using the Newton–Raphson
iterative procedure at the position level to determine dependent coordinates can be avoided.
Equations of Motion 253

Example 6.6 The equations of motion of the wheel/rail system formulated in


Example 6.5 using the augmented formulation can also be formulated using the
embedding technique, which leads to a number of equations equal to the number of
degrees of freedom of the system (Shabana et al. 2008). To this end, the constraints
at the acceleration level, r̈ wc = 𝟎, are used to write two dependent accelerations q̈ d in
terms of the independent acceleration q̈ i . In this wheel/rail example, the dependent
accelerations q̈ d can be selected to be R̈ wx and R̈ wy , while the independent acceleration q̈ i
can be selected as 𝜃̈ w . Using the condition r̈ wc = 𝟎 of Example 6.5, one can write
[ ] [ ] [ ]
R̈ wx −r cos 𝜙 w r 𝜃̇ w 𝜙̇ sin 𝜙
q̈ d = = ̈
𝜃 +
R̈ wy r sin 𝜙 r 𝜃̇ w 𝜙̇ cos 𝜙
Using this equation, the vector of the system accelerations can be written in terms of
independent acceleration in the form q̈ = Bdi q̈ i + 𝛄i , where in this case q̈ i reduces to
the scalar 𝜃̈ w . One therefore has
[ ] ⎡R̈ wx ⎤ ⎡−r cos 𝜙⎤ ⎡ r 𝜃̇ w 𝜙̇ sin 𝜙 ⎤
q̈ d ⎢ w⎥ ⎢ ⎥ ̈w ⎢ ̇ w ̇ ⎥
q̈ = = ⎢R̈ y ⎥ = ⎢ r sin 𝜙 ⎥ 𝜃 + ⎢r 𝜃 𝜙 cos 𝜙⎥
q̈ i ⎢ ̈w ⎥ ⎢ ⎥ ⎢ ⎥
⎣𝜃 ⎦ ⎣ 1 ⎦ ⎣ 0 ⎦
Using this equation, the velocity transformation matrix Bdi and quadratic velocity vector
𝛄i are recognized as
⎡−r cos 𝜙⎤ ⎡ r 𝜃̇ w 𝜙̇ sin 𝜙 ⎤
⎢ ⎥ ⎢ ⎥
Bdi = ⎢ r sin 𝜙 ⎥ , 𝛄i = ⎢r 𝜃̇ w 𝜙̇ cos 𝜙⎥
⎢ ⎥ ⎢ ⎥
⎣ 1 ⎦ ⎣ 0 ⎦
The equations of motion of the system were obtained in Example 6.5 as Mq̈ = Qe +
Qc , where the vectors and matrices in this equation were defined in Example 6.5. It is
clear that
⎡ Nx ⎤
[ ]⎢ ⎥
BTdi Qc = −r cos 𝜙 r sin 𝜙 1 ⎢ Ny ⎥=0
⎢ ⎥
⎣Nx r cos 𝜙 − Ny r sin 𝜙⎦
Therefore, substituting q̈ = Bdi q̈ i + 𝛄i into the equations of motion Mq̈ = Qe + Qc and
premultiplying by the transpose of the velocity transformation matrix BTdi , one obtains
(Shabana et al. 2008)
( w ) ( )
J + mw r 2 𝜃̈ w = M w − r Fxw cos 𝜙 + Fyw sin 𝜙 − mw gr sin 𝜙
This scalar equation, which does not have any constraint forces, is the system equation
of motion, which can be solved for the angular acceleration 𝜃̈ w . If no external forces
and moments are applied to the wheel, and if the wheel is represented by a cylinder
with a mass moment of inertia J w = mw r 2 /2, the preceding equation, upon using

(Continued)
254 Mathematical Foundation of Railroad Vehicle Systems

the kinematic relationship r 𝜃̇ w = (R − r) 𝜙̇ obtained in Example 6.5, reduces to


(3 (R − r) ∕2) 𝜙̈ + g sin 𝜙 = 0.
The wheel/rail problem considered in this example can be used to demonstrate that
the embedding technique has its roots in D’Alembert’s principle. By treating inertia
forces the same way as applied forces, one can eliminate constraint forces by equating
the moments of these forces about the contact point. This leads to one scalar equation
of motion because the system has one degree of freedom. This can be demonstrated
using the free-body diagram shown in Figure 4. The moments of the applied force about
( )
contact point c is M w − r Fxw cos 𝜙 + Fyw sin 𝜙 − mw gr sin 𝜙. The moment of the inertia
forces about the same point is J w 𝜃̈ w − mw R̈ wx r cos 𝜙 + mw R̈ wy r sin 𝜙. By using the con-
straint equations at the acceleration level to write accelerations R̈ wx and R̈ wy in terms of
independent acceleration 𝜃̈ w , it can be shown that the moment of the inertia forces about
( )
the contact point is J w + mw r 2 𝜃̈ w . Therefore, the same equation of motion obtained
previously in this example using the embedding technique can be obtained by equat-
ing the moments of the inertia and applied forces. However, the embedding technique
can be used as the basis for developing general MBS algorithms that are based on the
connectivity conditions and do not require using free-body diagrams.

6.8 WHEEL/RAIL CONSTRAINT CONTACT FORCES


Wheel/rail contact formulations were discussed in Chapter 5. Two fundamentally different
methods can be used to mathematically define wheel/rail interaction forces: the constraint
approach, which does not allow for wheel/rail separation; and the elastic approach, which
allows for wheel/rail separations. In the constraint approach, the wheel/rail normal con-
tact force is determined as a reaction force. In the elastic approach, on the other hand, the
wheel/rail contact force is defined using a compliant force model formulated using assumed
stiffness and damping coefficients. The two approaches (constraint and elastic) considered
in this book require the formulation and solution of a system of algebraic equations.

Wheel and Rail Contact Surfaces As discussed in the preceding chapters, formulat-
ing the three-dimensional wheel/rail contact problem requires the parameterization of the
wheel and rail surfaces. As shown in Figure 5 and discussed in Chapter 5, the wheel and
rail surface parameters can be written in a vector form as
[ ]T
s = sw1 sw2 sr1 sr2 (6.66)
where superscripts w and r refer, respectively, to the wheel and rail surfaces. In the preced-
ing equation, sw1 and sw2 are, respectively, the wheel lateral and angular surface parameters;
and sr1 and sr2 are, respectively, the rail longitudinal and lateral surface parameters, as shown
in Figure 5. This surface-parameter description is used to capture the three-dimensional
nature of the wheel/rail contact problem.
The locations of a potential contact point Pk , k = w, r, on the wheel and rail surfaces
with respect to the wheel and rail coordinate systems can be written, respectively, as
( ) [ ( ) ( ) ( )]T
ukP sk1 , sk2 = xk sk1 , sk2 yk sk1 , sk2 zk sk1 , sk2 , k = w, r, as discussed in Chapters 3–5.
The global position of the potential contact point can be written as rkP = Rk + Ak ukP , k = w, r.
Equations of Motion 255

Zwp
w
s 2
sw1
Ywp

Xwp
gwp(sw1 )

(a) (b)

Figure 6.5 Wheel and rail surfaces.

The tangent and normal vectors to the wheel and rail surfaces can be defined in the body
coordinate systems, respectively, as
k 𝜕ukP k 𝜕ukP k k
t1 = , t2 = , nk = t1 × t2 , k = w, r (6.67)
𝜕sk1 𝜕sk2
k k
These vectors can be defined in the global coordinate system as tk1 = Ak t1 , tk2 = Ak t2 and
nk = Ak nk , k = w, r.

Contact Constraint Approach In the contact constraint approach, as discussed in


Chapter 5, wheel/rail separation is not allowed. In this case, the wheel is assumed to have
five degrees of freedom with respect to the rail. The freedom of the wheel to move with
respect to the rail along the normal to the surface at the contact point is eliminated. Because
four geometric surface parameters are introduced as non-generalized coordinates that must
be solved for in order to determine the location of the contact point, the elimination of one
relative degree of freedom between the wheel and rail requires imposing the following five
algebraic constraint equations for a non-conformal contact j
( )
⎡ rw − rr j ⎤
⎢ (P P

( w r wj rj ) ⎢ wT r )j ⎥
C q ,q ,s ,s = ⎢ 1
j t n ⎥=𝟎 (6.68)
⎢ ( T )j ⎥
⎢ tw2 nr ⎥
⎣ ⎦
which can be written in an alternate and equivalent form, upon using the definition rwr
P =
rwP − rrP , as
( )j
⎡ tr1T rwr
P

⎢ ( T )j ⎥
⎢ tr2 rwr
P

( ) ⎢( T )j ⎥
Cj qw , qr , swj , srj = ⎢ nr rwrP
⎥=𝟎 (6.69)
⎢( )j

⎢ twT nr ⎥
⎢ 1 ⎥
⎢ ( w r )j ⎥
⎣ t 2 •n ⎦
256 Mathematical Foundation of Railroad Vehicle Systems

As discussed in Chapter 5, by using the contact constraint equations written in terms of the
[ ]T
surface parameters s = sw1 sw2 sr1 sr2 , the constraint functions that include the effect of
all joints and specified motion trajectories as well as the contact constraints can be writ-
ten in a vector form as C(q, s) = 0, where in this case s contains the surface parameters
[ T T T ]T
associated with all the contacts in the system – that is, s = s1 s2 · · · sncc – ncc is
[ ]T
the number of contacts, and sj = swj 1
wj rj rj
s2 s1 s2 is the vector of surface parameters
associated with contact j. It follows that 𝛿C = Cq 𝛿q + Cs 𝛿s = 0, where Cq and Cs are, respec-
tively, the constraint Jacobian matrices associated with vectors q and s. Therefore, one has
𝛌T (Cq 𝛿q + Cs 𝛿s) = 0, where 𝛌 is the vector of Lagrange multipliers. Adding this equation to
( )
the equation defining the Lagrange–D’Alembert principle 𝛿qT Mq̈ − Qe − Qv = 0, previ-
ously obtained in this chapter, one has
( )
𝛿qT Mq̈ + CTq 𝛌 − Qe − Qv + 𝛿sT CTs 𝛌 = 0 (6.70)

Using a procedure similar to the one discussed in Section 6.2 of this chapter, one can show
that the preceding equation leads to

Mq̈ + CTq 𝛌 = Qe + Qv , CTs 𝛌 = 𝟎 (6.71)

Differentiating the constraint equations C(q, s) = 0 twice with respect to time, one obtains

Cq q̈ + Cs s̈ = Qd (6.72)

where Qd is the vector that absorbs terms that are not linear in accelerations. Combining
the preceding two equations, one has the augmented form of the equations of motion in the
case of the wheel/rail contact constraint approach as
⎡ M 𝟎 CTq ⎤ ⎡q̈ ⎤ ⎡Qe + Qv ⎤
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ 𝟎 𝟎 Cs ⎥ ⎢ s̈ ⎥ = ⎢ 𝟎 ⎥
T (6.73)
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣Cq Cs 𝟎 ⎦ ⎣𝛌⎦ ⎣ Qd ⎦
There are important observations about the augmented form of the equations of motion of
Eq. 73, which accounts for the wheel/rail contact constraint conditions. These observations
are summarized as follows:

● Equation 73 can be solved for the second time derivatives of the generalized coordinates
̈ the second time derivatives of the non-generalized coordinates s̈ , and the vector of
q,
Lagrange multipliers 𝛌. If the constraint functions are linearly independent and the mass
matrix is nonsingular, one can show that the coefficient matrix in Eq. 73 has a full row
rank and is nonsingular.
● It is clear from Eq. 73 that there are no inertia or applied forces associated with the vector
of surface parameters s. Therefore, the surface parameters are geometric parameters and
can be considered non-generalized coordinates, as discussed in the literature (Shabana
and Sany 2001).
● Equation 73 implies that the system differential equations of motion and the wheel/rail
contact constraints must be solved simultaneously for the system generalized and
non-generalized coordinates as well as the constraint forces. This is necessary to cor-
rectly account for couplings between the generalized and non-generalized coordinates.
Equations of Motion 257

In the augmented contact constraint formulation, the total vector of coordinates can be
[ ]T
written as p = qT sT .
[ ]
● The system constraint Jacobian matrix can be written as Cp = 𝜕C∕𝜕p = 𝜕C∕𝜕q 𝜕C∕𝜕s ,
[ ]
which can be written as Cp = Cq Cs , where Cq = 𝜕C/𝜕q and Cs = 𝜕C/𝜕s. Using the
system constraint Jacobian matrix Cp to identify the dependent and independent coordi-
nates, the system degrees of freedom can include both generalized and non-generalized
[ ]T
coordinates. In this case, the total vector of coordinates can be written as p = pTi pTd ,
where pi and pd are, respectively, the vectors of independent and dependent coordinates.
Both vectors can include generalized and non-generalized coordinates.
● Because four surface parameters are used for each contact, which is described using five
algebraic equations, the constraint conditions for one contact lead to one independent
reaction force. This force is the normal contact constraint force. Equation 73 shows that
CTs 𝛌 = 𝟎 (Shabana and Sany 2001). In this equation, there are five Lagrange multipliers
jT
associated with five contact constraint equations. For a contact j, one has Cs 𝛌j = 𝟎, j =
1, 2, … , ncc . That is, each contact j introduces five independent contact constraints and
jT
five Lagrange multipliers. The equation Cs 𝛌j = 𝟎 for contact j represents a system of four
scalar equations in five unknown Lagrange multipliers 𝛌j . If the constraint equations are
j
linearly independent, the rank of the 5 × 4 constraint Jacobian matrix Cs associated with
jT
the surface parameters of contact j is 4, and therefore, the equation Cs 𝛌j = 𝟎 ensures that
there is only one independent Lagrange multiplier. That’s is, the five contact constraints,
which eliminate the relative degree of freedom of the wheel to move with respect to the
rail in the normal direction, can be used to determine only one independent constraint
force. This constraint force is used in the augmented contact formulation to determine
the wheel/rail normal contact force. This normal force is used with the wheel and rail
geometry to determine the tangential creep forces, as discussed in Chapter 5. The creep
forces acting on the wheel and rail can be entered into the equations of motion using
vector Qe .
It is clear, therefore, that when the augmented contact constraint formulation is used,
the normal contact force is obtained as a reaction force. Because this normal force is used
to determine the tangential creep forces included in vector Qe on the right-hand side of
Eq. 73, the value of the Lagrange multipliers associated with the contact constraints from
the previous time step can be stored and used to determine the tangential creep force. This
approximation has proven to work well and allows avoiding the use of an iterative procedure
to solve Eq. 73, which represents a large system of equations in the case of complex railroad
vehicle systems.

Elimination of Surface Parameters Four of the contact constraint equations can be


used to eliminate surface parameters (non-generalized coordinates) by writing them in
terms of generalized coordinates. In this case, contact between the wheel and rail can be
represented by one algebraic constraint equation imposed on the generalized coordinates.
This constraint equation eliminates the relative motion between the wheel and rail in a
direction normal to the contact surfaces at the contact point. For simplicity, the superscript
j that refers to the contact number is dropped in the remainder of this section. To explain
the procedure for eliminating the surface parameters, the constraint equations in Eq. 69 can
258 Mathematical Foundation of Railroad Vehicle Systems

be rearranged and written as


T
P ⎤
⎡ tr1 rwr
⎢ rT wr ⎥
[ ] ⎢ t2 rP ⎥
( w r w r) Cd ⎢ T ⎥
C q ,q ,s ,s = n
= ⎢ tw1 nr ⎥ = 𝟎 (6.74)
C ⎢ wT r ⎥
⎢ t2 n ⎥
⎢ rT wr ⎥
⎣n rP ⎦
where
[ T ]T }
Cd (qw , qr , sw , sr ) = tr1 rwr tw1 nr tw2 nr = 𝟎,
T T T
P tr2 rwr
P
T
(6.75)
Cn (qw , qr , sw , sr ) = nr rwr
P =0

Therefore, one has

Cds 𝛿s = −Cdq 𝛿q, Cds ṡ = −Cdq q,


̇ Cds s̈ = −Cdq q̈ + 𝛄dC (6.76)

in which Cds = 𝜕Cd ∕𝜕s, Cdq = 𝜕Cd ∕𝜕q, and 𝛄dC = −Ċ ds ṡ − Ċ dq q.
̇ Assuming the case of non-
conformal contact in which matrix Cds is a nonsingular matrix, Eq. 76 leads to
( −1 ) ( −1 ) ( −1 )
𝛿s = − Cds Cdq 𝛿q, ṡ = − Cds Cdq q, ̇ s̈ = − Cds Cdq q̈ + 𝛄dC (6.77)

where 𝛄dC = Cds 𝛄dC . Using Eq. 77, the second equation of Eq. 75 leads to
−1

( ( −1 ))
𝛿Cn = Csn 𝛿s + Cqn 𝛿q = Cqn − Csn Cds Cdq 𝛿q = 0, ⎫
( ( )) ⎪
̇Csn = Csn ṡ + Cqn q̇ = Cqn − Csn Cds −1 Cdq q̇ = 0 ⎬ (6.78)
( ) ( ( ))
̈Csn = Csn s̈ + Cqn q̈ + Ċ sn ṡ + Ċ qn q̇ + Csn 𝛄d = Cqn − Csn Cds −1 Cdq q̈ + 𝛾 n = 0⎪

C C
( )
where 𝛾Cn = Ċ sn ṡ + Ċ qn q̇ + Csn 𝛄dC is a vector that absorbs terms that are not linear in acceler-
ations, Csn = 𝜕Cn ∕𝜕s, and Cqn = 𝜕Cn ∕𝜕q. Using the preceding equation, the following algo-
rithm can be used to eliminate surface parameters and represent the wheel/rail interaction
using one constraint equation defined by Cn :

1. Given the configurations of the wheel and rail defined, respectively, by the generalized
coordinates qw and qr , the four constraint equations Cd of Eq. 75 can be solved itera-
tively using a Newton–Raphson algorithm to determine the four surface parameters sw
and sr . The wheel and rail coordinates qw and qr are assumed to be known from the
numerical integration of the system equations of motion. In the Newton–Raphson itera-
tions, the equation (𝜕Cd /𝜕s)Δs = − Cd is solved for the Newton differences Δs, which are
used to update s until convergence is achieved. The result of this numerical procedure
is equivalent to writing s = s(qw , qr ).
2. After determining the surface parameters, their derivatives can be determined using
Eq. 77 in terms of the derivatives of the wheel and rail coordinates. Furthermore, because
the surface parameters are numerically computed using wheel and rail coordinates, the
Jacobian matrix of constraint Cn defined by differentiation with respect to the gener-
( ( −1 ))
alized coordinates can be assumed equal to Cqn − Csn Cds Cdq . This is the Jacobian
matrix that must be used in the position and velocity analysis steps.
Equations of Motion 259

3. At the acceleration analysis step, the Jacobian matrix of constraint Cn is again assumed
( ( −1 ))
to take the form Cqn − Csn Cds Cdq , while the element in vector Qd associated with
constraint Cn is set equal to −𝛾Cn .
The use of this algorithm can be analytically explained by writing the Lagrange–
D’Alembert principle as (Shabana et al. 2008)
( T
) ( T )
𝛿qT Mq̈ + Cdq 𝛌d + Cnq T 𝛌n − Q + 𝛿sT Cds 𝛌d + Cns T 𝛌n = 0 (6.79)

where 𝛌d and 𝛌n are Lagrange multipliers associated with constraint equations Cd and Cn ,
respectively. The first equation in Eq. 77 can be written as
𝛿s = Bdn 𝛿q (6.80)
( )−1
where Bdn = − Cds Cdq . Therefore, Eq. 79 reduces to
( ( )T )
𝛿qT Mq̈ + Cnq + Cns Bdn 𝛌n − Q = 0 (6.81)

The forces associated with contact constraints Cd are eliminated from this equation as a
consequence of eliminating the surface parameters. This is the result of using the identity
( ( ) )T
T T T T −1 T
Cdq 𝛌d + Bdn Cds 𝛌d = Cdq 𝛌d + − Cds Cdq Cds 𝛌d = 𝟎. Also, as previously mentioned,
the second and third equations in Eq. 77, ṡ = Bdn q̇ and s̈ = Bdn q̈ + 𝛄dC , respectively, can be
used to eliminate ṡ and s̈ from the time derivatives of the constraint equation. Therefore, by
eliminating the non-generalized coordinates (surface parameters) and constraint equations
Cd , one can write the augmented form of the equations of motion as
[ T
][ ] [ ]
M Cq q̈ Qe + Qv
= (6.82)
Cq 𝟎 𝛌 Qd

where Cq is the Jacobian matrix of all constraint equations C after eliminating s and Cd , 𝛌
is the vector of Lagrange multipliers associated with constraint equations C, and Qd is the
vector that results from the differentiation of C twice with respect to time: that is, Cq q̈ = Qd .
Constraint functions C, constraint Jacobian matrix Cq , and vector Qd are implicit functions
of the surface parameters, and therefore, the coupling between the generalized coordinates
and the non-generalized surface parameters is not ignored by eliminating s and Cd .
The computer implementation of the procedure discussed in this section for the elimi-
nation of the surface parameters has shown that no clear computational advantage results
from eliminating surface parameters s and constraint functions Cd . This is mainly due to
the significant mathematical operations required for this elimination, as demonstrated by
the analysis presented in this section. Therefore, when the constraint contact formulation
is used, it is recommended to use Eq. 73, which is written in terms of both the generalized
and non-generalized coordinates.

6.9 WHEEL/RAIL ELASTIC CONTACT FORCES


The analysis presented in the preceding section shows that when the constraint contact for-
mulation is used, there is no wheel/rail separation, and the wheel is assumed to remain in
260 Mathematical Foundation of Railroad Vehicle Systems

contact with the rail regardless of the magnitude of applied forces. The constraint contact
formulation can provide insight and be more efficient in many simulation scenarios, par-
ticularly in the case of smooth dynamics and when the wheel remains in contact with the
rail. Nonetheless, such a constraint contact formulation can have serious limitations when
examining some important scenarios, including derailments. For this reason, it is impor-
tant to have an implementation of another approach that allows for wheel/rail separation,
as discussed in Chapter 5. In this approach, referred to as the elastic contact formulation,
wheel/rail interaction forces are described using a compliant force model.
In the elastic contact formulation, no kinematic constraints are imposed on the relative
motion between the wheel and rail, and consequently, the unconstrained wheel is assumed
to have six degrees of freedom with respect to the rail if no other motion constraints are
introduced. In this formulation, the compliant force model is written as the sum of elas-
tic and damping forces expressed in terms of assumed stiffness and damping coefficients,
respectively. Whereas in the contact constraint formulation, the location of the contact
point is determined by solving constraint equations at the position level, in the elastic con-
tact formulation, several methods can be used to determine the location of the wheel/rail
contact point. These methods include discrete nodal search, the use of lookup tables, and
solving a set of algebraic equations. In this section, the method of solving algebraic equations
to determine the location of the contact point online is used to formulate wheel/rail contact
forces based on the elastic contact approach. The implementation of the other two methods,
nodal search and lookup tables, is straightforward, and the same force model described in
this section can be used once the location of the wheel/rail contact point is determined.
As discussed in Chapter 5, the algebraic equations used in the elastic force formulation
to determine the contact point online are the same as four of the equations used in the
constraint contact formulation. These equations are the four algebraic equations given in
the first equation of Eq. 75, which are reproduced here for convenience (Escalona 2002;
Pombo and Ambrosio 2003; Shabana et al. 2005)
[ T ]T
tw1 nr tw2 nr = 𝟎
T T T
g = tr1 rwr
P tr2 rwr
P (6.83)
where all the vectors that appear in this equation are the same as previously defined. As
discussed in Chapter 5, given the wheel and rail configurations defined, respectively, by
vectors qw and qr at a given point in time, the preceding nonlinear algebraic equations
can be solved for the surface parameters s. In this case, g = g(s) = 0 is a system of four
algebraic equations that can be solved for the four unknown surface parameters. Using the
Newton–Raphson method, the system of algebraic equations (𝜕g/𝜕s)Δs = − g(s) is solved
iteratively for the Newton-differences Δs, which are used to update the surface parameters
until convergence is achieved. The converged solution defines the surface parameters that
can be used to determine the locations and velocities of the contact points on the wheel and
rail using the equations
}
rkP = Rk + Ak ukP ,
(6.84)
vPk = ṙ kP = Ṙ k + 𝛚k × ukP , k = w, r
Using the first equation in Eq. 84, the distance (penetration) between surfaces along the
normal is evaluated using the equation
T
𝛿 = rwr nr (6.85)
Equations of Motion 261

In the case of wheel/rail penetrations, normal contact forces can be calculated using the
assumptions of non-conformal contact and Hertz contact theory. The following expression
for wheel/rail normal contact force F n has been used in the literature (Shabana et al. 2004)
( )
Fn = Fns + Fnd = −kns 𝛿 1.5 − cnd |𝛿| 𝛿̇ (6.86)
where 𝛿 measures the penetration, F ns is the normal Hertzian (elastic) contact force, F nd
is the normal damping force, kns is the Hertzian stiffness coefficient that depends on the
surface curvatures and properties of the materials of the two bodies in contact, and cnd is a
damping coefficient. The time rate of penetration 𝛿̇ is evaluated using the velocity equation
( )T
in Eq. 84 as 𝛿̇ ≈ ṙ wP − ṙ rP nr = ṙ wr
T
r
P n . The absolute value of penetration |𝛿| is included in
the expression of the normal damping force to ensure zero normal contact force when 𝛿 = 0.
( )T
Discussion of defining the time rate of penetration, using the equation 𝛿̇ ≈ ṙ wP − ṙ rP nr =
ṙ wr
T
r
P n , was provided in the preceding chapter.
After determining the normal contact force F n using the elastic contact formulation, this
normal force can be used with material properties and the dimensions of the contact area
to determine the tangential creep forces. The tangential creep forces are entered into the
equations of motion using the vector of generalized applied forces Qe . It is clear from these
force definitions that normal contact force determined using the elastic approach is not a
constraint force. Therefore, the algebraic equations of Eq. 83 should not be viewed as con-
straints imposed on the motion of the system. These algebraic equations are not enforced
at the velocity and acceleration levels, and there are no constraint forces or Lagrange mul-
tipliers associated with them. Instead, a compliant force model that allows for wheel/rail
separation is used.

6.10 OTHER FORCE ELEMENTS


In addition to wheel/rail contact forces, other forces can have significant effects on the
dynamics and stability of railroad vehicle systems. For example, suspension forces influ-
ence the vehicle critical speed; beyond this critical speed, the vehicle becomes unstable.
Most rail vehicles have two types of suspensions: the primary suspension, which connects
the wheelsets to the bogie frame or equalizer bar, and the secondary suspension used to
connect the car body to the bogie. These suspension elements can be represented mathemat-
ically using combinations of spring-damper-actuator and friction elements. Furthermore,
bearings are used to mount wheelsets on bogie frames, and bushings are widely used in
developing virtual railroad models. Figure 6 shows examples of the force elements used
in railroad vehicles. This section discusses the formulations of several force elements used
in the virtual prototyping of railroad vehicle systems.

Spring-Damper-Actuator Element This force element can be used to connect two


arbitrary bodies in the vehicle system using a spring, damper, and/or actuator. The stiffness
and damping coefficients and actuator force can be represented using linear or nonlin-
ear functions defined analytically or numerically using spline functions. Figure 7 shows
two bodies i and j connected by the spring-damper-actuator force element. The attachment
point on body i is denoted as Pi , while the attachment point on body j is Pj . The stiffness
262 Mathematical Foundation of Railroad Vehicle Systems

ij
V
ij
θ
i
i Z
X
i
i Z
X
i
Pi k Y
i
Y fa cr Tm kr

c Pj j j
Z Y Z Z
j
j
Y j
X j
Y
o X X

(a) Spring-damper-actuator element (b) Rotational spring-damper-motor element

bj
Z
i
i Z bj
X bj X
Y j
P
Pi
i
Y cs i
b3
ks
bi
Pj Z
Z Y j
Z bi Qi X bi
j Y
Y Pi
o X X
j

(c) Series spring-damper element (d) Bushing element

Figure 6.6 Force elements. (a) Spring-damper-actuator element; (b) rotational


spring-damper-motor element; (c) series spring-damper element; (d) bushing element.

coefficient of the spring is assumed to be k, and its undeformed and current lengths are lo
and l, respectively. The damping coefficient is assumed to be c, and the actuator force is f a .
The force produced by this element can be written as
( )
Fs = k l − l0 + cl̇ + fa (6.87)
( )
ij ij ij j | j|
This force can be used to define the force vector Fs = Fŝ
rP , where ̂
rP = riP − rP ∕ |riP − rP |
| |
is a unit vector along the line connecting points Pi and Pj on bodies i and j, respectively.
Equations of Motion 263

i
Z
i
X
Pi k
i
Y
fa

c Pj Z
j
Z Y
j
j
Y
X
o X

Figure 6.7 Spring-damper-actuator element.

Therefore, the virtual work of the spring-damper-actuator force element can be written as
( )T ( )
ij ijT ij ij j
𝛿We = −Fs 𝛿rP = − Fŝ rP 𝛿riP − 𝛿rP
( )T ( )
ij j
= − Fŝ rP LiP 𝛿qi − LP 𝛿qj (6.88)
[ ]
where LkP = I −Ak ũ Pk Gk , k = i, j, and ũ Pk is the skew-symmetric matrix associated with
vector ukP , which defines the position of point Pk , k = i, j, with respect to the coordinate
system of body k. Using Eq. 88, the virtual work of the spring-damper-actuator force
ij T jT j
element can be written as 𝛿We = Qie 𝛿qi + Qe 𝛿qj , where Qie and Qe are the generalized
spring-damper-actuator forces associated, respectively, with the generalized coordinates of
bodies i and j given by
[( ) ] [ ]
( Qie R) ij
Fs ⎡ ij
Fs ⎤⎫
Qie=
T
−LiP Fŝ
ij
= ( i) = −
rP =− ⎢ ( ) ⎥⎪
Qe 𝜃 −Gi ũ iP Ai Fs
T T T ij
⎢ GiT uiP × Fijs ⎥⎪
⎣ ⎦⎪
( ) ⎬
[ ]
) ⎡ Qe ⎤ ⎡ ⎤
j
( Fs
ij ij
Fs ⎪
j jT
Qe = LP Fŝ
ij ⎢
rP = ( ) = R⎥
= ⎢ ( ) ⎥ ⎪
⎢ Qj ⎥ −Gj ũ P Aj Fs
T jT T ij
⎢ Gj ujP × Fijs ⎥
T

⎣ e 𝜃⎦ ⎣ ⎦ ⎭
(6.89)
k
where ukP = Ak ukP , k = i, j, and Gk and G are, respectively, the matrices that relate angular
velocity vectors 𝛚k and 𝛚k to the time derivatives of the orientation parameters, as previ-
ously discussed in this book. Since the spring-damper-actuator force formulation presented
k
in this section is written in terms of the two matrices Gk and G , k = i, j, such a formulation
can be used with any set of orientation parameters.

Rotational Spring-Damper-Motor Element The coefficients that enter into the def-
inition of the rotational spring-damper-motor element can also be defined analytically or
numerically as linear or nonlinear functions of the system coordinates and time. The torque
264 Mathematical Foundation of Railroad Vehicle Systems

ij
V
ij
θ
i
Z
i
X

i
Y

cr Tm kr

j
Z

j
Y
j
X

Figure 6.8 Rotational spring-damper-motor element.

produced by this force element when connecting two arbitrary bodies i and j in the system,
as shown in Figure 8, is written in terms of the rotational spring stiffness coefficient kr ;
ij
current and initial relative angles between the two bodies 𝜃 ij and 𝜃o , respectively, about the
element axis v ; damping coefficient cr ; and motor torque T m . The axis of rotation vij can be
ij

defined on body i as vij = Ai vij , where vij is a constant vector defined in the coordinate sys-
tem of body i. The torque produced by the rotational spring-damper-motor element about
axis vij can be written as
( )
Ts = kr 𝜃 ij − 𝜃o + cr 𝜃̇ ij + Tm
ij
(6.90)
ij
The virtual work of this torque can be written as 𝛿We = −Ts 𝛿𝜃 ij , where 𝛿𝜃 ij =
T ( )
vij 𝛿𝛑i − 𝛿𝛑j , and 𝛿𝛑i and 𝛿𝛑j are virtual rotations about the axes of the Cartesian
coordinate system, which can be written as 𝛿𝛑i = 𝛚i 𝛿t = Gi 𝛿𝛉i and 𝛿𝛑j = 𝛚j 𝛿t = Gj 𝛿𝛉j .
ij
Therefore, the virtual work 𝛿We = −Ts 𝛿𝜃 ij can be written as
ij T ( )
𝛿We = −Ts 𝛿𝜃 ij = −Ts vij 𝛿𝛑i − 𝛿𝛑j
T ( ) T jT
= −Ts vij Gi 𝛿𝛉i − Gj 𝛿𝛉j = Qie 𝛿qi + Qe 𝛿qj (6.91)
j
where Qie and Qe are the generalized forces associated with the coordinates of bodies i and
j, respectively, and defined as
( )
[( ) ] [ ] [ ]
Q i
𝟎 ⎡ Qje ⎤ 𝟎
e R
Qe = ( i ) =
i
T ( ) ,
j ⎢
Qe = ( ) = R⎥
T ( ) (6.92)
Qe 𝜃 −Gi Ts vij ⎢ Qj ⎥ Gj Ts vij
⎣ e ⎦ 𝜃
It is clear from the definitions of the generalized forces in this equation that the rotational
spring-damper-motor element does not produce generalized forces associated with the ref-
erence translations of the two bodies connected by this element.
Equations of Motion 265

i
Z
i
X

Pi
i
Y cs
ks

Pj
Z Y Z
j

j
o Y
X X
j

Figure 6.9 Series spring-damper element.

Series Spring-Damper Element The series spring-damper element, shown in


Figure 9, consists of a spring and damper connected in series. As shown in Figure 9, the
spring-damper can be used to connect two arbitrary bodies i and j. The attachment points
on bodies i and j are assumed to be Pi and Pj , respectively. The spring stiffness and damping
coefficients are assumed to be ks and cs , respectively. These coefficients can be defined
analytically or numerically as linear or nonlinear functions of the coordinates. In the series
spring-damper force element, widely used in railroad vehicle systems, the spring force
is equal to the damping force: that is, Fs = ks ds = cs ḋ d , where ds and dd are, respectively,
the spring deflection and relative displacement between the two ends of the damper, as
shown in Figure 9. The total displacement of the element is the sum of these two displace-
ments: that is, dt = ds + dd . Differentiating this equation with respect to time, one obtains
ḋ t = ḋ s + ḋ d . In the special case where ks is assumed to be constant, one has Ḟ s = ks ḋ s , and
consequently, the equation ḋ t = ḋ s + ḋ d can be written as ḋ t = Ḟ s ∕ks + Fs ∕cs , which leads to
( )
̇Fs + ks Fs = ks ḋ t (6.93)
cs
The total element displacement dt can be written in terms of the generalized coordinates
ij j
of bodies i and j. This can be accomplished by defining the vector rP = riP − rP , which can
be used to define the relative velocity along the unit vector that connects the two points Pi
and Pj . Equation 93 can be integrated numerically during dynamic simulations to define
force F s . Once this force is defined, a procedure similar to the one used for the translational
spring-damper-actuator force element can be used to determine the generalized forces
associated with the generalized coordinates of the two bodies connected by the series
spring-damper element.
( ) ( )
If ks is not constant, one has Ḟ s = ks ḋ s +(k̇ s ds , from) which ḋ s = F∕k
̇ s + k̇ s ∕ks ds . This
( ) ( )2
equation can be written as ḋ = F∕k ̇ s + k̇ ∕ k s F . In this case, ḋ = ḋ + ḋ leads to
s s s t s d
the more general equation
( ( )2 )
k̇ s cs + ks
Ḟ s + Fs = ks ḋ t (6.94)
ks cs
This equation reduces to Eq. 93 if ks is constant.
266 Mathematical Foundation of Railroad Vehicle Systems

Bushing Element The bushing element can be used to connect two arbitrary bodies i
and j in the system, as shown in Figure 10. The attachment points on bodies i and j are
assumed to be points Pi and Pj , respectively. An assumption is made that the relative trans-
lation and rotation between the two bodies at the bushing definition point are small. For
the convenience of describing the relative rotations between the two bodies, two bushing
coordinate systems on bodies i and j are introduced. These two coordinate systems can be
selected such that their axes are assumed to be initially parallel. The bushing elements,
often made of rubber materials and widely used in developing virtual prototyping models
in a wide class of applications including railroad vehicle systems, produce restoring and
damping forces in different translational and rotational directions. The bushing stiffness
and damping coefficients can be described in the virtual models analytically or numerically
using linear and/or nonlinear characteristics.

bj
Z
bj
bj X
Y j
P

i
b3

bi
Z

bi Qi X
bi
Y
Pi

Figure 6.10 Bushing element.

To develop the mathematical model of the bushing element and associate the relative
displacements and rotations between the two bodies connected by the bushing element
with the bushing material properties, the bushing coordinate systems X bi Y bi Z bi and
X bj Y bj Z bj , which are assumed to be rigidly attached to bodies i and j connected by the
bushing element, as shown in Figure 10, are used. Because the bushing coordinate system
X bi Y bi Z bi is assumed to be attached to body i, this coordinate system can be defined using
two points Pi and Qi on body i. These two points can be used to define one of the axes of
the bushing coordinate system in body i coordinate system X i Y i Z i using the unit vector
i ( ) | |
b3 = uiP − uiQ ∕ |uiP − uiQ |, where uiP and uiQ are, respectively, the constant local position
| |
vectors of points Pi and Qi . Using this axis, a simple procedure
[ can be used] to complete an
bi i i i i i
orthogonal triad defined by the orthogonal matrix A = b1 b2 b3 , where b1 and b2
i
are two unit vectors orthogonal to the unit vector b3 . The axes of the bushing coordinate
system X bi Y bi Z bi can be defined in the global coordinate system using the transformation
bi [ ] i
Abi = Ai A = bi1 bi2 bi3 , where bik = Ai bk , k = 1, 2, 3. The transformation matrix Abi
Equations of Motion 267

is used to define the relative displacements and rotations between the two bodies in the
bushing coordinate system, thereby giving these displacements and rotations obvious
meanings that can be associated with the bushing material properties, including the
stiffness and damping in different directions.
To define, on body j, the bushing coordinate system X bj Y bj Z bj , which has axes that are
initially parallel to the axes of the bushing coordinate
[ system] X bi Y bi Z bi on body i, before
bj jT bi j j j
the start of the simulation, matrix Ao = Ao Ao = b1 b2 b3 is defined, where subscript
o
bj
o refers to the initial configuration. Matrix Ao defines the axes of a coordinate system
X bj Y bj Z bj rigidly attached to body j that coincide with the axes of the bushing coordinate
system X bi Y bi Z bi rigidly attached to body i in the initial configuration. As body j moves
bj
with respect to body i, one can define the matrix Abj = Aj Ao that defines the orientation of
the coordinate system X bj Y bj Z bj in the global coordinate system.
The relative displacement and velocity between two points Pj and Pi can be defined,
respectively, in the bushing coordinate system X bi Y bi Z bi as
( )
T ji T j ⎫
dbji = Abi rP = Abi Rj + Aj uP − Ri − Ai uiP ⎪
( )⎬ (6.95)
vbji = Abi ṙ P = Abi Ṙ j + 𝛚j × uP − Ṙ i − 𝛚i × uiP ⎪
T ji T j

The relative rotations between the two bodies can be associated with the bushing axes using
T
matrix Abji = Abi Abj . In the case of small rotations, matrix Abji can be written in terms of
bji bji bji T
small relative rotations 𝛽1 , 𝛽2 , and 𝛽3 , which can be extracted from matrix Abji = Abi Abj ,
which can be computed in a straightforward manner if the coordinates of the two bodies are
known from the numerical integration. Similarly, the components of the relative angular
T ( )
velocity vector can be defined in the bushing coordinate system as 𝛚bji = Abi 𝛚j − 𝛚i . The
translational bushing stiffness and damping properties can, in general, be defined using the
3 × 3 stiffness and damping matrices Kbt and Dbt . Similarly, the rotational bushing stiffness
and damping coefficients can be defined using the 3 × 3 stiffness and damping matrices Kbr
and Dbr . The elements of the stiffness and damping matrices Kbt , Kbr , Dbt , and Dbr can be deter-
mined using experimental measurements. Using these matrices, the bushing restoring and
damping forces and moments can be defined in the bushing coordinate system X bi Y bi Z bi ,
respectively, as
b b
FR = Kbt dbji + Dbt vbji , M𝛼 = Kbr 𝛃bji + Dbr 𝛚bji (6.96)
[ ]T
where 𝛃bji = 𝛽1bji 𝛽2bji 𝛽3bji are the small relative rotations between the two bodies. The
force and moment vectors of Eq. 96 can be defined, respectively, in the global coordinate
b b
system XYZ as FbR = Abi FR and Mb𝛼 = Abi M𝛼 . Using these force and moment vectors, the
generalized bushing forces associated with the generalized coordinates of the two bodies
can be written as
T ( ) ⎫
QiR = FbR , Qi𝜃 = Gi Mb𝛼 + uiP × FbR
( )⎪ ⎬ (6.97)
j j T j
QR = −FbR , Q𝜃 = −Gj Mb𝛼 + uP × FbR ⎪

The generalized bushing forces defined in this equation can be added to the vector of the
system generalized applied forces Qe .
268 Mathematical Foundation of Railroad Vehicle Systems

6.11 TRAJECTORY COORDINATES

The absolute Cartesian coordinates used in this chapter to formulate the nonlinear dynamic
equations of motion of a railroad vehicle system allow developing general computational
MBS algorithms. This is particularly true when such algorithms are generalized for flexible
body dynamics, an important generalization for strength and durability investigations that
are routinely conducted for tank car strength and rail integrity. Other sets of coordinates,
however, have been used in the literature to develop specialized railroad vehicle system
software. While using specialized algorithms can make the implementation of flexible-body
formulations more cumbersome, these specialized algorithms can be efficient in develop-
ing software for specific applications that require the use of fewer degrees of freedom for
each body. Examples of these applications are in the important area of longitudinal train
dynamics (LTD) in which the focus is on the longitudinal coupler and braking forces as
well as energy consumption.

Trajectory Coordinates and Track Geometry This section presents the formulation
of the equations of motion using trajectory coordinates, which are used to develop special-
ized railroad vehicle and LTD algorithms. As demonstrated in Chapter 3, trajectory and
Cartesian coordinates are related by coordinate transformation, thereby allowing for sys-
tematically converting the equations of motion expressed in terms of Cartesian coordinates
to equations expressed in terms of trajectory coordinates. As explained in Chapter 3 and
shown in Figure 11, the general displacement of a rigid body i in the system can be described
using the six trajectory coordinates
[ ]T
pi = si yir zir 𝜓 ir 𝜙ir 𝜃 ir (6.98)

where si is the arc length coordinate of the track space curve; yir and zir are, respectively,
the lateral and vertical displacements of the origin Oi of the body coordinate system X i Y i Z i
relative to the trajectory body coordinate system X ti Y ti Z ti that follows the body, as shown
in Figure 11; and 𝜓 ir , 𝜙ir , and 𝜃 ir are, respectively, the yaw, roll, and pitch angles that define
the orientation of the body coordinate system X i Y i Z i with respect to the trajectory body
coordinate system X ti Y ti Z ti . As discussed in Chapter 3, if arc length parameter si is known,
the location of the origin of the trajectory body coordinate system X ti Y ti Z ti can be defined
and written as Rti = Rti (si ).

Euler Angles and Geometry As discussed in Chapters 3 and 4, another set of Euler
angles, expressed as field variables in arc length parameter si , can be used to describe track
geometry. Therefore, arc length parameter si is sufficient to define the orientation of the
trajectory body coordinate system X ti Y ti Z ti in terms of three Euler angles 𝜓 ti (si ), 𝜃 ti (si ), and
𝜙ti (si ) about the three axes Z ti , −Y ti , and −X ti . These three Euler angles, used to define
the orientation of the trajectory body coordinate system X ti Y ti Z ti , are considered geometric
field variables that only depend on the arc length parameter si and should be distinguished
from the time-dependent yaw, roll, and pitch Euler angles 𝜓 ir (t), 𝜙ir (t), and 𝜃 ir (t) previously
introduced to define the orientation of the body coordinate system with respect to its trajec-
tory body coordinate system. As explained in Chapter 3, these geometry Euler angles can
Equations of Motion 269

X
Y
O

Figure 6.11 Trajectory coordinates.

be used to construct the following transformation matrix that defines the orientation of the
trajectory body coordinate system X ti Y ti Z ti :

⎡cos 𝜓 ti cos 𝜃 ti − sin 𝜓 ti cos 𝜙ti + cos 𝜓 ti sin 𝜃 ti sin 𝜙ti − sin 𝜓 ti sin 𝜙ti − cos 𝜓 ti sin 𝜃 ti cos 𝜙ti ⎤
A = ⎢ sin 𝜓 ti cos 𝜃 ti cos 𝜓 ti cos 𝜙ti + sin 𝜓 ti sin 𝜃 ti sin 𝜙ti
ti
cos 𝜓 ti sin 𝜙ti − sin 𝜓 ti sin 𝜃 ti cos 𝜙ti ⎥
⎢ ⎥
⎣ sin 𝜃 ti
− cos 𝜃 sin 𝜙
ti ti
cos 𝜃 ti cos 𝜙ti ⎦
(6.99)

If the geometry of the track space curve is given, the field-variable Euler angles 𝜓 ti = 𝜓 ti (si ),
𝜃 ti = 𝜃 ti (si ), and 𝜙ti = 𝜙ti (si ) can be determined for a given value of arc length parameter
[ ( ) ( ) ( )]T
si . That is, vector 𝛉ti = 𝜓 ti si 𝜃 ti si 𝜙ti si at a given si is assumed to be known
[ ( ) ( ) ( )]T
from the specified track geometry. Determining 𝛉 = 𝜓 ti si 𝜃 ti si 𝜙ti si
ti for a given
value of si using the track data file produced by the track preprocessor was discussed in more
detail in Chapter 4.

Position, Velocity, and Acceleration Vectors It was shown in Chapter 3 that, using
the trajectory body coordinate system, the global position vector of the origin Oi of the body
i coordinate system can be defined as
( ) ( )
Ri = Rti si + Ati si uir (6.100)
[ ]T
where vector uir = 0 yir zir is the position vector of reference point Oi with respect
to the origin of the trajectory body coordinate system X ti Y ti Z ti . As discussed in Chapter 3,
because the location of point Oi in the longitudinal direction is assumed to be defined using
arc length si , the first element of vector uir is equal to zero. The absolute velocity and accel-
eration vectors of reference point Oi , respectively, can be obtained by differentiating Eq. 100
once and twice with respect to time. This leads to

Ṙ i = Li ṗ i , ̈ i = Li p̈ i + 𝛄iR
R (6.101)
270 Mathematical Foundation of Railroad Vehicle Systems

[(( ti i ) ( ti i ) ir ) ti ti ]
where Li = 𝜕R ∕𝜕s + 𝜕A ∕𝜕s u a2 a3 𝟎 is a 3 × 6 matrix in which a2ti and a3ti
are the second and third columns of the transformation matrix Ati of Eq. 99, 𝛄iR absorbs the
terms that are not linear in generalized accelerations, and 0 is the 3 × 3 null matrix.
The absolute angular velocity vector of body i can be written, as explained in Chapter 3,
as 𝛚i = 𝛚ti + 𝛚ir , which is the sum of the absolute angular velocity vector 𝛚ti = Gti 𝛉̇ ti of
the trajectory body coordinate system and vector 𝛚ir = Ati Gir 𝛉̇ ir , which defines the angular
velocity of the body with respect to the trajectory body coordinate system. Therefore, the
absolute angular velocity 𝛚i of body i can be written in terms of the time derivatives of the
trajectory coordinates as

𝛚i = Gti 𝛉̇ ti + Ati Gir 𝛉̇ ir = Hi ṗ i (6.102)

where matrices Gti and Gir are, respectively, defined as

⎡0 sin 𝜓 ti − cos 𝜓 ti cos 𝜃 ti ⎤ ⎡0 cos 𝜓 ir − sin 𝜓 ir cos 𝜙ir ⎤


⎢ ⎥ ⎢ ⎥
Gti = ⎢0 − cos 𝜓 ti − sin 𝜓 ti cos 𝜃 ti ⎥ , Gir = ⎢0 sin 𝜓 ir cos 𝜓 ir cos 𝜙ir ⎥
⎢ ⎥ ⎢ ⎥
⎣1 0 − sin 𝜃 ti ⎦ ⎣1 0 sin 𝜙ir ⎦
(6.103)
[( ( )) ( )]
and Hi = Gti 𝜕𝛉ti ∕𝜕si 𝟎 𝟎 Ati Gir is a 3 × 6 matrix in which 0 is the three-
dimensional zero vector. Differentiating Eq. 102 with respect to time, the absolute angular
acceleration vector 𝛂i of body i can be expressed in terms of the second time derivatives
of the trajectory coordinates as 𝛂i = Hi p̈ i + 𝛄i𝛼 , where 𝛄i𝛼 is a vector that absorbs terms
that are not linear in accelerations (Sanborn et al. 2007; Sinokrot et al. 2008; Shabana
et al. 2008). This equation that defines 𝛂i can also be used to write the absolute angular
acceleration vector defined in the body coordinate system X i Y i Z i as
T i
𝛂i = Ai 𝛂i = H p̈ i + 𝛄i𝛼 (6.104)
i T T T T
where H = Ai Hi = Air Ati Hi , 𝛄i𝛼 = Ai 𝛄i𝛼 , and Air is the matrix that defines the orien-
tation of the body coordinate system X i Y i Z i with respect to the trajectory body coordinate
system X ti Y ti Z ti . This matrix, as explained in Chapter 3, can be written in terms of the Euler
[ ]T
angle generalized coordinates 𝛉ir = 𝜓 ir 𝜙ir 𝜃 ir as

⎡cos 𝜓 ir cos 𝜃 ir − sin 𝜓 ir sin 𝜙ir sin 𝜃 ir − sin 𝜓 ir cos 𝜙ir cos 𝜓 ir sin 𝜃 ir + sin 𝜓 ir sin 𝜙ir cos 𝜃 ir ⎤
Air = ⎢sin 𝜓 ir cos 𝜃 ir + cos 𝜓 ir sin 𝜙ir sin 𝜃 ir cos 𝜓 ir cos 𝜙ir sin 𝜓 ir sin 𝜃 ir − cos 𝜓 ir sin 𝜙ir cos 𝜃 ir ⎥
⎢ ⎥
⎣ − cos 𝜙ir sin 𝜃 ir sin 𝜙ir cos 𝜙ir cos 𝜃 ir ⎦
(6.105)

Acceleration Transformation Combining the second equation in Eqs. 101 and 104,
one obtains (Sanborn et al. 2007; Sinokrot et al. 2008)
[ ] [ i] [ ]
R̈i L 𝛄i
= i p̈ i + Ri (6.106)
𝛂 i
H 𝛄𝛼

In this acceleration transformation, the absolute angular acceleration vector 𝛂i is used


instead of vector 𝛂i because the mass matrix associated with the absolute Cartesian
Equations of Motion 271

[ ]T
accelerations R ̈ iT 𝛂iT is constant in the Newton–Euler formulation of the equations
of motion. Equation 106 can be used with the Newton–Euler equations to obtain the
equations of motion in terms of trajectory coordinates without using any small-angle
assumptions; therefore, these equations of motion include all the nonlinear effects that
can be significant in some motion scenarios.

Trajectory-Coordinate Equations of Motion The form of the Newton–Euler


equations given in Eq. 47 is used with the acceleration transformation of Eq. 106 to obtain
the nonlinear equations of motion expressed in terms of trajectory coordinates. To this
end, the absolute accelerations in Eq. 106, written in terms of trajectory accelerations, are
substituted in Eq. 47, and Eq. 47 is premultiplied by the transpose of the coefficient matrix
of p̈ i in Eq. 106. This leads to (Sanborn et al. 2007; Sinokrot et al. 2008; Shabana et al. 2008)
Mipp p̈ i = Qipe + Qipv (6.107)
where

T T
Mipp = mi Li Li + H i I i𝜃𝜃 Hi
T T ⎪
Qipe = Li Fie + H i Mie ⎬ (6.108)
T T [ i i ( i i )]⎪
Qipv = −mi Li 𝛄iR − H i I 𝜃𝜃 𝛄 𝛼 + 𝛚 × I 𝜃𝜃 𝛚 ⎭
i

Mass matrix Mipp in this equation is not constant as it is the case when absolute Cartesian
coordinates are used with the Newton–Euler equations of Eq. 47.
In the case of constrained motion, kinematic algebraic constraint functions can be formu-
lated in terms of trajectory coordinates. In this case, the augmented form of the equations
of motion can be written as (Shabana et al. 2008)
[ ][ ] [ ]
Mpp CTp p̈ Qpe + Qpv
= (6.109)
Cp 𝟎 𝛌 Qd
[ ]T
where p = p1T p2T … pnb T is the vector of system trajectory coordinates, nb is the
total number of bodies in the system, Mpp is the system mass matrix associated with the
trajectory coordinates, Qpe is the vector of system applied forces associated with the tra-
jectory coordinates, Qpv is the vector of centrifugal and Coriolis forces, Cp is the Jacobian
matrix of the kinematic constraint equations, 𝛌 is the vector of Lagrange multipliers, and
Qd is the vector resulting from the differentiation of the system constraint equations twice
with respect to time.

Trajectory Coordinate Constraints. Using trajectory coordinates allows developing


simpler formulations for some types of motion constraints. For example, if it is desired to
specify the forward velocity V of a vehicle along the track centerline, one can use the con-
stant forward velocity constraint, which can be expressed in terms of trajectory coordinate
si as (Shabana et al. 2008)
C = si − V t − si0 = 0 (6.110)
where si0 is the initial arc length coordinate. The constraint condition of Eq. 110 ensures that
the origin of the trajectory body coordinate system moves with a constant velocity ṡ i = V
272 Mathematical Foundation of Railroad Vehicle Systems

along the track centerline. The constraint of Eq. 110 is a linear function in si . This linearity
cannot be achieved if absolute Cartesian coordinates are used.
Similar linear conditions can be developed when constraints are imposed on the yaw, roll,
or pitch angles of the vehicle. For example, the constraint that specifies the yaw motion of
a body with respect to the trajectory body coordinate system can be simply written in the
case of the trajectory coordinates as
C = 𝜓 ir − 𝜓0ir = 0 (6.111)
where 𝜓0iris the yaw angle in the initial configuration. In general, one can specify any of
the trajectory coordinates as a function of time using the constraint equation
C = pik − f (t) = 0 (6.112)
where pik is the kth trajectory coordinate in vector pi
of Eq. 98, and f (t) is a given function of
time. Using the simple trajectory constraints of Eqs. 111 and 112 leads to constant elements
in the constraint Jacobian matrix and allows for a straightforward elimination of depen-
dent coordinates. Using trajectory coordinates to formulate other constraint equations is
demonstrated by the following example.

Example 6.7 The wheelset shown in Figure 12 is denoted as body i and subjected
to the following three kinematic constraint conditions: (i) the forward velocity of the
wheelset along the track centerline is assumed to be g(t); (ii) the wheelset is assumed
to rotate about its own axle with a constant angular velocity 𝜔i (pitch rotation); and
(iii) the wheelset is assumed to have zero displacement along the Z ti axis of the trajec-
tory body coordinate system. The three constraint equations imposed on the motion of
the wheelset can be written mathematically as
⎡C1 ⎤ ⎡ si − g (t) ⎤
⎢ ⎥ ⎢ ⎥
C (p, t) = ⎢C2 ⎥ = ⎢𝜃 ir − 𝜃oir − 𝜔i t⎥ = 𝟎
⎢C ⎥ ⎢ zir − zir ⎥
⎣ 3⎦ ⎣ o ⎦

Figure 6.12 Trajectory coordinate constraints.


Equations of Motion 273

where 𝜃oir is the initial value of the pitch rotation and zoir is the initial value of coordinate
zir . Differentiating the constraint equations with respect to time, one obtains Cp ṗ + Ct =
𝟎, where the constraint Jacobian matrix and vector Ct are defined as
⎡1 0 0 0 0 0⎤ ⎡−dg (t) ∕dt⎤
⎢ ⎥ ⎢ ⎥
Cp = ⎢0 0 0 0 0 1⎥ , Ct = ⎢ −𝜔i ⎥
⎢ ⎥ ⎢ ⎥
⎣0 0 1 0 0 0⎦ ⎣ 0 ⎦
The constraint equations at the acceleration level are defined as Cp p̈ = Qd , where
[ ]T
Qd = d2 g (t) ∕dt2 0 0
Using these kinematic relationships, the embedding technique can be used systemati-
cally to obtain three equations of motion for the wheelset. It is clear that the degrees of
freedom of this system are yir , 𝜓 ir , and 𝜙ir : that is, the vector of independent coordinates
[ ]T [ ]T
is pi = yir 𝜓 ir 𝜙ir , and the vector of dependent coordinates is pd = si zir 𝜃 ir .
[ ]
The constraint Jacobian matrix can be written in a partitioned form as Cp = Cpd Cpi ,
where
⎡1 0 0⎤ ⎡0 0 0⎤
⎢ ⎥ ⎢ ⎥
Cpd = ⎢0 0 1⎥ , Cpi = ⎢0 0 0⎥
⎢ ⎥ ⎢ ⎥
⎣0 1 0⎦ ⎣0 0 0⎦
It is clear from this partitioning that the system velocities can be written in terms of the
independent velocities as
⎡ ṡ i ⎤ ⎡dg (t) ∕dt⎤ ⎡0 0 0⎤ ⎡dg (t) ∕dt⎤
⎢ ir ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ ẏ ⎥ ⎢ ẏ
ir
⎥ ⎢1 0 0⎥ ir
⎡ ẏ ⎤ ⎢
0 ⎥
⎢ ż ir ⎥ ⎢ 0 ⎥ ⎢0 0 0⎥ ⎢ ir ⎥ ⎢ 0 ⎥
ṗ = ⎢ ir ⎥ = ⎢ ⎥=⎢ ⎥ ⎢𝜓̇ ⎥ + ⎢ ⎥
⎢𝜓̇ ⎥ ⎢ 𝜓̇ ir ⎥ ⎢0 1 0⎥ ⎢ ir ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎣ 𝜙̇ ⎦ ⎢ ⎥
⎢ 𝜙̇ ir ⎥ ⎢ 𝜙̇ ir ⎥ ⎢0 0 1⎥ ⎢ 0 ⎥
⎢ ̇ ir ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 𝜃 ⎦ ⎣ 𝜔i ⎦ ⎣0 0 0⎦ ⎣ 𝜔 i ⎦
Using this equation, the total vector of accelerations can be written in the form of the
second equation in Eq. 63 as p̈ = Bdi p̈ i + 𝛄i , where
2
⎡0 0 0⎤ ⎡d2 g (t) ∕dt ⎤
⎢ ⎥ ⎢ ⎥
⎢1 0 0⎥ ⎢ 0 ⎥
⎢0 0 0⎥ ⎢ 0 ⎥
Bdi = ⎢ ⎥, 𝛄i = ⎢ ⎥
⎢0 1 0⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢0 0 1⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎣0 0 0⎦ ⎣ 0 ⎦
Substituting p̈ = Bdi p̈ i + 𝛄i in Eq. 107, given as Mipp p̈ i = Qipe + Qipv , and premultiplying
by the transpose of the velocity transformation matrix Bdi , one obtains three equations of

(Continued)
274 Mathematical Foundation of Railroad Vehicle Systems

motion from which the constraint forces are eliminated. These equations can be written
in the form of Eq. 64 as
( )
Mii p̈ i = BTdi Qpe + Qpv − Mpp 𝛄i
where Mii = BTdi Mpp Bdi .

6.12 LONGITUDINAL TRAIN DYNAMICS (LTD)

LTD algorithms are used in the analysis of railroad vehicle coupler and braking forces as
well as energy consumption of long trains traveling for long distances. The train models can
include 100 or 200 rail cars. For this reason, it is necessary to develop efficient algorithms
in which fewer degrees of freedom are used for each rail car. In most LTD algorithms, a
single-degree-of-freedom rail car is used in modeling long trains. This single degree of free-
dom is assumed to be the longitudinal coordinate si for a rail car i. This section discusses
formulations that employ one or two degrees of freedom for the rail car.

Single-Degree-of-Freedom LTD Model When using a single degree of freedom for


each rail car, it is assumed that the degree of freedom is the arc length coordinate si and all
other coordinates are constrained. In this case, reference point Oi of the rail car does not
deviate from the track centerline. Therefore, no other translational or rotational degrees of
freedom are allowed except the motion in the longitudinal direction. Using these simplify-
ing assumptions, one can write the equations of motion of each rail car i in the train in terms
of the arc length coordinate si only. Since in this case yir = yiro and zir = zoir , where subscript
o refers to the initial configuration, the position of reference point Oi can be written in the
track coordinate system, which is assumed to be fixed and to coincide with the global coor-
( ) ( ) [ ]T
dinate system, as Ri = Rti si + Ati si uiro , where uiro = 0 yiro zoir is a constant vector.
In this case, the absolute velocity of the reference point can be written as
( ) ( )
Ṙ i = Ṙ ti si + Ȧ ti si uiro = Li ṡ i (6.113)
( ) ( )
where Li = 𝜕Rti ∕𝜕si + 𝜕Ati ∕𝜕si uiro . Since all the rotations of the rail car with respect
to the trajectory body coordinate system X ti Y ti Z ti are constrained – that is, 𝜓 ir , 𝜙ir , and 𝜃 ir
remain constant – the angular velocity of the rail car 𝛚i is equal to the angular velocity of its
( )
trajectory body coordinate system 𝛚ti = G 𝛉̇ ti . In this case, one has 𝛉̇ ti = 𝜕𝛉ti ∕𝜕si ṡ i , and
ti

(Shabana et al. 2008)


⎡ sin 𝜃 ti 0 −1⎤
ti⎢ ⎥
G = ⎢− cos 𝜃 sin 𝜙 − cos 𝜙
ti ti ti 0⎥ (6.114)
⎢ ⎥
⎣ cos 𝜃 cos 𝜙 − sin 𝜙ti 0 ⎦
ti ti

Therefore, the angular velocity of rail car i can be written as


( ti )
̇
ti ti ti 𝜕𝛉
𝛚 =𝛚 =G 𝛉 =G
i ti
ṡ i (6.115)
𝜕si
Equations of Motion 275

By differentiating Eqs. 113 and 115 with respect to time, one obtains Eq. 106, which can be
written in this case as
[ ] [ i] [ i]
̈i
R L 𝛄R
i s̈ +
i
= (6.116)
𝛂i
H 𝛄i𝛼
where
( ti ) ( 2 ti ( 2 ti ) )
𝜕𝛉 𝜕 R 𝜕 A ( i )2 ⎫
i
H =G
ti
, 𝛄 i
= L̇ i ṡ i = + uiro ṡ ⎪
𝜕s i R
𝜕s i2
𝜕s i2

(( )( ) ( 2 ti ))
ti
( ) ⎬ (6.117)
𝜕G 𝜕𝛉 ti ti 𝜕 𝛉 2 ⎪
𝛄i𝛼 = + G ̇
s i
𝜕si 𝜕si 𝜕si2


Substituting Eq. 116 into the Newton–Euler equations given in Eq. 47, and premultiplying
by the transpose of the coefficient vector of s̈ i in Eq. 116, the equation of motion of the
one-degree-of-freedom rail car can be written as (Sanborn et al. 2007; Sinokrot et al. 2008;
Shabana et al. 2008)
i i
Mpp s̈ = Qipe + Qipv (6.118)
T
i = mi Li Li + H I H , iT i i T iT i T
where Mpp 𝜃𝜃 Qipe = Li Fie + H Me , and Qipv = −mi Li 𝛄iR −
iT
[ i ( i )]
H I𝜃𝜃 𝛄𝛼 + 𝛚i × I𝜃𝜃 𝛚i .
i

Example 6.8 In the case of a rail car i negotiating a curve with zero grade and superel-
evation, one has 𝜃 ti = 𝜙ti = 0, and the only nonzero track geometry angle is 𝜓 ti . In this
( ) ( ) ( ) ( )( )
case, one has Li = 𝜕Rti ∕𝜕si + 𝜕Ati ∕𝜕si uiro = 𝜕Rti ∕𝜕si + 𝜕Ati ∕𝜕𝜓 ti 𝜕𝜓 ti ∕𝜕si uiro .
[ ]T ( )
Using Eq. 99 and the definition of uiro as uiro = 0 yiro zoir , one has 𝜕Rti ∕𝜕si =
[ ]T ( ) [ ]T
cos 𝜓 ti sin 𝜓 ti 0 and 𝜕Ati ∕𝜕𝜓 ti uiro = −yiro cos 𝜓 ti sin 𝜓 ti 0 . Therefore, Li can
be defined as (Shabana et al. 2008)
⎡cos 𝜓 ti ⎤
( ) ⎢ ⎥
Li = 1 − yiro 𝜓sti ⎢ sin 𝜓 ti ⎥
⎢ ⎥
⎣ 0 ⎦
( )
where 𝜓sti = 𝜕𝜓 ti ∕𝜕si . In this special case of the track geometry, one has
⎡0 0 −1⎤
ti⎢ ⎥
G = ⎢0 −1 0 ⎥
⎢ ⎥
⎣1 0 0⎦
i ti ( ti i )
Therefore, the coefficient matrix H = G 𝜕𝛉 ∕𝜕s in Eq. 116 can be written as

( ) ⎡0⎤
i ti 𝜕𝛉ti ⎢ ⎥
H =G =⎢0⎥
𝜕si ⎢ ti ⎥
⎣𝜓s ⎦

(Continued)
276 Mathematical Foundation of Railroad Vehicle Systems

ti
Since 𝜕G ∕𝜕si = 𝟎, one can evaluate vectors 𝛄iR and 𝛄i𝛼 of Eq. 117, respectively, as
( 2 ti ( 2 ti ) )
𝜕 R 𝜕 A ( i )2
𝛄iR = + uiro ṡ
𝜕s i2
𝜕s i2

⎡ ⎡ sin 𝜓 ti ⎤ ⎡cos 𝜓 ti ⎤⎤
⎢ ti ( ) ⎢ ⎥ ⎢ ⎥⎥ ( )2
= − ⎢𝜓s 1 − yiro 𝜓sti ⎢− cos 𝜓 ti ⎥ + yiro 𝜓sti ⎢ sin 𝜓 ti ⎥⎥ ṡ i
⎢ ⎢ ⎥ ⎢ ⎥⎥
⎣ ⎣ 0 ⎦ ⎣ 0 ⎦⎦
and
(( )( ) ( )) ⎡0⎤
𝜕G
ti
𝜕𝛉ti ti 𝜕 2 𝛉ti ( i )2 ⎢ ⎥ ( i )2
𝛄i𝛼 = +G ṡ = ⎢ 0 ⎥ ṡ
𝜕si 𝜕si 𝜕si2 ⎢ ti ⎥
⎣𝜓ss ⎦
2
where 𝜓ssti = 𝜕 2 𝜓 ti ∕𝜕si .
i =
Using the matrices and vectors developed in this example, the mass matrix Mpp
T iT i i
mi Li Li + H I𝜃𝜃 H in Eq. 118 can be evaluated as (Shabana et al. 2008)
i T iT i i
Mpp = mi Li Li + H I𝜃𝜃 H
( )2 ( )2
= mi 1 − yiro 𝜓sti + izz 𝜓sti
The scalars Qipe , and Qipv can be evaluated, respectively, as
T iT i
Qipe = Li Fie + H Me
( )( i ) i
= 1 − yiro 𝜓sti Fex cos 𝜓 ti + Fey
i
sin 𝜓 ti + 𝜓sti M ez
and
T iT
[ i ( i )]
Qipv = −mi Li 𝛄iR − H I𝜃𝜃 𝛄i𝛼 + 𝛚i × I𝜃𝜃 𝛚i
( ( ) ) ( )2
= mi yiro 1 − yiro 𝜓sti − izz 𝜓sti 𝜓ssti ṡ i
( i )
where in this special example of planar motion, the gyroscopic moment 𝛚i × I𝜃𝜃 𝛚i is
[ i ]
i T is the force vector defined in the global coordinate
equal to zero, Fie = Fex i
Fey Fez
i
[ ]T
i i i
system XYZ, and M𝛼 = M 𝛼x M 𝛼y M 𝛼z is the moment vector defined in the rail car
coordinate system X i Y i Z i . Substituting Mppi
, Qipe , and Qipv into the equation of motion of
Eq. 118, one obtains (Shabana et al. 2008)
( ( )2 ( ti )2 ) i ( )( i ) i
mi 1 − yiro 𝜓sti + Izz i
𝜓s s̈ = 1 − yiro 𝜓sti Fex cos 𝜓 ti + Fey
i
sin 𝜓 ti + 𝜓sti M ez
[ ( ) ] ( )2
+ mi yiro 1 − yiro 𝜓sti − Izz
i
𝜓sti 𝜓ssti ṡ i
If yiro = 0, the preceding equation reduces to
[ ( ti )2 ] i ( )2
i
mi + Izz
i
𝜓s s̈ = Fex
i
cos 𝜓 ti + Fey
i
sin 𝜓 ti + 𝜓sti M ez − Izz
i
𝜓sti 𝜓ssti ṡ i

Furthermore, if the track is assumed to be tangent, 𝜓 ti = 𝜓sti = 𝜓ssti = 0, and the preced-
ing equation is further simplified to mi s̈ i = Fex
i .
Equations of Motion 277

Two-Degree-of-Freedom Rail Car Model The single-degree-of-freedom rail car


model has been widely used in the development of LTD algorithms designed to effi-
ciently solve problems of long freight trains that can have more than 100 rail cars.
These problems focus on the longitudinal coupler and braking forces as well as energy
consumptions. However, the simplified single-degree-of-freedom rail car model cannot be
used to capture important phenomena such as wheelset hunting, which is the result of
coupling between the lateral and yaw displacements of the wheelsets. Therefore, simplified
two-degree-of-freedom models have been developed in the literature to study the hunting
phenomenon as well as wheelset stability.
In the two-degree-of-freedom model discussed in this section, the lateral displacement yir
and yaw angle 𝜓 ir are assumed to be the two independent coordinates of the wheelset. As
in the case of the single-degree-of-freedom model, the two-degree-of-freedom model can
be derived as a special case from the general trajectory coordinate formulation previously
introduced in this chapter. This can be achieved by imposing proper kinematic constraint
conditions on the other four trajectory coordinates si , zir , 𝜙ir , and 𝜃 ir . In the general case
of profiled wheels and rails, the displacement in the normal direction zir and roll angle
𝜙ir depend on the lateral displacement yir and yaw angle 𝜓 ir : that is, zir = zir (yir , 𝜓 ir ) and
𝜙ir = 𝜙ir (yir , 𝜓 ir ). If the forward velocity and pitch rotations of the wheelset are prescribed,
one can write two constraints on si and 𝜃 ir . These two constraint equations can be writ-
ten, respectively, as C1 = si − sio − Vt = 0 and C2 = 𝜃 ir − 𝜃oir − 𝜔p t = 0, where V and 𝜔p are,
respectively, the specified forward velocity and pitch angular rotation of the wheelset, which
are assumed in this section to be constant in order to develop a model that can be used in
stability and hunting analysis at prescribed velocities. The other two constraints on zir and
𝜙ir can be written, respectively, as C3 = zir − f 1 (yir , 𝜓 ir ) = 0 and C4 = 𝜙ir − f 2 (yir , 𝜓 ir ) = 0,
where f 1 and f 2 are two functions that can be specified depending on the profile and track
geometries. It is important to point out that in the case of general track geometry, these
two functions also depend on the longitudinal arc length coordinate si . Other simplifying
assumptions are also made in the literature. Since hunting stability is, for the most part,
investigated in the case of tangent tracks, the two functions f 1 and f 2 are assumed in this
section to have the forms f 1 = f 1 (yir , 𝜓 ir ) and f 2 = f 2 (yir , 𝜓 ir ), respectively: that is, these two
functions are not dependent on other coordinates, including the arc length parameter si .
The four constraint equations can then be written in vector form as

⎡C1 ⎤ ⎡ si − sio − Vt ⎤
⎢ ⎥ ⎢ ir ⎥
⎢C2 ⎥ ⎢ 𝜃 − 𝜃oir − 𝜔p t ⎥
C (p, t) = ⎢ ⎥ = ⎢ ir ( ir ir ) ⎥ = 𝟎 (6.119)
⎢C3 ⎥ ⎢ z − f1 y , 𝜓 ⎥
⎢C ⎥ ⎢𝜙ir − f (yir , 𝜓 ir )⎥
⎣ 4⎦ ⎣ 2 ⎦

The constraint equations at the velocity level can be written as

⎡ ṡ i − V ⎤
⎢ ⎥
𝜃̇ − 𝜔p
ir
̇C (p, t) = Cp ṗ + Ct = ⎢⎢ ( ) ir ( ) ir ⎥

⎢ ż − 𝜕f1 ∕𝜕y ẏ − 𝜕f1 ∕𝜕𝜓 𝜓̇ ⎥
ir ir ir

⎢𝜙̇ ir − (𝜕f ∕𝜕yir ) ẏ ir − (𝜕f ∕𝜕𝜓 ir ) 𝜓̇ ir ⎥


⎣ 2 2 ⎦
278 Mathematical Foundation of Railroad Vehicle Systems

⎡1 ⎡ ṡ i ⎤
0 0 0 0 0⎤ ⎢ ir ⎥
⎢ ⎥ ⎢ ẏ ⎥ ⎡ −V ⎤
⎢0 0 0 0 0 1⎥ ⎢ ż ir ⎥ ⎢−𝜔 ⎥
=⎢ ⎥ ⎢ ⎥ + ⎢ p⎥ = 𝟎
(
⎢0 − 𝜕f ∕𝜕yir ) (
1 − 𝜕f1 ∕𝜕𝜓 ir
)
0 0⎥ ⎢𝜓̇ ⎥ ⎢ 0 ⎥
ir
⎢ 1 ⎥ ⎢ ̇ ir ⎥ ⎢ ⎥
⎢0 − (𝜕f ∕𝜕yir ) ( ) 𝜙
0 − 𝜕f2 ∕𝜕𝜓 ir 1 0⎥⎦ ⎢ ̇ ir ⎥
⎣ 0 ⎦
⎣ 2
⎣𝜃 ⎦
(6.120)

In this equation, the 4 × 6 constraint Jacobian matrix Cp and the four-dimensional vector
Ct can be recognized as

⎡1 0 0 0 0 0⎤
⎢ ⎥ ⎡ −V ⎤
⎢0 0 0 0 0 1⎥⎥ ⎢−𝜔 ⎥

Cp = ⎢ ( ) ( ) ⎥, Ct = ⎢ p ⎥ (6.121)
⎢0 − 𝜕f1 ∕𝜕yir 1 − 𝜕f1 ∕𝜕𝜓 ir 0 0⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦
⎢0 − (𝜕f ∕𝜕yir ) ( )
0 − 𝜕f2 ∕𝜕𝜓 ir 1 0⎥⎦
⎣ 2

The constraint equations at the acceleration level can be written as

⎡ s̈ i ⎤
⎢ ⎥
⎢ 𝜃̈ ir ⎥
̈C (p, t) = Cp p̈ − Qd = ⎢ ⎥ (6.122)
⎢ z̈ ir − (𝜕f ∕𝜕yir ) ÿ ir − (𝜕f ∕𝜕𝜓 ir ) 𝜓̈ ir − D ⎥
⎢ 1 1 s1 ⎥
⎢ ir ( ) ir ( ) ir ⎥
⎣𝜙̈ − 𝜕f2 ∕𝜕y ÿ − 𝜕f2 ∕𝜕𝜓 𝜓̈ − Ds
ir ir ⎦
2

2 2
where Dsk = (𝜕 2 fk ∕𝜕yir )(ẏ ir )2 + (𝜕 2 fk ∕𝜕𝜓 ir )(𝜓̇ ir )2 + 2(𝜕 2 fk ∕𝜕yir 𝜕𝜓 ir )ẏ ir 𝜓̇ ir , k = 1, 2, and
vector Qd is defined as

Qd = [0 0 Ds1 Ds2 ]T (6.123)

The constraint Jacobian matrix Cp and vector Qd can be used to obtain the augmented form
of the equations of motion of this two-degree-of-freedom system. This augmented form, in
the case of the trajectory coordinates, is defined by Eq. 109.
Alternatively, the embedding technique can be used to obtain two equations of motion
that do not include any constraint forces. This can be achieved by writing dependent
accelerations in terms of independent accelerations. It is clear from the definition of the
Equations of Motion 279

constraint Jacobian matrix that the sub-Jacobians matrices associated with the dependent
[ ]T [ ]T
and independent coordinates pd = si zir 𝜙ir 𝜃 ir and pi = yir 𝜓 ir can be written as

⎡1 0 0 0⎤ ⎡ 0 0 ⎤
⎢0 0 0 1⎥⎥ ⎢ 0 0 ⎥
Cpd =⎢ , Cpi = ⎢ ( ) ( )⎥ (6.124)
⎢0 1 0 0⎥ ⎢− 𝜕f1 ∕𝜕y ir − 𝜕f1 ∕𝜕𝜓 ⎥
ir
⎢ ⎥ ⎢ ( ) ( )⎥
⎣0 0 1 0⎦ ⎣− 𝜕f2 ∕𝜕yir − 𝜕f2 ∕𝜕𝜓 ir ⎦

Therefore, the total vector of accelerations can be written in terms of the independent accel-
erations as

⎡ s̈ ⎤ ⎡
i
0 0 ⎤ ⎡ 0 ⎤
⎢ ir ⎥ ⎢ ⎥ ⎢ ⎥
⎢ ÿ ⎥ ⎢ 1 0 ⎥ ⎢ 0 ⎥
⎢ ir ⎥ ⎢( ) ( )⎥ [ ir ] ⎢ ⎥
̈
⎢ ⎥ ⎢ 1
z 𝜕f ∕𝜕yir 𝜕f ∕𝜕𝜓 ir
⎥ ̈
y ⎢−D ⎥
+⎢ ⎥
1 s1
p̈ = ⎢ ⎥ = ⎢ ⎥ (6.125)
⎢𝜓̈ ⎥ ⎢
ir 0 1 ⎥ 𝜓̈
ir ⎢ 0 ⎥
⎢ ir ⎥ ⎢( ) ( )⎥ ⎢ ⎥
⎢ 𝜙̈ ⎥ ⎢ 𝜕f2 ∕𝜕y
ir 𝜕f2 ∕𝜕𝜓 ir ⎥ ⎢−D ⎥
⎢ ir ⎥ ⎢ ⎥ ⎢ s2⎥

⎣ 𝜃̈ ⎦ ⎣ 0 0 ⎦ ⎢ 0 ⎥
⎣ ⎦
This equation can be written as p̈ = Bdi p̈ i + 𝛄i , where velocity transformation matrix Bdi
and vector 𝛄i are given, respectively, by

⎡ 0 0 ⎤
⎢ ⎥
⎢ 1 0 ⎥
⎢( ) ( )⎥
⎢ 𝜕f1 ∕𝜕y 𝜕f1 ∕𝜕𝜓 ⎥
ir ir
Bdi = ⎢ ⎥ (6.126)
⎢ 0 1 ⎥
⎢( ) ( )⎥
⎢ 𝜕f2 ∕𝜕y 𝜕f2 ∕𝜕𝜓 ir ⎥
ir

⎢ ⎥
⎣ 0 0 ⎦

and
[ ]T
𝜸i = 0 0 −Ds1 0 −Ds2 0 (6.127)

Using velocity transformation matrix Bdi and vector 𝛄i , a number of equations of motion
equal to the number of the system degrees of freedom can be obtained and written, as pre-
viously explained, as
( )
Mii p̈ i = BTdi Qpe + Qpv − Mpp 𝛄i (6.128)

where Mii = BTdi Mpp Bdi , and the vector and matrices that appear in Eq. 128 are the same as
those of Eq. 107.
280 Mathematical Foundation of Railroad Vehicle Systems

6.13 HUNTING STABILITY


Hunting stability has been investigated in the literature using simplified two-degree-of-
freedom models whose motion is governed by linear differential equations. These simplified
models are obtained using the more general equations developed in the preceding section
by making several assumptions. In hunting stability investigations, a tangent track is often
used, and the assumption of small angles is made. By making these assumptions, one can
show that the nonlinear equations of Eq. 128 reduce to a set of linear differential equations
that can be used to explain important dynamic behaviors of the railroad vehicle system.
Therefore, the equations used in this section can be considered a special form of the more
general and nonlinear equations of Eq. 128.

Nonlinear Dynamic Equations Consider a wheelset i that consists of two wheels, as


shown in Figure 13. In the case of a tangent track, the vector that defines the global position
of origin Oi of the wheelset coordinate system X i Y i Z i can be written as
[ ]T [ ]T
Ri = Rix Riy Riz = si yir zir (6.129)
Because of this definition of vector Ri , and because of the assumption of a tangent track used
in the development presented in this section, it is expected that Eq. 107 will reduce to the
Newton–Euler equations of Eq. 44. Using vector Ri , the global position vector of a contact
point on wheel k = r, l, where r and l refer, respectively, to the right and left wheel, can be
written as rik = Ri + Ai uki ki
P , where uP is the position vector of contact point P with respect
to wheel reference point; and Ai is the transformation matrix that defines the wheelset
[ ]T
orientation in terms of the three Euler angles 𝛉i = 𝜓 ir 𝜙ir 𝜃 ir . Using the assumption
[ ] i
of tangent track, matrix Li of Eq. 101 reduces to Li = I 𝟎 , matrix H of Eq. 103 reduces
i
[ ] i ir
i
to H = 𝟎 G , Ati = I, Gti is a constant matrix, 𝛚ti is zero, and G = G that assumes the
i
form of matrix G defined in Eq. 21. Using these results, the vectors and matrices of Eq. 106
reduce to (Shabana et al. 2008)
[ ] [ ] [ ] [ ]
Li I 𝟎 𝛄iR 𝟎
i
B = i = , 𝛄 = i = ̇ i ir
i
(6.130)
H 𝟎 Gi 𝛄𝛼 G 𝛉̇
Therefore, one can write a matrix equation similar to Eq. 107 as
Mipp p̈ i = Qipe + Qipv (6.131)

w
Z
w
Y

w
X

Figure 6.13 Wheelset model.


Equations of Motion 281

where
[ ] ⎫
mi I 𝟎
,⎪
iT iT i i
Mipp = mi L Li + H I𝜃𝜃 H Mwp = iT i
𝟎 G I𝜃𝜃 G
i ⎪
[ ] ⎪
T Fie ⎪
T i i
Qipe = Li Fie + H Me = , ⎪

iT i
G Me
[ ( )] ⎬ (6.132)
T iT i
Qipv = −mi Li 𝛄iR − H I𝜃𝜃 𝛄i𝛼 + 𝛚i × I𝜃𝜃 𝛚i
i ⎪

⎡ ⎤ ⎪
𝟎 ⎪
= ⎢ [ ( )] ⎥ ⎪
⎢ −Gi Ii𝜃𝜃 Ġ 𝛉̇ i + 𝛚i × Ii𝜃𝜃 𝛚i
T i
⎥ ⎪
⎣ ⎦ ⎭

To obtain the two equations of motion that govern the hunting stability of the wheelset,
Eqs. 126 and 127 are used and rewritten as

⎡0 0 ⎤ ⎡ 0 ⎤
⎢ ⎥ ⎢ 0 ⎥
1 0⎥
[( ) ] ⎢ [( ) ] ⎢ ⎥
Bdi ⎢f ⎥ 𝛄i R ⎢−D ⎥
1y f1𝜓 ⎥
Bdi = ( )R = ⎢ 𝛄i = ( ) = ⎢ 1⎥
s
, (6.133)
Bdi 𝜃 ⎢0 1 ⎥ 𝛄i 𝜃 ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢f2y f2𝜓 ⎥ ⎢−Ds2 ⎥
⎢ ⎥ ⎢ ⎥
⎣0 0 ⎦ ⎣ 0 ⎦

where the first derivatives f 1y = 𝜕f 1 /𝜕yir , f 2y = 𝜕f 2 /𝜕𝜓 ir , f 1𝜓 = 𝜕f 1 /𝜕yir , and f 2𝜓 = 𝜕f 2 /𝜕𝜓 ir ;


2 2 2
the second derivatives f1yy = 𝜕 2 f1 ∕𝜕yir , f2yy = 𝜕 2 f2 ∕𝜕yir , f1𝜓𝜓 = 𝜕 2 f1 ∕𝜕𝜓 ir , f2𝜓𝜓 =
2
𝜕 2 f2 ∕𝜕𝜓 ir , Dsk = fkyy (ẏ ir )2 + fk𝜓𝜓 (𝜓̇ ir )2 + 2fky𝜓 ẏ ir 𝜓̇ ir , k = 1, 2; and

⎡ 0 0 ⎤ ⎡ 0 1 ⎤ ⎫
( ) ⎢ ⎥ ( ) ⎢ ⎥ ⎪
Bdi R = ⎢ 1 0 ⎥, Bdi 𝜃 = ⎢ f2y f2𝜓 ⎥ ,⎪
⎢ ⎥ ⎢ ⎥ ⎪
⎣ f1y f1𝜓 ⎦ ⎣ 0 0 ⎦ ⎪
⎬ (6.134)
⎡ 0 ⎤ ⎡ 0 ⎤ ⎪
( ) ⎢ ⎥ ( ) ⎢ ⎥ ⎪
𝛄i R = ⎢ 0 ⎥, 𝛄i 𝜃 = ⎢ −Ds2 ⎥ ⎪
⎢ ⎥ ⎢ ⎥ ⎪
⎣ −Ds1 ⎦ ⎣ 0 ⎦ ⎭

Using these definitions, the time derivatives of the Euler angles and the angular velocity
vector can be written as
( )
⎡𝜓̇ ir ⎤ ⎡ 𝜓̇ ir ⎤ ⎡−𝜓̇ ir cos 𝜙ir sin 𝜃 ir + ẏ ir f2y + 𝜓̇ ir f2𝜓 cos 𝜃 ir ⎤
⎢ ̇ ir ⎥ ⎢ ir ⎥ ⎢ ⎥
⎢ 𝜙 ⎥ = ⎢ẏ f2y + 𝜓̇ f2𝜓 ⎥ , 𝛚i = ⎢ 𝜔p + 𝜓̇ ir sin 𝜙ir
ir

⎢ ̇ ir ⎥ ⎢ ⎥ ⎢ ir (
ir cos 𝜃 ir + ẏ ir f + 𝜓̇ ir f
) ir ⎥
⎣𝜃 ⎦ ⎣ 𝜔p ⎦ ⎣ 𝜓
̇ cos 𝜙 2y 2𝜓 sin 𝜃 ⎦
(6.135)
282 Mathematical Foundation of Railroad Vehicle Systems

Assuming ixx = izz and the products of inertia of the wheelset are equal to zero, one has

⎡f2y cos 𝜃 ir 𝛼1 ⎤ [ ( )2 ]
i( ) ⎢ ⎥ ( )T iT i i ( ) iixx f2y iixx f2y f2𝜓
G Bdi 𝜃
=⎢ 0 sin 𝜙ir ⎥ , Bdi 𝜃 G I 𝜃𝜃 G Bdi 𝜃 =
⎢ ⎥ iixx f2y f2𝜓 ii𝜓𝜓
⎣ f2y sin 𝜃 𝛼2 ⎦
ir

(6.136)
In this equation, 𝛼 1 = f 2𝜓 cos 𝜃 ir − cos 𝜙ir sin 𝜃 ir , 𝛼 2 = f 2𝜓 sin 𝜃 ir + cos 𝜙ir cos 𝜃 ir , and ii𝜓𝜓 =
(( ) )
2
iiyy sin2 𝜙ir + iixx f2y + cos2 𝜙ir . Substituting Eq. 133 into Eq. 131, and premultiplying by
the transpose of transformation matrix Bdi , one obtains two scalar differential equations of
motion, which can be written in the form of Eq. 128 as
( )
Mii p̈ i = BTdi Qipe + Qipv − Mipp 𝛄i (6.137)
[ ir ] T
where pi = y 𝜓 ir and
( )
⎡mi 1 + (f )2 + ii (f )2 f f + ii f f ⎤ ⎫

T i
Mii = Bdi Mpp Bdi = ⎢ 1y xx 2y 1y 1𝜓 xx 2y 2𝜓
⎥ ⎪
⎢ ( )2 ⎥
i
⎣f1y f1𝜓 + ixx f2y f2𝜓 f1𝜓 + ii𝜓𝜓 ⎦ ⎪
[ ] ⎪
Feyi
+ Fez i
f1y + Qi𝜃2 f2y [ i ] ⎪
i T
T i
Bdi Qpe = i f i i
, Qpe = Fex Fey Fez Q𝜃1 Q𝜃2 Q𝜃3 ⎪
i i i i i
Fez 1𝜓 + Q𝜃1 + Q𝜃2 f2𝜓 ⎬
( ) ( ) T [ i ( i )] ⎪
T ̇ i
BTdi Qipv − Mipp 𝛄i = − Bdi 𝜃 G I𝜃𝜃 G 𝛉̇ i + 𝛚 × I𝜃𝜃 𝛚
i i i

[ ] ⎪
mi f1y Ds1 + iiyy f2y Ds2 ⎪
+ ⎪
i i
m f1𝜓 Ds1 + iyy f2𝜓 Ds2 ⎪

(6.138)
Equation 137, which is a nonlinear matrix equation, has two scalar equations that can be
solved using direct numerical integration methods. Vector Qipe in the preceding equation
includes the wheel/rail normal and tangential contact forces, which can be nonlinear func-
tions in creepages if nonlinear contact theory is used.

Creepage Definitions Vector Qipe of Eq. 138 includes the effect of creep forces, which
are expressed in terms of velocity creepages as explained in Chapter 5. For the simple
wheelset model considered in this section, closed-form expressions can be obtained for
creepages in terms of the prescribed velocities. Creepages at wheel/rail contact points can
be conveniently defined by introducing an intermediate wheelset coordinate system that
does not rotate with the wheelset about its Y i axis. This coordinate system is defined by
transformation matrix Aii , which is expressed in terms of two angles 𝜓 ir and 𝜙ir as
⎡cos 𝜓 ir − sin 𝜓 ir cos 𝜙ir sin 𝜓 ir sin 𝜙ir ⎤
⎢ ⎥
A = ⎢ sin 𝜓 ir cos 𝜓 ir cos 𝜙ir cos 𝜓 ir sin 𝜙ir ⎥
ii
(6.139)
⎢ ⎥
⎣ 0 sin 𝜙ir cos 𝜙ir ⎦
Using this transformation, the global position vector at contact point Pik is given by
P = R + A uP , k = 1, 2, where k = 1 for the right contact, k = 2 for the left contact,
rik i ii iik
Equations of Motion 283

y γ

Rl Rr

Figure 6.14 Right and left wheel rolling radius and conicity.

[ ]
k −r k T , dk is the distance of the contact point from the origin of the wheelset
uiik
P = 0 d
coordinate system in the lateral direction, and r k is the rolling radius at the contact point.
For the right and left wheels, one has, respectively, r 1 = Rr = r 0 − 𝛾yir and r 2 = Rl = r 0 + 𝛾yir ,
where r o and 𝛾 are, respectively, the nominal rolling radius and the conicity, as shown in
Figure 14. It follows that the absolute velocity of a contact point defined in the intermediate
wheelset coordinate system can be written as vPik = Aii Ṙ i + 𝛚ii × uiik
T
̇
ii ir
P , where 𝛚 = G 𝛉 ,
ii
ii i
where G can be defined using G by assuming 𝜃 = 0 since the intermediate wheelset
ir

coordinate system is not affected by the pitch rotation. This leads to


⎡ 0 1 0⎤
ii⎢ ⎥
G = ⎢ sin 𝜙 0 1⎥
ir (6.140)
⎢ ⎥
⎣cos 𝜙 0 0⎦
ir

[ ]T
Consequently, one can write 𝛚ii = 𝜙̇ ir 𝜃̇ ir + 𝜓̇ ir sin 𝜙ir 𝜓̇ ir cos 𝜙ir . Using small-angle
assumptions and neglecting higher-order terms, the angular velocity vector 𝛚ii reduces
[ ]T
to 𝛚ii = 𝜙̇ ir 𝜃̇ ir 𝜓̇ ir . Using this definition of the angular velocity and the small-angle
assumption in transformation matrix Aii , one can show that velocity vector vPik = Aii Ṙ i +
T

𝛚ii × uiik
P can be written as

⎡vik
px ⎤ ⎡ V ⎤ ⎡−dk 𝜓̇ ir − r k 𝜃̇ ir ⎤
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
py ⎥ = ⎢ẏ − V𝜓 ⎥ + ⎢
vPik = ⎢vik ir ir r k 𝜙̇ ir ⎥ (6.141)
⎢ ik ⎥ ⎢ ⎥ ⎢ ̇ ⎥
⎣vpz ⎦ ⎣ ż ir
⎦ ⎣ d 𝜙
k ir

As discussed in Chapter 5, longitudinal, lateral, and spin creepages are defined, respectively,
[ ]T √
as 𝜁xk = vik
px ∕V, 𝜁y = vpx ∕V, and 𝜁s = 𝛚 n ∕V, where n = 0 (−1)
k ik k ii • ik ik k+1
𝛾 1 ∕ 1 + (𝛾)2 ,
k = 1, 2, is a unit vector along the normal at the contact point. This unit normal vector is
defined in the intermediate wheelset coordinate system. Therefore, by using Eq. 141, the
following linear creepage expressions can be obtained:
284 Mathematical Foundation of Railroad Vehicle Systems

( ) ( ) ( ) ( )
⎡ d 𝛾 ⎤ ⎡ d 𝛾 ir ⎤
⎢ 𝜓̇ +
ir
yir ⎥ ⎢ − V 𝜓̇ − r y ⎥
ir
V r
⎡𝜁x ⎤ ⎢ ⎡𝜁x ⎤ ⎢
1 2
( )
o ⎥ ( )
o ⎥
⎢ 1 ⎥ ⎢ ẏ ir ro − 𝛾yir 𝜙̇ ir ⎥ ⎢ 2 ⎥ ⎢ ẏ ir ro + 𝛾yir 𝜙̇ ir ⎥
⎢𝜁y ⎥ = ⎢ + − 𝜓 ⎥,
ir ⎢𝜁y ⎥ = ⎢ + −𝜓 ⎥
ir (6.142)
⎢ 1⎥ ⎢ V ( ir
V
) ⎥ ⎢ 2⎥ ⎢ V ( ir
V
) ⎥
⎣𝜁s ⎦ ⎢ ⎥ ⎣𝜁s ⎦ ⎢ ⎥
𝜓̇ + 𝛾 𝜃̇ ir 𝜓̇ − 𝛾 𝜃̇ ir
⎢ ⎥ ⎢ ⎥
⎣ V ⎦ ⎣ V ⎦

In obtaining these linear creepage expressions, the higher-order term yir 𝜓̇ ir is neglected,
and it is assumed that d1 = − d − yir and d2 = d − yir , where d is half the gage. Furthermore,
in the case of conical wheels, contact angle 𝛿 c is the same as the conicity, and this leads to
the definition of the normal vector nik .

Linear Creep Force Model As discussed in Chapter 5, in Kalker’s linear creep theory,
longitudinal, lateral, and spin creep forces can be written, respectively, in terms of creepages
as F crx = −cxx 𝜁x , F cry = −cyy 𝜁y − cys 𝜁s , and M crs = csy 𝜁y − css 𝜁s , where cxx , cyy , cys = csy , and
css are creep coefficients that are assumed to be functions of the material properties and
dimensions of the contact area. These forces are determined in a contact frame, which can
be different from the wheelset and global coordinate systems. If lateral displacement yir
and yaw angle 𝜓 ir are assumed small, the creep coefficients can be assumed the same for
the right and left contacts. In this case, one has only the generalized creep forces associated
i = ∑2
( )
with coordinates yir and 𝜓 ir , defined, respectively, as Fey k=1 −cyy 𝜁y − cys 𝜁s and Q𝜃1 =
k k i
∑2 ( )
k=1 d(−1) cxx 𝜁x + csy 𝜁y − css 𝜁s . The first term in generalized force Q𝜃1 is a yaw moment
k k k k i

attributed to longitudinal creep forces at the two contacts. Using Eq. 142, one can write the
following expression for generalized creep forces:

2cyy ( ⎫
ro 𝛾 ) ir 2cys ir

i =−
Fey 1+ ẏ + 2c11 𝜓 ir − 𝜓̇
V d V ⎪
( )⎬ (6.143)
2cys ( ro 𝛾 ) ir 2c d 𝛾
Qi𝜃1 = 1+ ẏ − 2cys 𝜓 ir − ss 𝜓̇ ir − 2dcxx 𝜓̇ ir + yir ⎪
V d V V ro ⎪

Linearized Dynamic Equations A linear set of equations that sheds light on wheelset
dynamics and stability can be obtained from Eq. 137. Because yaw angle 𝜓 ir is small in most
railroad vehicle applications, zir and 𝜙ir can be assumed to depend only on yir . Using this
assumption, one has

⎡0 0⎤ ⎡ 0 ⎤
⎢ ⎥ ⎢ ⎥
⎢1 0⎥ ⎢ 0 ⎥
[( ) ] ⎢ ⎥ [( ) ] ⎢ ( ir )2 ⎥
Bdi R ⎢f1y 0⎥ 𝛄i R ⎢−f ẏ ⎥
𝛄i = ( ) = ⎢ ⎥
1yy
Bdi = ( ) = ⎢ ⎥, ⎢ ⎥
(6.144)
Bdi 𝜃 ⎢0 1⎥ 𝛄i 𝜃 0
⎢ ⎥ ⎢ ⎥
⎢f2y 0⎥ ⎢−f (ẏ ir )2 ⎥
⎢ ⎥ ⎢ 2yy ⎥
⎣0 0⎦ ⎢ ⎥
⎣ 0 ⎦
Equations of Motion 285

and the matrices and vectors in Eq. 138 reduce to


( ( )) ( )
⎡ mi 1 + f 2 + ii f 2 0 ⎤ ⎫
Mii = BTdi Mipp Bdi = ⎢
1y xx 2y
⎥ ⎪
⎢ i ⎥ ⎪
⎣ 0 i𝜓𝜓 ⎦ ⎪
[ i i i
] ⎪
Fey + Fez f1y + Q𝜃2 f2y ⎪
T i
Bdi Qpe = ⎪
i
Q𝜃1 ⎬ (6.145)
( ) ( ) [ ( )]⎪
BTdi Qipv − Mipp 𝛄i = − Bdi 𝜃 G i I i𝜃𝜃 Ġ i 𝛉̇ i + 𝛚i × I𝜃𝜃 𝛚i ⎪
T T i


[ ( ) ( )2 ] ⎪
mi f1y f1yy + iiyy f2y f2yy ẏ ir ⎪
+ ⎪
0 ⎭
Using the assumptions of small yaw and roll angles, a small change in the profile function,
and small lateral displacements and velocities, and recalling that iixx = iizz for a symmet-
ric wheelset, one obtains the linearized equations of motion of the two-degree-of-freedom
wheelset, which can be written as
[ ][ ] [ ]
mi 0 ÿ ir Feyi
+ Fezi
f1y + Qi𝜃2 f2y
= (6.146)
0 iizz 𝜓̈ ir 𝜃1 Qi − iiyy 𝜙̇ ir 𝜃̇ ir
If the assumption of pure rolling is made, the pitch angular velocity can be written as 𝜃̇ ir =
V∕ro , where V is the constant forward velocity of the wheelset and r o is the nominal rolling
radius. The applied forces and moments on the right-hand side of the preceding equations
include the creep forces and moments. Force Fey i is the lateral creep force, and Qi is the spin
𝜃1
creep moment. Substituting Eq. 143 into Eq. 146 and assuming zero applied forces except
creep forces and moments and gravity forces, one obtains
[ ][ ] [ ][ ] [ ][ ] [ ]
mi 0 ÿ ir dyy dy𝜓 ẏ ir kyy ky𝜓 yir 0
+ + = (6.147)
i
0 izz 𝜓̈ ir d𝜓y d𝜓𝜓 𝜓̇ ir k𝜓y k𝜓𝜓 𝜓 ir 0
where
( ( ))
dyy = 2cyy 1 + ro 𝛾∕d ∕V, dy𝜓 = 2cys ∕V, ⎫
( ( ) (i ) ) ( ) ⎪
dy𝜓 = −2cys ro 𝛾∕d + iyy 𝛾(V) ∕ro d ∕V, d𝜓𝜓 = 2 css + d cxx ∕V,⎬
2 2 (6.148)

kyy = kw1 , ky𝜓 = −2cyy , k𝜓y = 2dcxx 𝛾∕ro , k𝜓𝜓 = 2cys + kw2 ⎭
and kw1 and kw2 are stiffness coefficients resulting from accounting for gravity force. In
the case of a suspended wheelset like the one shown in Figure 15, suspension spring and
damper elements are used to connect the wheelset to the frame in the longitudinal and
lateral directions. The preceding equation can be modified to account for the effect of sus-
pension forces as
[ ][ ] [ ][ ] [ ][ ] [ ]
mi 0 ÿ ir dyy dy𝜓 ẏ ir kyy ky𝜓 yir 0
+ + = (6.149)
0 izzi 𝜓̈ ir
d𝜓y d𝜓𝜓 𝜓̇ ir
k𝜓y k𝜓𝜓 𝜓 ir 0
where
( ( ( )) ) ( ( ) ) }
dyy = 2cyy 1 + ro 𝛾∕d + 2dy V ∕V, d𝜓𝜓 = 2 css + d2 cxx + 2dx (b)2 V ∕V,
kyy = kw1 + 2ky , k𝜓𝜓 = 2cys + kw2 + 2kx b2
(6.150)
286 Mathematical Foundation of Railroad Vehicle Systems

dy, ky Y dy, ky

X
dx, kx dx, kx

Figure 6.15 Suspended wheelset.

where dx and dy are suspension damping coefficients, and kx and ky are suspension stiffness
coefficients, as shown in Figure 15. The effect of the suspension of the wheelset is examined
in the remainder of this section.

Effect of the Primary Suspension Unsuspended wheelsets are unstable if the effects
of spin creepage and gravity force are neglected. To investigate this instability, the primary
suspension coefficients in Eq. 149 are assumed to be zero. For such models, the hunting fre-
quency can be approximated using Klingel’s formula (Klingel 1883), which is derived using
kinematic equations√ without regard to the forces applied to the wheelset. This formula is
( )
given by f = (V∕2𝜋) 𝛾∕ ro d . Figure 16 shows the results of an eigenvalue analysis based
on the equations developed in this section, which account for the effect of creep forces. As
shown in Figure 16, this comparative analysis leads to a solution that is in good agreement
with the solution obtained using Klingel’s formula.
While unsuspended wheelsets are unstable, suspended wheelsets have critical speeds that
depend on suspension characteristics. For a given wheelset with specified suspension char-
acteristics, there is a critical speed below which the wheelset is stable and above which

9
8
7

6 Klingel’s formula
Frequency (Hz)

4 Linear analysis

1
0
0 25 50 75 100 125 150 175 200
Speed (m/s)

Figure 6.16 Eigenvalue analysis.


Equations of Motion 287

the wheelset is unstable. To examine the effect of the primary suspension on stability, the
wheelset shown in Figure 15 is assumed to be connected to the frame with a primary sus-
pension. This system can be unstable if the forward velocity V of the wheelset exceeds
the critical speed. To investigate the effects of the primary suspension and conicity on the
stability and critical speed of the wheelset, the data of Table 1 are used (Shabana et al. 2008).
The root loci results of Figure 17 show that as the wheelset forward velocity V increases,
the real part of the eigenvalues becomes positive, leading to unstable behavior. The effect
of the wheel profile conicity 𝛾 on the critical speed is demonstrated by the results shown in
Figure 18. In vibration theory, the ratio between the real part 𝛼 and the imaginary part 𝜔 of

Table 6.1 Data for the suspended wheelset model.

Parameter description Value

Wheelset mass mi 1568 kg


Inertia moment iixx 656 kg m2
Inertia moment iiyy 168 kg m2
Inertia moment iizz 656 kg m2
Longitudinal spring stiffness kx 1.35 × 105 N/m
Lateral spring stiffness ky 2.50 × 105 N/m
Longitudinal damping coefficient dx 0 N/(m s)
Lateral damping coefficient dy 0 N/(m s)
Distance between the longitudinal springs 2b 1.8 m
Half gage distance d 1.435 m
Nominal rolling radius r o 0.4566 m
Conicity 𝛾 1/40

Source: Courtesy of Shabana, A.A., Zaazaa, K.E., and Sugiyama, H.

40 V = 120 m/s
35 V = 70 m/s
30 First root
25
20
15
10 V = 40 m/s
5
0
Im

–5
–10
–15
–20 V
–25
Second root
–30
–35
–40
–2 0 2 4 6 8 10 12 14
Re

Figure 6.17 Root loci results.


288 Mathematical Foundation of Railroad Vehicle Systems

0.5

0.4 1/10
1/20
1/40
Degree of Stability

0.3

0.2

0.1
Conicity
0.0
–0.1
–0.2
0 20 40 60 80 100 120 140 160 180 200
Speed (m/s)
Figure 6.18 Conicity effect.

the eigenvalue defines the degree of stability. The critical speed is reached when the ratio
𝛼/𝜔 is equal to zero, which is the case for a critically stable system (sustained oscillations).
Because increasing the conicity causes the wheel/rail contact forces to add more energy to
the system, such an increase can lead to unstable behavior. This instability can be clearly
seen at low speeds by lowering the stiffness coefficients of the primary suspension.

6.14 MBS MODELING OF ELECTROMECHANICAL


SYSTEMS
Electromechanical systems are integral parts of modern vehicle systems, including railroad
systems. The equations that govern the electric circuits required to produce and control
forces such as levitation forces can be systematically integrated with the constrained
dynamic equations used in MBS railroad vehicle algorithms. In the preceding chapter,
maglev trains were discussed, and an expression for the magnetic levitation force F z was
obtained as
( )2
( ) 1 nco i (t)
Fz = Fz dz , i = (6.151)
Pp Ap RM
where dz = dz (t) is the distance between the pole faces; Ap = lp wp ; lp and wp are, respectively,
the length and width of the pole face; Pp is the permeability of the free space; nco is the
number of turns of the coil; i is the current; and RM = RM (dz ) is the reluctance of the mutual
flux. The force of Eq. 151 can be used to define magnetic levitation force vector Fml in the
global coordinate system. This force represents the magnetic force between the vehicle and
the guide. Using this force vector, the generalized forces associated with vehicle v and guide
g can be written, respectively, as
[ ] [ ]
Fml g Fml
v
Qml = T ( ) , Qml = − T ( ) (6.152)
Gv uv × Fml Gg ug × Fml
Equations of Motion 289

where uk , k = v, g, is the position of the point of application of the force with respect to
the origin of the respective body coordinate system. The formulations presented in this
chapter also allow for using distributed magnetic levitation forces, which can be used to
determine the generalized forces associated with the generalized coordinates of the vehicle
and the guide. In the case of distributed forces, an integration over the area of application
of the magnetic levitation force can be performed and used to determine the magnetic lev-
itation generalized forces, which can be added to the equation of motion defined by Eq. 55.
Current i and the distance between pole faces dz depend on time, and therefore, they are,
in general, functions of the system coordinates.
A more general approach that accounts for the time rate of change in the current
combines electric circuit equations with vehicle dynamic equations of motion. Using this
approach, power-supply voltage vs can be written in terms of current i, circuit resistance Rc ,
and circuit inductance Lc as vs = Rc i(t) + d(Lc (t)i(t))/dt, where, as defined in the preceding
chapter, Lc (t) = (nco )2 (RL + RM )/RL RM , RL = 1/PL is the reluctance of the leakage flux, and
PL is the permeance of the leakage flux. Power-supply voltage vs can be written as

vs = Rc i (t) + Lc (t) (di∕dt) + i (t) L̇ c (t) (6.153)

Inductance Lc , which depends on air gap dz , is defined as


( )2
( ) Pp nco Ap
Lc = Lc dz = LL + ( ) (6.154)
2 dz (t) + cp
where Pp is the permeability of the free space, cp is a constant defined in the preceding
chapter, and LL = (nco )2 PL = (nco )2 /RL . Substituting Eq. 154 into Eq. 153 and rearranging
the terms, one obtains
( ( )2 ) ( ( )2 )
Pp nco Ap di (t) Pp nco Ap ḋ z (t)
LL + ( ) = ( )2 − Rc i (t) + vs (6.155)
2 dz (t) + cp dt 2 d (t) + c
z p

This is a first-order ordinary differential equation in current i(t). It is a non-homogeneous


equation because of the presence of power-supply voltage vs on the right-hand side. There-
fore, the value of the magnetic levitation force, which depends on air gap dz , is determined
by the solution of the preceding equation, which depends on the value of voltage vs . As
discussed in Chapter 5, stability issues must be taken into consideration in the design of
maglev train systems. The magnetic levitation force decreases as air gap dz increases, and
this can lead to instability of the maglev suspension system. For this reason, a feedback con-
trol system is needed to achieve a stable electromagnetic suspension. The appropriate value
of power-supply voltage vs required to ensure stability can be determined using a feedback
control system. To this end, the power-supply voltage is allowed to vary as a function of
( ) ( ) ( )
air gap dz and its derivative ḋ z as vs dz , ḋ z = vso + f dz , ḋ z , where f dz , ḋ z is a control
function that can be evaluated if air gap dz and its derivative ḋ z are measured, and vso is the
( )
power voltage at the initial equilibrium configuration. Using the control function f dz , ḋ z ,
the value of the voltage can be adjusted to ensure that dz remains constant and ensure stable
behavior of the maglev suspension.
The electric current differential equation defined by Eq. 155 can be systematically
integrated with the MBS constrained dynamic equations developed in this chapter.
290 Mathematical Foundation of Railroad Vehicle Systems

To demonstrate the procedure used to achieve this integration, the case of a maglev
train system consisting of ns electromagnetic suspensions is considered. In this case,
all the electric currents of all the suspension circuits can be written in vector form as
[ ]T
i = i (t) = i1 (t) i2 (t) · · · ins (t) . Using Eq. 155 for each suspension circuit, one can
̇ where the function fms (i, q, q)
write di∕dt = fms (i, q, q), ̇ can be defined using Eq. 155,
and q is the vector of the system generalized coordinates. The electric current differential
equation di∕dt = fms (i, q, q) ̇ can be combined with the augmented form of the equations
of motion defined by Eq. 55 to obtain the MBS equation, which can be used to model the
maglev train system as
⎡ M CTq 𝟎⎤ ⎡ q̈ ⎤ ⎡ Q (i, q, q)̇ ⎤
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢Cq 𝟎 𝟎⎥ ⎢ 𝛌 ⎥ = ⎢ Qd (q, q) ̇ ⎥ (6.156)
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣𝟎 ̇ ⎦
𝟎 I ⎦ ⎣di∕dt⎦ ⎣fms (i, q, q)
where I is the ns × ns identity matrix, Q = Qe + Qv + Qml , and Qml is a vector that contains
all the generalized magnetic levitation forces. The preceding equation can be solved for q, ̈ 𝛌
and di/dt. Acceleration vector q̈ and the time derivative of the electric current di/dt can
be integrated to determine generalized coordinates q, generalized velocities q, ̇ and electric
currents i. It is important, however, to point out that, because of the lack of coupling in the
coefficient matrix of the preceding equation between the derivatives of the electric currents
and other unknown accelerations and Lagrange multipliers, it is more efficient to solve the
following two equations separately instead of solving Eq. 156:
[ ][ ] [ ]
M CTq q̈ Q (i, q, q)̇ di
= , ̇
= fms (i, q, q) (6.157)
Cq 𝟎 𝛌 Qd (q, q)̇ dt
Clearly, the second equation in this equation does not require any matrix inversion or LU
factorization, and therefore, there is no need to combine the two equations of Eq. 157 in the
matrix equation of Eq. 156.
291

Chapter 7

PANTOGRAPH/CATENARY SYSTEMS

Pantograph/catenary systems are used to supply power for electrically operated trains.
Most freight trains that travel for long distances are driven by diesel locomotives. However,
such diesel-powered locomotives have a speed limit of approximately 238 km/h (148 mph);
therefore, diesel-powered engines cannot be used for high-speed passenger trains that
operate at speeds of 300 km/h or higher. Extensive experimentation is currently being
conducted to significantly increase train speed to a level that minimizes travel time. There-
fore, pantograph/catenary technology is necessary: it represents the only viable option,
particularly as the demand for higher train speeds increases. Such technology is also more
environmentally friendly as compared to reliance on fuel, due to its higher efficiency and
lower emissions. Using pantograph/catenary systems is not limited to high-speed trains;
such systems are also used for urban transportation systems such as trams, trolleys, and
electrically powered buses that travel for short distances at lower speeds. Electrification
systems are less costly, quieter, and more reliable and can produce the greater amount of
power needed for higher speeds.
A pantograph, which is often mounted on the top of a train, tram, or bus, is used to collect
power through contact with an overhead power line called a catenary. The use of overhead
electric lines to provide power for transportation systems started in the late 1800s. Such
technology was first proposed for urban transportation, including trams and trolleys, and
later for buses; therefore, as previously mentioned, the use of this technology is not limited
to pantograph/catenary systems for high-speed trains (Figure 1).
It is also important to point out that pantograph/catenary systems, although they are
the most widely used method for providing power for high-speed trains, are not the only
method for supplying electric power to rail transportation systems. A third rail, batteries,
and ground-level power supplies are also used, particularly for urban rail transportation
systems.
Electrically powered trains use both overhead AC and DC voltage. AC voltage is normally
used with an overhead arrangement, while DC voltage is used with a third rail. The volt-
age used in some countries for electric train operations can reach 25 kV AC at 50–60 Hz.
AC electric supplies are used with induction motors. However, because of the DC-motor
advantage, an AC supply can be converted to DC voltage.
This chapter presents mathematical formulations that can be used for the virtual
prototyping and design of pantograph/catenary systems. The focus of the chapter is on
developing computational approaches that can be integrated with general multibody
Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,
First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
292 Mathematical Foundation of Railroad Vehicle Systems

Pantograph/catenary system

Figure 7.1 Pantograph/catenary system. Source: Vaccaro (2019).

system (MBS) algorithms for computer modeling and simulation of electrically operated
railroad vehicle systems.
A pantograph can be modeled as a mechanical system that consists of interconnected
rigid or flexible components whose motion equations can be developed using the MBS con-
strained dynamic formulations discussed in Chapter 6. Catenary wires, on the other hand,
are flexible cables that experience large displacements as the result of pantograph/catenary
contact and aerodynamic forces. These displacements can be described using the finite
element (FE) absolute nodal coordinate formulation (ANCF), which allows for accurately
capturing initial wire geometries. The structure of the pantograph/catenary system is first
discussed, and nonlinear kinematic and dynamic governing equations are developed. To
ensure the efficient and safe operation of pantograph/catenary systems, and to minimize
loss of contact, which can lead to arcing, several important issues must be taken into consid-
eration, including proper design of the system to maintain contact between the pantograph
and catenary wire, the effect of aerodynamic forces, wear control and reduction, and tem-
perature effects. Some of these topics are discussed in this chapter.

7.1 PANTOGRAPH/CATENARY DESIGN

As shown in Figure 1, a pantograph/catenary system consists of two main subsystems:


the pantograph, which is mounted on the top of the vehicle and is used to collect power,
and the electric cable, called a catenary, which transmits electric power through contact.
A pantograph can be modeled as a mechanical system consisting of interconnected com-
ponents, while the catenary, which consists of several cables, can be treated as a structural
system that experiences large displacements as the result of contact with the pantograph
Pantograph/Catenary Systems 293

and aerodynamic forces. The catenary and pantograph systems are described in this section
in more detail, and some of the important issues related to their design are highlighted to
provide a better understanding of the challenges encountered in the operation, design, per-
formance evaluation, and computer simulations of these systems.

Pantograph Two types of pantograph systems are in common use: single-arm and
double-arm pantograph systems. These two systems are shown in Figure 2. The single-arm
pantograph system was invented in 1955 by Louis Faiveley. It is Z-shaped and sometimes
referred to as a half-pantograph, and is more widely used because of its compactness,
lighter weight, and response and reliability at higher speeds; in addition, it requires
less power to control. Double-arm pantographs are also used because they are more
fault-tolerant as compared to single-arm pantographs.

(a) (b)

Figure 7.2 Pantograph types. Sources: (a) Vaccaro (2019); (b) hpgruesen/Pixabay.

Figure 3 shows the main components of a single-arm pantograph system. As shown,


a pantograph consists of interconnected components that include the thrust rod, lower
arm, balance rod, upper arm, crossbar, plunger, and pan-head. The pantograph system is
designed to maintain contact between the pan-head and the catenary contact wire. To main-
tain this contact and avoid separations – which can result in air gaps and arcing during the
operation of a railroad vehicle – pneumatic actuators are used, to provide sufficient uplift
forces. The uplift force should not be very high, to avoid component wear that can adversely
affect the performance of the pantograph/catenary system. Springs are also used as passive
controllers to improve the stability of the pantograph mechanism. Carbon strips are used on
the pan-head to collect electric current from the catenary wire, which is transmitted to the
pantograph system. The use of carbon material, as compared to copper, is preferred for the
pan-head strip due to the undesirable higher friction force that results from copper/copper
contact.

Pantograph Data In some computer simulation models developed for pantographs, the
pantograph lower arm is connected to the car body with a revolute joint; the lower link is
connected to the car body with a spherical joint and to the upper arm with a spherical joint;
the lower arm is connected to the upper arm with a revolute joint; the plunger is connected
to the upper arm with a revolute joint and the upper link with a spherical joint; and the
294 Mathematical Foundation of Railroad Vehicle Systems

Plunger

Pan-head
Upper link

Upper arm
Balance rod

Lower arm

Lower link

Figure 7.3 Single-arm pantograph.

lower arm is connected to the upper link with a spherical joint (Pappalardo et al. 2016;
Daocharoenporn et al. 2019). In the model developed by Pappalardo et al. and Daocharoen-
porn et al., the lower arm mass is assumed to be 32.18 kg, and its mass moments of inertia are
assumed to be 0.31, 10.43, and 10.65 kg⋅m2 . The upper arm mass is assumed to be 15.60 kg,
and its mass moments of inertia are assumed to be 0.15, 7.76, and 7.86 kg⋅m2 . The lower
link mass is assumed to be 3.10 kg, and its mass moments of inertia are assumed to be 0.05,
0.46, and 0.46 kg⋅m2 . The upper link mass is assumed to be 1.15 kg, and its mass moments
of inertia are assumed to be 0.05, 0.48, and 0.48 kg⋅m2 . The plunger mass is assumed to be
1.51 kg, and its mass moments of inertia are assumed to be 0.07, 0.05, and 0.076 kg⋅m2 . The
pan-head mass is assumed to be 9.50 kg, and its mass moments of inertia are assumed to be
1.59, 0.21, and 1.78 kg⋅m2 . When multiple pantographs are used to power high-speed trains,
the minimum distance between two pantographs should not be less than 200 m. These data
depend on the pantograph model used for specific train models and therefore should not
be used with all simulation models. They are presented here for the purpose of providing
examples of inertia data that can be used in computer simulation models.

Catenary Figures 4 and 5 show a catenary system used to provide electric power for the
operation of railroad vehicles. Different names are used to refer to catenary systems, includ-
ing overhead line and overhead wire. In addition to their use for electrically operated trains,
catenary systems are used to supply the power required for the operation of urban vehi-
cles such as trams, trolleys, and electric buses. In the case of intercity trains, catenary wires
receive electricity from feeder stations located at different distances along the track. The
feeder stations receive electricity from high-voltage electric grids. Electric current is col-
lected by the train using the pantograph, which has a contact strip attached to the pan-head.
The pan-head carbon strip maintains contact with the catenary by applying uplift force.
The combined pantograph/catenary system is also referred to as a current collection sys-
tem. Current collection systems allow electric current to flow through to the train and back
Pantograph/Catenary Systems 295

Figure 7.4 Catenary system. Source: Jackin/Adobe Stock.

to the feeder station through wheel/rail contact. This arrangement for the flow of current
represents a closed electric circuit.

Catenary Structure The most common catenary system consists of contact and mes-
senger wires connected by droppers that allow only tensile forces. The catenary system also
includes supports, moving brackets, and steady arms (pull-off fittings). The contact wire
provides electric power to the pantograph through contact with the pan-head, as previously
mentioned. Because of the friction resulting from pantograph/catenary contact and the fre-
quent trips normally made by trains, fatigue and wear of the catenary wires and pan-head of
the pantograph are important design considerations. In addition to contact forces, the cate-
nary can be subjected to significant cross-wind forces that result in severe oscillations. As a

Vertical droppers

Figure 7.5 Catenary system. Source: peuceta/Adobe Stock.


296 Mathematical Foundation of Railroad Vehicle Systems

result of these forces, undesirable changes are possible in the geometry of the contact wire,
which can negatively impact current-collection quality; these changes must be considered
at the design stage.
To achieve high-quality, high-speed current collection, it is necessary to avoid significant
deviations in the contact wire geometry, which can result from severe vibrations or large
deformations. One of the methods used to avoid such changes in the contact wire geom-
etry is to support the contact wire using a messenger wire, which prevents slanting and
severe oscillations of the contact wire resulting from pantograph/catenary interaction forces
and other forces such as aerodynamic forces. Nonetheless, both messenger and contact
wires may experience nonlinear elastic deformations; therefore, it is necessary to examine
their dynamic response using computer simulations. In the pantograph/catenary computer
model developed by Seo et al. (2005), the mass density, cross-sectional area, and Young’s
modulus of the contact wire were assumed to be 3200 kg/m3 , 16.0 mm2 , and 9 × 109 N/m2 ,
respectively. Additional catenary data used in computer simulations were provided by Pap-
palardo et al. (2016) and Daocharoenporn et al. (2019).
The contact wire is connected to the messenger wire by vertical droppers, shown in
Figure 5. The droppers provide additional stiffness that helps maintain system stability. The
vertical droppers are designed to have high tensile stiffness but zero resistance to compres-
sive forces. The steady horizontal arm is used to maintain the shape of the catenary such
that pan-head wear is reduced. In developing computer models, the connection between
the steady arm and the support can be modeled using a pin joint to allow for contact wire
oscillations. As pointed out by Seo et al. (2005), the tension in the droppers between the
contact wire and the messenger wire, span length of the support, gap, number of droppers,
etc. have an effect on the vertical stiffness of catenary systems. In the computer simulations
performed by Seo et al. (2005), the distance between two droppers is assumed to be 12.5 m.
While the contact wire is modeled using ANCF finite elements, discrete springs and
dampers are used to model the droppers. The stiffness and damping coefficients of the
droppers are assumed to be 2.0 × 105 N/m and 100 N⋅s/m, respectively.

Catenary Geometry on Tangent Tracks Because the catenary and messenger wires
are approximately straight lines between the supports, when a vehicle negotiates a tan-
gent track, contact between the pantograph carbon strip used for current collection and the
catenary contact wire can be localized; this results in a small contact area of high stress and
nonuniform wear that can cause damage to the carbon strip. For this reason, the catenary is
zigzagged over tangent track sections to achieve more uniform wear by allowing the contact
wire to slide over the pantograph carbon strip. This zigzagging geometry is not required in
curved track sections since as the vehicle negotiates a curve, the contact wire sweeps over
the pantograph carbon strip. This issue is revisited when the pantograph/catenary wear
problem is discussed in a later section of this chapter.

Catenary Pretension Pretension contributes to greater bending stiffness as a result of


the coupling between axial and bending deformations. This additional geometric stiffen-
ing is necessary to maintain catenary stability and ensure stable contact with the pan-head.
As the stiffness of the contact wire increases, the speed of the propagation of elastic waves
Pantograph/Catenary Systems 297

resulting from pan-head/catenary interaction also increases to a level that is much higher
than the vehicle operating speed. In addition to achieving stable pan-head/catenary con-
tact, the speed of the propagation of the elastic waves is an important factor that must be
considered. To avoid resonance, the frequency of the wire oscillations and the speed of the
elastic waves must be much higher than the train operating speed, and this can be achieved
using pretension applied to the catenary wires.
To make the pretension independent of weather conditions, including temperature varia-
tions, a constant pretension of 9000–20 000 N per wire is normally produced using balance
weights or hydraulic tensioners. The weights are used to balance axial cable strains that
cause pretension and increase the wire bending stiffness. When weights are used to produce
wire pretension, the positions of the weights can change as the temperature changes. To
prevent the weights from swaying, their motion is restricted by using sliding bars or tubes.
Weight movements also impose a limit on the maximum length of the catenary wires. Due
to temperature variations, the wires can expand or contract, and this deformation is pro-
portional to the wire length and inversely proportional to the axial stiffness of the wire. For
this reason, the maximum length of the wires between two pretension weight locations is
limited to slightly less than 2 km. To prevent the catenary wires from sliding along the track
as the result of the balance weights moving when the temperature varies, anchors fixed to
the catenary poles or supports are placed close to the middle of the distance between the
weight locations to constrain the motion of the catenary wire in the longitudinal direction
of the track. Some other designs are also used that include stoppers to prevent the wires
from sliding in the case of unexpected failure.
As previously mentioned, pretensioning the contact wire increases its stiffness and the
speed of elastic wave propagation in the wire. The speed of elastic wave propagation depends
not only on the contact wire stiffness but also on its inertia and material properties. It is
recommended that the maximum train speed should not exceed 70% of the speed of elas-
tic wave propagation in the contact wire. That is, for a train traveling at 300 km/h, the
speed of wave propagation in the contact wire should be greater than 430 km/h. This is
necessary to ensure safe operation of the train and avoid the loss of pantograph/catenary
contact.

Voltage Specifications The voltage used for electrically operated trains depends on the
country and region; different countries use different voltages for different rail transporta-
tion systems. Both AC and DC voltage are used. AC voltage is normally used with conductor
motors. Although DC voltage systems are more commonly used with a third rail, due to the
DC-motor advantage, an AC electric supply can be converted to a DC system. Some of the
most commonly used voltages, shown in Table 1, were selected for international standard-
ization (CENELEC 2007; IEC 2007). The voltage range allowed for standardized voltages
depends on the number of trains collecting current and the distance of these trains from
feed stations.
For example, the planned traction power supply system for a future California
high-speed train will utilize a 25 kV and 60 Hz catenary and a negative (−25 kV) longitudi-
nal feeder together with autotransformers placed approximately every 8 km (5 miles). The
2 × 25 kV – 60 Hz arrangement is used to reduce the number of feed stations (Hsiao 2010).
298 Mathematical Foundation of Railroad Vehicle Systems

Table 7.1 Voltage standardization

Voltage
Min. Min. Max. Max.
Electrification system non-permanent permanent Nominal permanent non-permanent

600 V DC 400 V 400 V 600 V 720 V 800 V


750 V DC 500 V 500 V 750 V 900 V 1000 V
1500 V DC 1000 V 1000 V 1500 V 1800 V 1950 V
3 kV DC 2 kV 2 kV 3 kV 3.6 kV 3.9 kV
15 KV AC, 16.7 Hz 11 kV 12 kV 15 kV 17.25 kV 18 kV
25 KV AC, 50 Hz (EN 50163) 17.5 kV 19 kV 25 kV 27.5 kV 29 kV
and 60 Hz (IEC 60850)

Source: CENELEC (2007); IEC (2007).

7.2 ANCF CATENARY KINEMATIC EQUATIONS


A large number of investigations have been devoted to studying catenary system dynamics
and vibration. Some of these investigations are based on linear models, while others propose
nonlinear models with varying degrees of complexity and assumptions. In some cases, the
catenary is modeled using simple discrete spring-damper elements. Such simplified models
do not account for the effect of distributed inertia and elasticity of catenary cables and do
not accurately predict the cable stresses that are necessary for credible evaluation of cable
integrity and durability. Such credible assessments can be made using continuum-based
models that account for the distributed inertia and elasticity of catenary wires. Such
continuum-based models allow for a more accurate evaluation of the speed of elastic
wave propagation and the implementation of different material models; these models also
account for geometric nonlinearities that result from large-amplitude oscillations.
Catenary design has a direct effect on the operation of electrically powered trains, as evi-
denced by the fact that pantograph/catenary contact is one of the main factors contributing
to limiting train speeds. As a result of the forces acting on the catenary system, elastic waves
are generated in the catenary cable. The train operating speed cannot exceed the speed of
the propagation of catenary elastic waves due to safety considerations (Kumaniecka and
Jacek 2008; Pappalardo et al. 2016). Therefore, it is recommended to use a continuum-based
approach for the catenary model that can be systematically integrated with general MBS
computational algorithms.

Catenary Models A catenary system consists of flexible wires that can be subjected
to significant vibrations, including aerodynamics and contact forces. These vibrations
can result in a loss of contact with the pan-head, leading to poor current collection that
can adversely affect the performance of the pantograph/catenary system. As a result of
catenary/pan-head contact, the contact wire can experience large nonlinear vibrations,
which must be accounted for to develop realistic virtual prototyping models. Capturing
the nonlinearity of catenary vibrations contributes to accurate prediction of the location
of the contact point between the catenary wire and pan-head in computer simulations. In
Pantograph/Catenary Systems 299

the literature, as previously mentioned, simplified models have been proposed; in some
of these models, lumped parameters are used; other models use discrete springs that have
no inertia. Other models include an analytical description of the catenary kinematics,
including the Fourier sine expansion method. Some other approaches employ the finite
element (FE) method using conventional beam elements that do not correctly capture
large displacements as a result of the type of nodal coordinates used. In some studies of
pantograph/catenary interaction, co-simulation techniques are used by using two different
computer programs: one for modeling the pantograph and vehicle system, and an FE
computer program for modeling the catenary system. When co-simulation techniques
are used, communication between different software using different formulations and
numerical procedures can be an issue.
In other investigations, continuum-based ANCF finite elements are used for modeling
catenary flexibility (Seo et al. 2006; Lee and Park 2012). ANCF finite elements can be
directly implemented in MBS computer programs, thereby eliminating the need to use
co-simulations. The FE/ANCF approach is general and allows for developing a detailed
catenary model that has a large number of degrees of freedom. This approach also allows
for using nonlinear material models and conveniently describing catenary geometries. In
this chapter, catenary equations of motion are developed using ANCF finite elements that
are suited for nonlinear large displacement analysis. These elements lead to a constant
mass matrix regardless of the magnitude of the catenary-wire displacements.

ANCF Finite Elements ANCF finite elements, introduced in Chapter 2 for the
geometry description, also allow for developing accurate, general, flexible body models
for predicting catenary vibrations and stresses. These continuum-based elements can
capture both the axial and bending deformation of catenary cables, and also account for
coupling between these two deformation modes. Two different three-dimensional ANCF
finite elements are good candidates for developing catenary dynamic equations of motion:
the ANCF cable element, which has a smaller number of degrees of freedom and does not
capture the effect of shear deformation; and the fully parameterized ANCF beam element
introduced in Chapter 2. The fully parameterized ANCF beam element captures the
effect of shear deformation and the change in cross-section dimensions that results from
cross-section stretch, axial, and bending displacements. That is, the effects of pretension
and temperature on the change in cross-section dimensions can be examined using the
ANCF fully parameterized beam element. Nonetheless, both the three-dimensional ANCF
cable and fully parameterized beam elements capture the in-plane and out-of-plane
bending deformations of the catenary. Furthermore, as a result of using ANCF position
vector gradients as nodal coordinates, complex geometries can be modeled; these features
allow for experimenting numerically with different catenary wire shapes. As previously
mentioned in Chapter 2, the fully parameterized ANCF beam element was mainly used
for the geometric description of the rail. In this chapter, ANCF finite elements are used to
develop both kinematic and dynamic equations of motion of catenary wires.

ANCF Cable Element The three-dimensional ANCF cable element can be used to
capture the extension and in- and out-of-plane bending deformations (Gerstmayr and
Shabana 2006; Shabana 2018). However, this element does not allow for deformation of the
300 Mathematical Foundation of Railroad Vehicle Systems

cross-section since the element shape function matrix depends only one parameter. The
ANCF cable element, shown in Figure 6, is assumed to have two nodal coordinates, and
each node has six coordinates: three translation coordinates and three gradient coordinates
that define the elements of the tangent vector at the node. Therefore, the vector of nodal
[ ( ) ]T
T
jk
coordinates at node k of an element j can be written as e = rjk rx
jk , k = 1, 2, where
jk
rjk is the global position vector of node k, rx = 𝜕rj ∕𝜕x is the position-gradient vector
evaluated at node k, and x is the element spatial coordinate in the longitudinal direction.
Because the ANCF cable element has 6 coordinates at each node, the total number of
element nodal coordinates is 12. Therefore, polynomials with 12 coefficients can be used
as a starting point for developing the element displacement field. For the cable element,
the following cubic polynomials can be used
j
⎡r1 ⎤ ⎡a0 + a1 x + a2 (x)2 + a3 (x)3 ⎤
⎢ ⎥ ⎢ ⎥
rj = rj (x) = ⎢r2j ⎥ = ⎢ b0 + b1 x + b2 (x)2 + b3 (x)3 ⎥ (7.1)
⎢ ⎥ ⎢ ⎥
⎢ j⎥ ⎢ 2 3 ⎥
⎣r3 ⎦ ⎣ 0 c + c1 x + c2 (x) + c3 (x) ⎦
where al , bl , and cl , l = 0,1,2,3, are the time-dependent polynomial coefficients that can
be replaced by the element position and position-gradient nodal coordinates defined
[( ) ( ) ]T [ ]T
T T T j2 T
= rj1 rj1
T T
by vector ej = ej1 ej2 x rj2 rx , which can also be written as
[ T T
] T
ej = rj (x = 0) rjx (x = 0) rj (x = l) rjx (x = l) . Using these definitions of the nodal
T T

coordinates, the cable element displacement field can be written as

rj = rj (x, t) = Sj (x) ej (t) (7.2)

In this equation, Sj (x) is the element shape function matrix, which can be written as
[ ]
Sj = s1 I s2 I s3 I s4 I (7.3)

r xj2

r xj1

r j1

Z
Y
r j2

Figure 7.6 ANCF cable element.


Pantograph/Catenary Systems 301

where I is the 3 × 3 identity matrix,


( ) ⎫
s1 = 1 − 3𝜉 2 + 2𝜉 3 , s2 = l 𝜉 − 2𝜉 2 + 𝜉 3 ,⎪
( ) ⎬ (7.4)
s3 = 3𝜉 2 − 2𝜉 3 , s4 = l −𝜉 2 + 𝜉 3 ⎪

are the shape functions, 𝜉 = x/l, and l is the length of the finite element.
The cable element described in this section is based on cubic interpolation; therefore, it
is suited for modeling bending problems. Because it is a three-dimensional element, it cap-
tures both in- and out-of-plane bending deformations. The cubic interpolation allows for
describing non-constant curvatures, a necessary feature for modeling bending problems.
This description is also consistent with the basic bending-vibration equation of beams,
which is a fourth-order partial differential equation. Using an interpolation lower than
cubic imposes restrictions on the definition of the curvature within the element and can
lead to undesirable discontinuity problems.
Furthermore, using position vector gradients as nodal coordinates allows for modeling
initially curved cables and large amplitudes of vibration. The cable element displacement
field, however, depends on only one spatial coordinate, the longitudinal coordinate x.
Therefore, the deformation of the element cross-section cannot be accounted for when this
element is used. In addition, the effect of shear deformation is not taken into consideration
in the formulation of the displacement field of this element. Nonetheless, because this
element has significantly fewer coordinates as compared to the fully parameterized ANCF
beam element, it has proven to be very efficient for modeling catenary system vibrations
and also does not suffer from the locking problems encountered when fully parameterized
ANCF elements are used.
The ANCF displacement field can be used to describe arbitrarily large rigid-body displace-
ments because it includes linear terms as part of the polynomial representation. Therefore,
such an element displacement field leads to zero strain under arbitrary rigid-body displace-
ments. Using this description eliminates the need to use an incremental solution procedure
and allows for direct and efficient implementation in MBS computational algorithms. Fur-
thermore, the ANCF displacement-field description allows for defining the absolute veloc-
ity and acceleration vectors as linear functions of the nodal velocities and accelerations. By
differentiating the position equation rj (x, t) = Sj (x)ej (t) once and twice with respect to time,
one obtains the absolute velocity and acceleration vectors, which can be written, respec-
tively, as ṙ j (x, t) = Sj (x) ė j (t) and r̈ j (x, t) = Sj (x) ë j (t).

Fully Parameterized ANCF Beam Element The three-dimensional, two-node, fully


parameterized ANCF beam element is shown in Figure 7. Each node has 12 coordinates
that include the 3 components of the node global position vector and 9 components of the
matrix of position vector gradients at the node. Therefore, the element has 24 degrees of
freedom: 12 at each of the element nodes (Yakoub and Shabana 2001). This element was
introduced in Chapter 2 and is used to define the rail geometry regardless of whether the rail
is modeled as a flexible body. If x, y, and z are used as the element spatial coordinates, the
geometry of the rail surface can be defined using the functional relationship z = z(x, y). This
relationship can be used for the geometry description, and it is not invoked in the analysis
302 Mathematical Foundation of Railroad Vehicle Systems

r2z
rx2

ry2

r2

Z
Y
O
r1 r1z
X
rx1

ry1

Figure 7.7 ANCF beam element.

of deformations. That is, during the deformation analysis step, the element spatial coordi-
nates x, y, and z are assumed independent, while in the search for the contact point on the
rail surface, for example, the functional relationship z = z(x, y) is used. It is also important
to mention that in a computer implementation, the solution algorithm can be designed to
allow using different elements for the geometry and deformation analyses. For example,
the ANCF fully parameterized element can be used to define the surface geometry, while
the ANCF cable element can be used in the deformation analysis to achieve efficient com-
puter simulations. In the case of wheel/rail contact, using the ANCF fully parameterized
element is necessary to capture the surface geometry of the rail accurately. In the case of a
pantograph/catenary system, the geometry is simpler, and using the ANCF fully parame-
terized element is not necessary unless the effects of deformation modes such as changing
cross-section dimensions need to be evaluated.
Because the fully parameterized ANCF beam element has 24 nodal coordinates, the fol-
lowing cubic interpolation can be used for beam element j

⎡r1j ⎤ ⎡a0 + a1 x + a2 y + a3 z + a4 xy + a5 xz + a6 (x)2 + a7 (y)3 ⎤


⎢ ⎥ ⎢ ⎥
rj = ⎢r2j ⎥ = ⎢ b0 + b1 x + b2 y + b3 z + b4 xy + b5 xz + b6 (x)2 + b7 (y)3 ⎥ (7.5)
⎢ ⎥ ⎢ ⎥
⎢r j ⎥ ⎢ c + c x + c y + c z + c xy + c xz + c (x)2 + c (y)3 ⎥
⎣ 3⎦ ⎣ 0 1 2 3 4 5 6 7 ⎦

where al , bl , and cl , l = 0, 1, …, 7, are the time-dependent polynomial coefficients that can be


replaced by the element position and position-gradient nodal coordinates defined by vector

[( ) ( ) ]T [ ]T
T T T j1 T j1 T j2 T j2 T j2 T
ej = = rj1 rj1
T T
ej1 ej2 x ry rz rj2 rx ry rz (7.6)
Pantograph/Catenary Systems 303

In this equation,
⎡e1 ⎤ ⎡e4 ⎤
⎢ ⎥ j1 j ⎢ ⎥
r = r (0, 0, 0) = ⎢e2 ⎥ ,
j1 j
rx = rx (0, 0, 0) = ⎢e5 ⎥ ,
⎢ ⎥ ⎢ ⎥
⎣e3 ⎦ ⎣e6 ⎦
⎡e7 ⎤ ⎡e10 ⎤
j1 j ⎢ ⎥ ⎢ ⎥
ry = ry (0, 0, 0) = ⎢e8 ⎥ , rj1
z = rjz (0, 0, 0) = ⎢e11 ⎥ (7.7)
⎢ ⎥ ⎢ ⎥
⎣e9 ⎦ ⎣e12 ⎦
are the coordinates of the first node of the element,
⎡e13 ⎤ ⎡e16 ⎤
⎢ ⎥ j2 j ⎢ ⎥
r = r (l, 0, 0) = ⎢e14 ⎥ ,
j2 j
rx = rx (l, 0, 0) = ⎢e17 ⎥ ,
⎢ ⎥ ⎢ ⎥
⎣e15 ⎦ ⎣e18 ⎦
⎡e19 ⎤ ⎡e22 ⎤
j2 j ⎢ ⎥ ⎢ ⎥
ry = ry (l, 0, 0) = ⎢e20 ⎥ , rj3
z = rjz (l, 0, 0) = ⎢e23 ⎥ (7.8)
⎢ ⎥ ⎢ ⎥
⎣e21 ⎦ ⎣e24 ⎦
j
are the coordinates of the second node of the element, l is the element length, and r𝛼 =
𝜕rj ∕𝜕𝛼, 𝛼 = x, y, z. Using the conditions of Eqs. 7 and 8, the coefficients of the polynomials
in Eq. 5 can be replaced by the vector of nodal coordinates of Eq. 6, which can be written
[ ]T
using Eqs. 7 and 8 as ej = e1 e2 … e24 . Therefore, one can show that the displacement
field of the fully parameterized ANCF beam element can be written as
rj = rj (x, t) = Sj (x) ej (t) (7.9)
[ ]T j
In this equation, x = x y z is the vector of spatial coordinates and S is the element
shape function matrix, which can be written as
[ ]
Sj = s1 I s2 I s3 I s4 I s5 I s6 I s7 I s8 I (7.10)
where the shape functions si , i = 1, 2, …, 8 are defined as (Yakoub and Shabana 2001)
( ) ⎫
s1 = 1 − 3𝜉 2 + 2𝜉 3 , s2 = l 𝜉 − 2𝜉 2 + 𝜉 3 , s3 = l (𝜂 − 𝜉𝜂) , ⎪
( ) ⎪
s4 = l (𝜍 − 𝜉𝜍) , s5 = 3𝜉 2 − 2𝜉 3 , s6 = l −𝜉 2 + 𝜉 3 , ⎬ (7.11)

s7 = l𝜉𝜂, s8 = l𝜉𝜍 ⎪

and 𝜉 = x/l, 𝜂 = y/l, and 𝜍 = z/l.
Unlike the cable element, which has a displacement field written in terms of only
one parameter x, the displacement field of the fully parameterized beam element is
expressed in terms of three parameters x, y, and z. Therefore, the fully parameterized
beam element is capable of capturing deformation modes that cannot be captured by the
cable element. These deformations, as previously mentioned, include shear deformation
and changing cross-section dimensions. This generality comes at the expense of higher
computational cost as a result of the increase in the number of coordinates and the exis-
tence of high-frequency modes that couple different displacements. Therefore, the fully
304 Mathematical Foundation of Railroad Vehicle Systems

parameterized beam element is recommended for use in catenary problems in which the
effects of these deformation modes are important and need to be investigated.

7.3 CATENARY INERTIA AND ELASTIC FORCES


The ANCF element displacement fields defined in the preceding section for the cable and
fully parameterized elements can be used to develop the expressions of the element inertia
and elastic forces. In the developments presented in this section, the catenary is assumed
to be discretized using ne finite elements. For both the gradient-deficient cable and fully
parameterized beam elements, the displacement field for element j can be written as
rj = rj (x, t) = Sj (x)ej (t), which in the case of the cable element x reduces to a scalar – that is,
x = x – and in the case of the fully parameterized beam element, x is a three-dimensional
[ ]T
vector defined as x = x y z . As a result of using the full parameterization for the beam
element, a complete set of gradient vectors can be determined. This allows for using a more
general approach for formulating the elastic forces at the expense of having high-frequency
deformation modes and encountering locking problems in some applications.

Catenary Inertia Forces Using the displacement field rj = rj (x, t) = Sj (x)ej (t), the abso-
lute velocity and acceleration vectors of an arbitrary point on finite element j can be written,
respectively, as vj = ṙ j = Sj ė j and aj = r̈ j = Sj ë j , j = 1, 2, … , ne . The virtual displacement of
the arbitrary point can be written as 𝛿rj = Sj 𝛿ej . Therefore, the virtual work of the inertia
forces of ANCF element j can then be defined as
( T T) ( )
j T
𝛿Wi = 𝜌j aj 𝛿rj dV j = 𝜌j ë j Sj Sj 𝛿ej dV j
∫ j ∫V j
{V }
T T
= ë j j j j j
𝜌 S S dV 𝛿ej , j = 1, 2, … , ne (7.12)
∫V j
where 𝜌j and V j are, respectively, the mass density and volume of the ANCF element. The
j [ T ]
preceding equation can be written as 𝛿Wi = ë j Mj 𝛿ej , where M j is the constant and sym-
metric mass matrix of ANCF element j defined as
T
Mj = 𝜌j Sj Sj dV j , j = 1, 2, … , ne (7.13)
∫V j
Because the element mass matrix is constant, the quadratic velocity centrifugal and Corio-
lis inertia forces are equal to zero in this formulation. Given the dimensions and material
properties of the catenary wires, the mass matrix in the preceding equation can be evaluated
only once in advance of the dynamic simulation. In the case of the ANCF cable element,
integration over the volume is not required, and the mass matrix can be evaluated by inte-
grating over the element length. In this case, the cable-element mass matrix can be simply
l
evaluated as Mj = ∫0 𝜌j Aj Sj Sj dx, where Aj is the element cross-section area.
T

j jT j
The virtual work of the inertia forces can then be written as 𝛿Wi = Qi 𝛿ej , where Qi is
the vector of the generalized inertia forces associated with the element nodal coordinates.
This vector can be written as
j
Qi = Mj ë j , j = 1, 2, … , ne (7.14)
Pantograph/Catenary Systems 305

In obtaining this simple expression of the inertia forces, no assumptions are made except for
the order of the interpolating polynomials. Therefore, this simple expression of the inertia
forces can be used in the case of very large displacements and does not impose any restric-
tions on the amount of deformation or rotation within the element except for the restrictions
imposed by the order of the interpolating polynomials. This ANCF feature is attributed to
using position vector gradients as nodal coordinates. This choice of coordinates allows for
describing arbitrary displacements, including general rigid-body displacements. The vir-
tual work and vector of generalized inertia forces are used in the development of catenary
equations of motion presented in this section.

Catenary Elastic Forces In the case of the fully parameterized ANCF beam element,
one can evaluate all the position gradient vectors. In this case, a general continuum mechan-
ics approach can be used in the formulation of the elastic forces. The general continuum
mechanics approach for formulating stress forces is documented in standard texts on con-
tinuum mechanics and the theory of elasticity. Using this approach, the virtual work of the
stresses of element j can be written in terms of the Green-Lagrange strain tensor 𝜺j and the
j
second Piola-Kirchhoff stress tensor 𝛔P2 as (Bonet and Wood 1997; Boresi and Chong 2000;
Ogden 1984; Shabana 2018)

j j
𝛿Ws = − 𝛔P2 ∶ 𝛿𝛆j dV j , j = 1, 2, … , ne (7.15)
∫V j
The stress and strain tensors used in this equation are defined in the reference configura-
tion. The virtual changes in the strains can be expressed in terms of the virtual changes of
the position vector gradients as
[( T ) T ( )]
1
𝛿𝛆j = 𝛿Jj Jj + Jj 𝛿Jj , j = 1, 2, … , ne (7.16)
2
[ ]
where Jj = 𝜕rj ∕𝜕x 𝜕rj ∕𝜕y 𝜕rj ∕𝜕z is the matrix of position vector gradients at an
arbitrary point on element j. The second Piola-Kirchhoff stresses are related to the
Green-Lagrange strains using the constitutive equations
j
𝛔P2 = Ej ∶ 𝛆j , j = 1, 2, … , ne (7.17)

where Ej is the fourth-order tensor of elastic coefficients. Substituting the preceding two
equations into Eq. 15, it can be shown that the virtual work of the stresses of ANCF element
j can be written as
j 1 ( j ) [( T ) T ( )]
𝛿Ws = − E ∶ 𝛆j ∶ 𝛿Jj Jj + Jj 𝛿Jj dV j
2 ∫V j
jT
= −Qs 𝛿ej , j = 1, 2, … , ne (7.18)
j
where Qs is the vector of the elastic forces associated with the nodal coordinates of ANCF
element j.
The cable element is a gradient-deficient element, and therefore, a complete set of gradi-
ent vectors cannot be defined. For this reason, classical beam theory formulations can be
used. In this case, the elastic forces of the cable element can be formulated using the virtual
306 Mathematical Foundation of Railroad Vehicle Systems

work or the strain energy. The virtual work of the elastic forces for the cable element can
be written as
l l
j j j 1
𝛿Ws = Ej Aj 𝜀11 𝛿𝜀11 dx + Ej I j 𝜅 j 𝛿𝜅 j dx, j = 1, 2, … , ne (7.19)
∫0 2 ∫0
j
where 𝜀11 is the axial strain, Ej is the modulus of elasticity, Aj is the cross-section area, I j is
the second moment of area, and 𝜅 j is the curvature. The elastic forces can also be evaluated
using the strain energy of cable element j. The strain energy can be written as
l ( )2 l ( )2
1 j 1
Uj = Ej Aj 𝜀11 dx + Ej I j 𝜅 j dx, j = 1, 2, … , ne (7.20)
2 ∫0 2 ∫0
Using this expression of strain energy, the vector of generalized elastic forces associated
j ( )
j T
with the element nodal coordinates can be defined as Qs = − 𝜕U j ∕𝜕e ( T . The) axial
j j j j j
strain 𝜀11 can be defined using position-gradient vector rx as 𝜀11 = (1∕2) rx rx − 1 . The
curvature 𝜅 j can be defined as the norm of the curvature vector obtained by differen-
j | j| | j|
tiating unit tangent rx ∕ |rx | with respect to the arc length s, where ds = |rx | dx: that is,
( ) [ ( | | ) ] (| | )
j | j| j | j| j | j|
𝜕 rx ∕ |rx | ∕𝜕s = 𝜕 rx ∕ |rx | ∕𝜕x (𝜕x∕𝜕s), which can be written as 𝜕 rx ∕ |rx | ∕𝜕s =
[ ( | | ) ]( | |
) | |
j | j| | j|
𝜕 rx ∕ |rx | ∕𝜕x 1∕ |rx | . Therefore, exact definitions of the axial strain and curvature
| | | |
can be used to develop the expression of the vector of the elastic forces of the ANCF
cable element.
While the mass matrices of the ANCF cable and fully parameterized beam elements
are constant and the generalized inertia forces are linear in the second derivatives of
the nodal coordinates, the generalized elastic (stress) forces of both elements are highly
nonlinear functions of the nodal coordinates. These elements, therefore, can capture the
elastic nonlinearities due to large displacements of catenary wires. As a result of using the
general continuum mechanics approach in the formulation of the elastic forces of the fully
parameterized beam element, and as a result of using a complete set of gradient vectors
that allow for the definition of all the components of the Green-Lagrange strain tensor,
the resulting nonlinear elastic forces account for more deformation modes as compared to
the gradient-deficient cable element. While these deformation modes can be important in
many applications, they can also be a source of high frequencies that force the numerical
integration routine to take smaller time steps, making computer simulations using fully
parameterized beam-element models less efficient as compared to simulations based on
cable-element models.

7.4 CATENARY EQUATIONS OF MOTION


In Chapter 6, the principle of virtual work in dynamics was derived using a system of
particles. A continuum can be assumed to consist of an infinite number of particles, and
therefore, the principle of virtual work is also applicable to flexible bodies. For example,
in the case of n particles, the virtual work of the inertia forces can be written as 𝛿Wi =
∑np i iT i p
̈ 𝛿r , where mi is the particle mass and ri is its global position vector. If a finite
i=1 m r
element j is assumed to consist of an infinite number of particles, the summation can be
Pantograph/Catenary Systems 307

replaced by integration and mi can be replaced by 𝜌j dV j , where dV j is an infinitesimal


volume. It follows that the virtual work of the inertia force of the finite element can be
written as 𝛿Wi = ∫V j 𝜌j r̈ j 𝛿rj dV j , as previously used in this section.
T

It was shown in Chapter 6 that the principle of virtual work can be written as 𝛿W i = 𝛿W e ,
where 𝛿W i is the virtual work of the system inertia forces and 𝛿W e is the virtual work of the
system applied forces. In the case of flexible bodies, the virtual work principle can account
for the effect of internal elastic forces by including the virtual work of the system elastic
forces 𝛿W s . In the case of unconstrained motion, the equations of motion of the finite ele-
ments that form a catenary can be developed using the principle of virtual work defined as
𝛿Wi = 𝛿Ws + 𝛿We (7.21)
In the case of a flexible catenary, the virtual work 𝛿W e of applied forces such as gravity, aero-
dynamics, and contact forces can be systematically developed. If, for example, an external
force vector F j acts at a point P defined by coordinates xP on finite element j, the virtual
work of this force vector can be written using the ANCF displacement field as
j T j T ( ) jT
𝛿We = Fj 𝛿rP = Fj Sj xP 𝛿ej = Qe 𝛿ej (7.22)
where Sj (xP ) is a constant matrix that defines the element shape function matrix at point
j
xP , and Qe is the vector of generalized forces associated with the element nodal coordinates
e as a result of the application of force vector Fj . This vector of generalized forces is defined
j
j T ( )
using the preceding equation as Qe = Sj xP Fj . The generalized forces associated with
the element nodal coordinates can be systematically defined in cases of both concentrated
and distributed external forces for both the fully parameterized and gradient-deficient finite
elements discussed in this section. Therefore, one can write the virtual work of the applied
forces acting on the catenary by summing up the virtual work of the forces of its finite
elements as

ne
j

ne
jT
𝛿We = 𝛿We = Qe 𝛿ej (7.23)
j=1 j=1

Using the expression for the virtual work of the inertia forces of the ANCF element
obtained in the preceding section, the virtual work of the inertia forces of the catenary can
be obtained by summing up the virtual work of the inertia forces of its ANCF elements as

ne
j

ne
( j j )T j
𝛿Wi = 𝛿Wi = M ë 𝛿e (7.24)
j=1 j=1

Using the expression of the virtual work of the elastic forces of the finite element obtained
in the preceding section, the virtual work of the elastic forces of the FE catenary model can
be written as

ne
j

ne
jT
𝛿Ws = 𝛿Ws = − Qs 𝛿ej (7.25)
j=1 j=1

Substituting the preceding three equations into the principle of virtual work of Eq. 21 yields
ne ( )T
∑ j j
Mj ë j + Qs − Qe 𝛿ej = 0 (7.26)
j=1
308 Mathematical Foundation of Railroad Vehicle Systems

This equation is equivalent to the following scalar equation, which is written in a more
explicit form in terms of the element vectors and matrices:
T
⎧⎡ 1 ⎫
⎪⎢M 𝟎 … 𝟎 ⎤ ⎡ ë 1 ⎤ ⎡ Q1s ⎤ ⎡ Q1e ⎤⎪ ⎡ 𝛿e1 ⎤
⎪⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎪ ⎢ ⎥
⎪⎢ 𝟎 M
2 … 𝟎 ⎥ ⎢ ë 2 ⎥ ⎢ Q2s ⎥ ⎢ Q2e ⎥⎪ ⎢ 𝛿e2 ⎥
⎨⎢ ⎥ ⎢ ⎥ + ⎢ ⎥ − ⎢ ⎥⎬ ⎢ ⎥=0 (7.27)
⎪⎢ ⋮ ⋮ ⋱ 𝟎 ⎥ ⎢ ⋮ ⎥ ⎢ ⋮ ⎥ ⎢ ⋮ ⎥⎪ ⎢ ⋮ ⎥
⎪⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎪ ⎢ ⎥
… M ⎥⎦ ⎢⎣ë ne ⎥⎦ ⎢⎣Qs e ⎥⎦ ⎢⎣Qe e ⎥⎦⎪ ⎢𝛿ene ⎥
n n
⎪⎣ 𝟎 𝟎 ne
⎣ ⎦
⎩ ⎭
In the FE analysis, the element Boolean matrix Bj is introduced to define the element con-
nectivity. This Boolean matrix is used to develop a standard procedure for the assembly
of the elements by mapping the element nodal coordinates to catenary nodal coordinate
vector
[ ̈ Therefore, one can write Eq. 26 as
e. In this case, one has]𝛿ej = Bj 𝛿e and ë j = Bj e.
∑ne ( j j j j
)T
j=1 M B e ̈ + Qs − Qe B 𝛿e = 0. In the case of unconstrained motion, the elements
j

of vector e are independent, and therefore, their coefficients in this equation must be equal
∑ne ( jT j j j j
)
B M B ë + Bj Qs − Bj Qe = 𝟎. Performing the summation in
T T
to zero. This leads to j=1
this equation, it can be shown that the catenary FE equations of motion can be written as
Më + Qs − Qe = 𝟎 (7.28)
∑ne jT j j ∑ne jT j
where M = j=1 B M B is the catenary symmetric mass matrix, Qs = j=1 B Qs is the
∑ne jT j
vector of catenary elastic forces, and Qe = j=1 B Qe is the vector of the catenary applied
forces. The mass matrix in the preceding equation remains constant and symmetric, and
therefore, a Cholesky transformation can be used to obtain an identity generalized mass
matrix, which has an optimum sparse matrix structure. Therefore, if an elastic contact for-
mulation (ECF) is used to describe pantograph/catenary interaction, the solution for the
catenary accelerations does not require an iterative LU factorization or finding the inverse of
the mass matrix during the dynamic simulation. After solving for the ANCF Cholesky accel-
erations, the ANCF nodal accelerations of the catenary can be obtained using the Cholesky
transformation (Shabana 2018).

7.5 PANTOGRAPH/CATENARY CONTACT FRAME


In the analysis presented in this chapter, it is assumed that catenary wires are modeled
using ANCF finite elements, which allow for systematic definitions of the gradient vec-
tors that enter into the formulation of pantograph/catenary contact forces. To develop a
unified approach that can be used for both the gradient-deficient cable and fully parameter-
ized beam elements, the contact conditions are formulated using a single position-gradient
vector. Recall that the gradient-deficient ANCF cable element has only the longitudinal
position-gradient vector rx , while the fully parameterized ANCF beam element has three
position vector gradients rx , ry , and rz . Therefore, to develop a unified approach that allows
for using either of the two ANCF elements, the contact equations are formulated using the
longitudinal gradient vector rx , which is tangent to the element centerline. In the notations
used in this section and the following sections, superscript c refers to the catenary; this is
Pantograph/Catenary Systems 309

with the understanding that positions of arbitrary points on the ANCF finite elements are
directly defined in the global coordinate system, and for such an ANCF catenary model,
there is no need to introduce a local body coordinate system.
Pantograph/catenary forces can have both normal and tangential components that
depend on the locations of the contact points between the pan-head and catenary. To
properly define the components of these contact forces, a coordinate system X c Y c Z c ,
referred to as the contact frame, is introduced at the contact point, as shown in Figure 8.
The orientation of this coordinate system in the global coordinate system is assumed to be
[ ]
defined by the orthogonal transformation matrix Ac = ic jc kc , where ic , jc , and kc are
orthogonal unit vectors along the axes of the coordinate system X c Y c Z c . The longitudinal
axes of the contact frame X c Y c Z c can be defined by unit vector ic along ANCF gradient
vector rcx of the catenary wire, which comes into contact with the pan-head. The other two
axes defined by unit vectors jc and kc can be determined as unit vectors perpendicular to
ANCF gradient vector rcx at the contact point. These three orthogonal unit vectors can be
defined as (Pappalardo et al. 2016)
rc nc nc
ic = ‖ xc ‖ , jc = ‖ 1c ‖ , kc = ‖ 2c ‖ (7.29)
‖rx ‖ ‖n1 ‖ ‖n2 ‖
and vectors rx , n1 , and n2 are
c c c

( ) ( )( ) ( )
⎡ rcx ⎤ ⎡ − rcx 1 rcx 2 ⎤ ⎡− rcx ⎤
⎢( ) ⎥ 1
⎢(( ) ) (( ) )2 ⎥ ⎢ 3

rcx = ⎢ rcx 2 ⎥ , nc1 = ⎢ rcx 1 + rcx 3 ⎥ , nc2 = ⎢ 0 ⎥
2
(7.30)
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢(rc ) ⎥ ⎢ ( c) ( c) ⎥ ⎢ (rc ) ⎥
⎣ x 3⎦ ⎣ − r r
x 2 x 3 ⎦ ⎣ x 1⎦
The norms of vectors nc1 and nc2 become zero if ANCF position gradient vector rcx is parallel
[ ]T
to vector j = 0 1 0 . This leads to a singularity that can be avoided by using a different
definition of the orthogonal vectors of Eq. 30. In this singular configuration, which is not
encountered in the case of the catenary system, the following three orthogonal vectors can

Zc

Xc
Cate
nary
cont
act w
c ire
Y

Figure 7.8 Pantograph/catenary contact.


310 Mathematical Foundation of Railroad Vehicle Systems

be used instead of those defined by the preceding equation (Gere and Weaver 1965; Shabana
2019):
( )
⎡ 0 ⎤ ⎡− rcx 2 ⎤ ⎡0⎤
⎢( c ) ⎥ ⎢ ⎥ ⎢ ⎥
rx = ⎢ rx 2 ⎥ , n1 = ⎢ 0 ⎥ , n2 = ⎢0⎥
c c c
(7.31)
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦ ⎣ 0 ⎦ ⎣1⎦
It is clear from the preceding two equations that in both the non-singular and singular
configurations, the pantograph/catenary contact frame X c Y c Z c is defined using a single
gradient vector, which is the longitudinal gradient vector rcx at the contact point. There-
fore, such a definition of the contact frame can be used with both the ANCF cable element,
which has only longitudinal gradient vector rx , and the fully parameterized ANCF beam
element, which has three gradient vectors rx , ry , and rz .
As in the case of wheel/rail contact, two fundamentally different formulations can be used
to predict pantograph/catenary contact forces: the constraint contact formulation (CCF) and
the elastic contact formulation (ECF). In the constraint contact formulation, the pan-head
is assumed to remain in contact with the catenary wire, and therefore, separations or pen-
etrations are not allowed. In this case, the freedom of the pan-head to translate along the
normal to the catenary wire is eliminated. The pan-head, however, can move relative to
the catenary wire in the longitudinal and lateral directions. The normal contact force in the
CCF approach is determined as a constraint (reaction) force. In the elastic contact formula-
tion, on the other hand, pantograph/catenary separations and penetrations are allowed and
no constraints are imposed on the motion of the pan-head with respect to the catenary wire.
In this case, the pantograph/catenary normal contact force is described using a compliant
force model with assumed stiffness and damping coefficients. The constraint and elastic
contact formulations are described in the following two sections, respectively.

7.6 CONSTRAINT CONTACT FORMULATION (CCF)


The constraint contact formulation is recommended in simulation scenarios in which it is
desirable to avoid the oscillations and high frequencies that result when the elastic contact
formulation is used. Avoiding these oscillations and high frequencies allows for smoother
solutions that can be used to shed light on dynamic behaviors and phenomena that can be
difficult to observe or identify when high-frequency oscillations are superimposed on the
solutions. For this reason, it is recommended to implement both constraint and elastic con-
tact formulations in computer software used for nonlinear dynamic simulations of railroad
vehicle systems.
Two CCF approaches can be used to formulate the constraint contact conditions between
the pantograph head and catenary. In the first approach, the pan-head is not allowed
to slide with respect to the catenary wire in the lateral direction, while in the second
approach, lateral sliding is allowed. Both constraint formulations require using the concept
of non-generalized coordinates, previously introduced in this book. As in the preceding
section, it is assumed in the analysis presented in this section that the catenary wire is
modeled using ANCF finite elements, while the pantograph is modeled as a system of
interconnected bodies. The pan-head is assumed to be a rigid body whose configuration is
Pantograph/Catenary Systems 311

[ T T ]T
defined using the Cartesian absolute coordinates qp = Rp 𝛉p introduced in Chapter
3, where superscript p refers to the pan-head, Rp is the vector that defines the global
position of the origin of the pan-head body coordinate system X p Y p Z p with respect to the
global coordinate system, and 𝛉p is the set of rotational coordinates used to define the
orientation of the pan-head body coordinate system. Using these coordinates, the global
position of a point on the pan-head can be written as rp = Rp + Ap up , where up is the
vector that defines the position of the point with respect to the origin of the pan-head body
coordinate system X p Y p Z p . The location of the origin of the coordinate system X p Y p Z p
is assumed to be the center of mass of the link to which the pan-head is attached. The
global position vector of an arbitrary point on the catenary wire can simply be written as
rc = Sc ec , where superscript c refers to the catenary, and Sc and ec are, respectively, the
shape function matrix and vector of nodal coordinates of the ANCF element on which the
contact point lies.

Longitudinal Relative Sliding If the effect of the relative lateral displacement between
the pan-head and catenary wire is not considered, the constraint contact conditions can be
represented by a sliding joint that describes the movement of the pan-head along the flexible
catenary (Seo et al. 2006). The sliding joint kinematic conditions, which are used to model
the relative motion between the rigid pan-head of the pantograph and the flexible catenary
cable, can be solved to determine the location of the contact points on the pan-head and
catenary wire. To determine the locations of the contact points, the arc length parameter
s that defines the distance traveled by the pan-head with respect to the catenary wire is
introduced. This arc length parameter, which is associated with the space curve that defines
the centerline of the ANCF cable or fully parameterized beam-element catenary model, can
be written in terms of the spatial longitudinal coordinate x of the finite element: that is,
s = s(x). Using this functional relationship, arc length parameter s is associated with the
reference configuration, despite the fact that its value depends on the location of the point
of contact between the pan-head and the catenary wire.
As a result of introducing arc length parameter s, three constraint equations are required
to solve for s (one equation) and eliminate the two translational degrees of freedom of
the pan-head with respect to the catenary wire in the lateral and normal directions (two
equations). These constraint contact conditions can then be written as
( )
Cs = Cs qp , ec , s, t = rp − rc = 𝟎 (7.32)

where rp = rp (qp , t) is the global position vector of contact point P on the pan-head,
rc = rc (ec , s, t) is the global position vector of contact point P on the catenary, qp = qp (t)
is the vector of generalized coordinates of the pantograph link to which the pan-head
is attached, ec = ec (t) is the vector of the nodal coordinates of the finite element on the
catenary on which the contact point lies, and s = s(t) is the non-generalized arc length
coordinate used to determine the location of the contact point on the catenary cable.
The vectors in the three constraint conditions of Eq. 32 are defined in the global coordi-
nate system. An equivalent formulation for the constraint conditions of Eq. 32 is to define
the components of these vectors in the pantograph/catenary contact frame X c Y c Z c intro-
duced in the preceding section and whose axes are defined using the longitudinal ANCF
312 Mathematical Foundation of Railroad Vehicle Systems

position vector gradients rx , as explained in the preceding section. In this case, the preced-
ing equation can be written in terms of vector components defined along the axes of the
coordinate system X c Y c Z c as (Pappalardo et al. 2016)
⎡ ic (rp − rc ) ⎤
T

⎢ T ⎥
Cs = ⎢ jc (rp − rc ) ⎥ = 𝟎 (7.33)
⎢ cT ( P )
c ⎥
⎣k r − r ⎦
where ic = ic (ec , s, t), jc = jc (ec , s, t), and kc = kc (ec , s, t) are the three unit vectors along
the axes of the coordinate system X c Y c Z c defined on the catenary cable at the contact point,
as previously explained. As explained in the preceding section, the definition of the panto-
graph/catenary contact frame is based on only one gradient vector instead of three gradient
vectors. This allows for developing a general sliding joint formulation that can be used with
both ANCF fully parameterized and gradient-deficient finite elements.

Longitudinal and Lateral Sliding The three constraint equations of Eq. 32 or, equiva-
lently, Eq. 33 can be used to determine the arc length parameter s and eliminate two degrees
of freedom of relative translation of the pan-head with respect to the catenary wire in the
lateral and normal directions. For example, given the generalized coordinates of pan-head
T
qp and catenary nodal coordinates ec , the first equation ic (rp − rc ) = 0 of Eq. 33 can be
solved iteratively to determine arc length parameter s. This solution for s can be substituted
into the last two equations of Eq. 33 to restrict pan-head motion with respect to the catenary
wire in lateral and normal directions.
The same three equations of Eq. 32 or, alternatively, Eq. 33 can be used to develop a
constraint contact formulation that allows for both longitudinal and lateral displacements
of the pan-head with respect to the catenary wire. In this case, the lateral component
of vector up in the equation rp = Rp + Ap up is allowed to vary and is treated as an
unknown geometric parameter. Without any loss of generality, the coordinate system of
the pan-head body can be selected such that local position vector up can be written as
[ p p p ]T p p p
up = u1 yl u3 , where u1 and u3 are assumed to be constants, while yl is a parameter
that can vary and can be determined by solving the constraint equations. In this case,
p
two of the constraint equations can be used to solve for the parameters s and yl , while
the remaining constraint equation eliminates the freedom of the pan-head to translate
with respect to the catenary wire along the normal direction. For example, given the
generalized coordinates of pan-head qp and catenary nodal coordinates ec , the first two
T T p
equations ic (rp − rc ) = 0 and jc (rp − rc ) = 0 of Eq. 33 can be used to solve for s and yl ,
cT p c
while the third equation k (r − r ) = 0 is used to eliminate the freedom of the pan-head
to move with respect to the catenary wire in the normal direction. By introducing the
p
additional parameter yl , lateral motion of the pan-head is allowed, and the constraints on
the generalized coordinates are reduced to only one because two constraint equations are
p
used to determine the unknown geometric parameters s and yl .

Augmented Formulation and Embedding Technique The constraint contact


conditions introduced in this section can be used with the augmented formulation or the
embedding technique discussed in Chapter 6. In the augmented formulation, the constraint
equations are combined with the differential equations of motion using the technique of
Pantograph/Catenary Systems 313

Lagrange multipliers. This approach allows for using non-generalized coordinates that
represent geometric variables that have no associated inertia or applied forces. The normal
pantograph/catenary contact forces in the augmented formulation can be evaluated using
Lagrange multipliers as reaction forces.
In alternate augmented formulations, the geometric parameters can be systematically
eliminated, leading to a number of constraint contact conditions equal to the number of
degrees of freedom eliminated. In this second augmented formulation, the non-generalized
geometric parameters do not appear in the equations of motion.
In the embedding technique, on the other hand, all the dependent generalized and
non-generalized coordinates are eliminated, leading to a number of equations of motion
equal to the number of system degrees of freedom. In general, a systematic procedure
can be used to eliminate the geometric parameters to obtain a number of constraint
equations equal to the number of degrees of freedom eliminated in the constraint contact
formulation.

Elimination of Geometric Parameters To demonstrate the systematic procedure for


eliminating geometric parameters, the case of the constraint formulation that only allows
for relative sliding is considered (Pappalardo et al. 2016). A similar procedure can be used
when relative lateral sliding between the pan-head and catenary contact wire is allowed.
If lateral sliding is not allowed, one geometric parameter s is introduced, and the relative
motion is governed by the constraints of Eq. 32 or Eq. 33. In this section, the procedure of
eliminating arc length parameter s is demonstrated using Eq. 33. The first scalar equation
T
in this equation, ic (rc − rp )T = 0, is equivalent to the equation Ces = (rc − rp )T rcx = 0. The
algebraic equation Ces = (rc − rp )T rcx = 0 can be used to eliminate arc length parameter s.
Assuming that the generalized coordinates are known from the numerical integration of
the equations of motion, the algebraic equation Ces = (rc − rp )T rcx = 0 can be considered a
nonlinear equation in arc length parameter s. Using a Newton–Raphson algorithm, this
( )
equation can be used to determine s by iteratively solving the equation 𝜕Ces ∕𝜕s Δs = −Ces ,
T ( )
where Δs is the Newton difference, 𝜕Ces ∕𝜕s = (𝜕rc ∕𝜕s)T rcx + rcp 𝜕rcx ∕𝜕s , and rcp = rc − rp .
Having determined s, the location of the contact point on the catenary can be determined for
( ) ( )
a given position of the pan-head. Using this solution, one can write Ces q 𝛿q + Ces s 𝛿s = 0,
( s) ( s) [ ] ( )
where Ce q is the row vector Ce q = 𝜕Ces ∕𝜕q = 𝜕Ces ∕𝜕qp 𝜕Ces ∕𝜕ec , Ces s is the scalar
( s) [ T T ]T
Ce s = 𝜕Ces ∕𝜕s, and q = qp ec is the vector of generalized coordinates of the pan-head
( )
body and ANCF catenary. Assuming that Ces s is different from zero, one has
(( ) ( ) )
𝛿s = − Ces q ∕ Ces s 𝛿q ⎫

(( ) ( ) ) ⎪
ṡ = − Ce q ∕ Ce s q̇
s s
⎬ (7.34)
(( ) ( ) ) ( ) ⎪
s̈ = − Ces q q̈ + Qse c ∕ Ces s ⎪

( s)
where Qe c is a vector that is quadratic in velocities, which arises from differentiating the
( ) ( )
constraint equation at the velocity level Ces q q̇ + Ces s ṡ = 0 with respect to time. This dif-
( ) ( ( ) ) ( ) ( ( ) )
ferentiation leads to Ces q q̈ + d Ces q ∕dt q̇ + Ces s s̈ + d Ces s ∕dt ṡ = 0. This equation
( ) ( ) ( ( ) ) ( ( ) )
defines Qse c as Qse c = d Ces q ∕dt q̇ + d Ces s ∕dt s.̇ The value of s obtained using the
314 Mathematical Foundation of Railroad Vehicle Systems

Newton–Raphson iterations and virtual change and the first and second time derivatives of
s given by Eq. 34 can be used to eliminate the dependence of the second and third equations
of Eq. 33 and their time derivatives on s and its time derivatives. These two equations can
be written at the position level as
[ T ]
jc p − rc
(r )
Csm = T ( ) =𝟎 (7.35)
kc rP − rc
The virtual change and the first and second time derivatives of these two equations can
written, respectively, as
( ) ( ) ⎫
𝛿Csm = Csm q 𝛿q + Csm s 𝛿s = 𝟎 ⎪
( ) ( ) ⎪
Ċ sm = Csm q q̇ + Csm s ṡ = 𝟎 ⎬ (7.36)
( ) ( ( ) ) ( ) ( ( ) ) ⎪
C̈ sm = Csm q q̈ + d Csm q ∕dt q̇ + Csm s s̈ + d Csm s ∕dt ṡ = 𝟎 ⎪

( ) ( )
where Csm q = 𝜕Csm ∕𝜕q and Csm s = 𝜕Csm ∕𝜕s. Using the results of Eq. 34 with Eq. 36, the
dependence of the constraint equations Csm and their derivatives on arc length parameter s
can be eliminated, leading to
( )
( s) 1 ( s ) ( s) ⎫
𝛿Cm =s Cm q − ( s ) Cm s Ce q 𝛿q = 𝟎 ⎪
Ce s ⎪
( ) ⎪
( s) 1 ( s ) ( s) ⎪
̇Csm = Cm q − ( s ) Cm s Ce q q̇ = 𝟎 ⎬ (7.37)
Ce s ⎪
( ( s ) ( s) ) ⎪
( ) Cm s Ce q ( ) ⎪
C̈ sm = Csm q − ( ) q̈ − Qsm c = 𝟎 ⎪
Ces s ⎭
( s)
where Qm c is a vector that is quadratic in velocities and results from differentiating the
constraint equations twice with respect to time. Because the vector of constraint functions
Csm has two scalar functions, eliminating the arc length parameter leads to a number of
constraint equations equal to the number the system degrees of freedom eliminated by the
pantograph/catenary contact conditions. The details of the scalars, vectors, and matrices
that appear in the equations used in this section are presented by Pappalardo et al. (2016).
A similar procedure can also be used in the case of longitudinal and lateral sliding to
p
eliminate the dependence on geometric parameters s and yl and obtain one constraint
equation expressed in terms of only the generalized coordinates, because one degree of
freedom is eliminated in this case.

7.7 ELASTIC CONTACT FORMULATION (ECF)

The elastic contact formulation allows for pan-head/catenary penetrations and separa-
tions, and therefore, no degrees of freedom are eliminated when this formulation is used.
Algebraic equations are used to solve for geometric parameters that define the location
of the potential contact points on the pan-head and catenary wire. Distance equations
Pantograph/Catenary Systems 315

are then used to check whether contacts occur between the pan-head and contact wire.
Predicting possible separations between the pan-head and contact wire is important since
such separations are the cause of arcing, which has an adverse effect on current-collection
quality. This section discusses two cases in which lateral displacement of the pan-head
relative to the contact wire is restricted or unrestricted, respectively. The equations
that govern these two cases are similar to the equations used in the constraint contact
formulation discussed in the preceding section.

Longitudinal Relative Sliding In this case, the pan-head is assumed to slide with
respect to the catenary wire in the longitudinal direction, while the pan-head displacement
relative to the catenary wire in the plane perpendicular to the longitudinal tangent is
assumed to be restricted. The algebraic equations of Eq. 33 can be used as the basis for
the elastic contact formulation discussed in this section. As in the CCF approach, arc
length geometric parameter s is used to define the location of the contact point between
the catenary and pan-head. The configuration of the pan-head is assumed to be known:
that is, the vector of generalized coordinates qp = qp (t) is assumed to be known from the
numerical integration of the equations of motion of the system. Similarly, the vector of
nodal coordinates ec = ec (t) of the catenary is assumed to be known. Knowing these gen-
T
eralized coordinates, the first equation in Eq. 33, ic (rp − rc ) = 0, can be used to determine
the arc length parameter s that defines the location of the contact point. Recalling that
[( ) ( ) ( ) ]T
ic = rcx ∕ ||rcx ||, where rcx = rcx 1 rcx 2 rcx 3 is the longitudinal ANCF gradient vector
tangent to the catenary wire at the potential contact point and s = s(t), the algebraic
T
equation ic (rp − rc ) = 0 can be written in a more convenient form as (Kulkarni et al. 2017)
( )T
g (s) = rc (s) − rp rcx (s) = 0 (7.38)
Knowing the system generalized coordinates, this equation can be considered a nonlin-
ear algebraic equation in arc length parameter s. Due to the nonlinearity of this equation,
an iterative Newton–Raphson algorithm can be used to obtain the solution for arc length
parameter s. In this Newton–Raphson iterative procedure, one must solve the equation
T ( )
(𝜕g/𝜕s)Δs = − g for Newton-differenceΔs, where gs = 𝜕g∕𝜕s = (𝜕rc ∕𝜕s)T rcx + rcp 𝜕rcx ∕𝜕s
and rcp = rc − rp . Assuming that gs ≠ 0, the solution of Eq. 38 defines arc length parameter
s , which can be used to define the location of the potential contact point on the catenary
wire rc (s). Using the value of s, the last two equations in Eq. 33 can be used to define the
following lateral and normal distances, respectively:
T ( ) T ( )
dl = jc rp − rc , dn = kc rP − rc (7.39)
These two distances along two orthogonal directions perpendicular to the longitudinal tan-
gent can be used to check whether contact occurs between the catenary pan-head and
contact wire. If d1 and d2 are smaller than certain tolerances, contact between the pan-head
and catenary wire is assumed.

Longitudinal and Lateral Sliding If lateral motion of the pan-head with respect to the
catenary wire is not restricted, the first two algebraic equations of Eq. 33 can be used to
p p
solve for geometric parameters s and yl , where yl defines the lateral relative displacement
316 Mathematical Foundation of Railroad Vehicle Systems

of the pan-head with respect to the catenary wire, as discussed in the preceding section.
These two nonlinear algebraic equations can be written as
( p) ( ( p) ) ⎫
g1 = g1 s, yl = (ic (s))T rp yl − rc (s) = 0 ⎪
( p) ( )T ( p ( p ) ) ⎬ (7.40)
g2 = g2 s, yl = jc (s) r yl − rc (s) = 0 ⎪

These two nonlinear algebraic equations can be solved iteratively using a Newton–Raphson
p
algorithm to determine s and yl . To this end, the following system of equations is con-
structed:
[ p
][ ] [ ]
𝜕g1 ∕𝜕s 𝜕g1 ∕𝜕yl Δs g1
p p
=− (7.41)
𝜕g2 ∕𝜕s 𝜕g2 ∕𝜕yl Δyl g2
p
The convergence of the Newton–Raphson iterations defines s and yl , which can be substi-
T ( )
tuted into the equation kc rP − rc to calculate the distance
( ( ))T ( P ( p p ) ( ))
dn = k c e c , s r q , yl − rc ec , s (7.42)

If dn is smaller than a given tolerance, contact between the pan-head and catenary wire is
assumed. If this condition is not satisfied, separation is assumed, and no interaction forces
between the pan-head and catenary wire are applied.

Constraint and Elastic Formulations The two different cases of contact, without and
p
with lateral sliding yl of the pan-head relative to the contact wire, are considered in both
the constraint and elastic contact formulations presented in this chapter. Capturing this
lateral sliding is necessary for developing accurate pantograph/catenary wear models in
many motion scenarios. Localized contact in a very small area leads to significant wear
that can damage the carbon strip used for current collection. Such wear can be reduced
if the contact wire slides laterally over the pan-head to avoid contact being confined to a
small area on the pan-head. During curve negotiations, the contact wire sweeps over the
pantograph carbon strip, so the contact point is not restricted to a localized region. In the
case of a tangent track, the catenary is staggered to assume a zigzagging shape in order for
the contact wire to slide laterally over the carbon strip and avoid having contact in a small
region. Such lateral sliding contributes to reducing the wear of the carbon strip. Therefore,
p
in investigations focused on wear, it is necessary to consider the effect of lateral sliding yl
to obtain realistic results that are consistent with actual catenary design.
Unlike the algebraic equations used in the constraint contact formulation described in
the preceding section, the algebraic equations used in the elastic contact formulation do
not eliminate degrees of freedom, and they do not represent motion constraints. The ECF
algebraic equations are mainly used to determine the geometric parameters that define the
locations of pan-head/catenary contact points. CCF algebraic equations eliminate degrees
of freedom and must be satisfied at the position, velocity, and acceleration levels. When
ECF algebraic equations are used, on the other hand, normal contact forces are defined
using compliant force models with assumed stiffness and damping coefficients. In the case
of separation, the compliant contact forces are assumed to be zero. When CCF algebraic
equations are used, on the other hand, normal contact forces are determined as constraint
Pantograph/Catenary Systems 317

(reaction) forces using the techniques described in Chapter 6. In this case, no assumption of
stiffness and damping coefficients needs to be made. Therefore, while the two approaches
employ similar algebraic equations, they are fundamentally different and require using dif-
ferent solution procedures, as is the case with wheel/rail contact formulations. The two
contact formulations lead to a different number of degrees of freedom and a different num-
ber of system constraint forces.

7.8 PANTOGRAPH/CATENARY EQUATIONS AND MBS


ALGORITHMS
The pantograph/catenary contact formulations presented in this chapter can be system-
atically integrated with general computational MBS algorithms that allow for developing
complex railroad vehicle system models. These MBS algorithms also allow for modeling
flexible bodies using the FE method and are based on fully nonlinear formulations of the
system equations of motion, as discussed in Chapter 6. The development of such general
computational algorithms for railroad vehicle systems requires introducing different sets
of coordinates required for modeling components with different degrees of flexibility. Some
components can be bulky and can be modeled as rigid bodies; other components are flex-
ible but experience only small deformations that can lead to very high stresses, as is the
case with stiff components; still other components, such as catenary wires, are very flexi-
ble and experience very large deformations. All these types of bodies, which have different
degrees of flexibility, can exist in one railroad vehicle model. Different coordinates are used
to describe the motion of rigid bodies, flexible bodies, and very flexible bodies.

Coordinate Types General MBS solution algorithms are designed to model systems
that consist of interconnected bodies, which can be rigid or flexible. Flexible bodies that
experience small deformations are often modeled in computational MBS algorithms using
the floating frame of reference (FFR) formulation. The FFR formulation, which allows for
coordinate reduction using component mode synthesis techniques, can be used to solve
small-deformation problems efficiently. In the FFR formulation, a coordinate system is
assigned to each flexible body. The configuration of the flexible body that experiences small
deformations is defined by a set of reference coordinates that define the translation and
orientation of the body coordinate system and a set of elastic coordinates that define the
deformation of the body with respect to its coordinate system.
The vector of generalized coordinates used to describe the motion of rigid bodies and the
reference motion of the body coordinate systems in the FFR formulation is denoted as qr .
For both rigid bodies and the coordinate systems of the FFR bodies, the vector of general-
ized coordinates qr includes Cartesian coordinates that define the global positions of the
origins of the coordinate systems of the rigid and FFR bodies. The small elastic deforma-
tions of FFR bodies are described using the vector of elastic coordinates qf , which defines
the deformations of FFR bodies with respect to their coordinate systems. The rigid-body
and FFR formulation are suited for developing models of pantograph systems that consist of
interconnected rigid and deformable bodies. Pantograph links can experience small defor-
mations, and therefore, the FFR formulation is more suited for such systems. The small
318 Mathematical Foundation of Railroad Vehicle Systems

deformations of relatively stiff members can lead to very high stresses that can result in
component failures.
In addition to the two sets of coordinates qr and qf , other coordinate types need to be
introduced to be able to develop realistic and more accurate virtual prototyping of railroad
vehicle system models. As discussed in this chapter, the catenary system can be modeled
using ANCF finite elements, which can be used to solve accurately nonlinear large dis-
placement and large deformation problems. General computational MBS algorithms allow
for introducing the ANCF coordinates e to develop catenary equations of motion with-
out the need to use incremental-rotation or co-simulation techniques. Furthermore, in the
augmented formulation of the equations of motion, the non-generalized geometric param-
eters s, used to describe curve and surface geometries and define the locations of wheel/rail
and pantograph/catenary contact points, enter into the formulation of the kinematic condi-
tions if the constraint contact formulations are used to model pantograph/catenary and/or
wheel/rail contacts.

MBS Formulation A general multibody system such as a railroad vehicle system may
consist of rigid bodies, flexible bodies, and very flexible bodies subjected to kinematic
motion constraints and external forces. The motion of the rigid bodies is governed by
the Newton-Euler equations presented in Chapter 6. The flexible bodies are assumed to
experience small deformations and can be efficiently modeled using the FFR formulation;
while the very flexible bodies can be modeled using ANCF finite elements. Therefore, the
vector of system generalized coordinates p used in MBS computational algorithms can be
[ ]T
written as p = qTr qTf eT sT , where qr defines the reference coordinates of the rigid
and FFR bodies, qf defines the small deformations of the FFR bodies with respect to their
body coordinate systems, e defines the vector of the system nodal coordinates of the ANCF
finite elements used to describe the nonlinear dynamics and large displacements, and s is
the vector of non-generalized coordinates or surface parameters used to describe the curve
and surface geometries in the formulation of the wheel/rail and/or pantograph/catenary
contact conditions. Using the augmented Lagrangian formulation, the system equations of
motion can be written using the principle of virtual work in dynamics as (Shabana 2019)
⎡Mrr Mrf 𝟎 𝟎 CTqr ⎤ ⎡q̈ r ⎤ ⎡ Qr ⎤
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ Mfr Mff 𝟎 𝟎 CTqf ⎥ ⎢q̈ f ⎥ ⎢ Qf ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ 𝟎 𝟎 Mee 𝟎 CTe ⎥ ⎢ ë ⎥ = ⎢ Qe ⎥ (7.43)
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ 𝟎 𝟎 𝟎 𝟎 CT⎥⎢ s ̈ ⎥ ⎢𝟎 ⎥
⎢ s ⎥⎢ ⎥ ⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ Cqr Cqf Ce Cs 𝟎 ⎦ ⎣ 𝛌 ⎦ ⎣Qc ⎦
where subscripts r, f , e, and s refer, respectively, to reference, flexible, ANCF, and
non-generalized coordinates; Mrr is the mass matrix associated with the reference motion
of the rigid and FFR bodies; Mff is the mass matrix associated with the FFR bodies that
experience small deformations; Mrf = MTfr is the mass matrix that defines the nonlinear
dynamic inertia coupling between the reference motion and the small elastic deformations
of the FFR bodies; Mee is the mass matrix of the ANCF finite elements used to describe the
large displacements of the flexible bodies; Cqr , Cqf , Ce , and Cs are the constraint Jacobian
Pantograph/Catenary Systems 319

matrices associated with coordinates qr , qf , e, and s, respectively; 𝝀 is the vector of


Lagrange multipliers, which can be used to define constraint forces; Qr , Qf , and Qe are the
generalized forces associated with the generalized coordinates qr , qf , and e, respectively;
and Qc is a quadratic velocity vector, which results from differentiating the constraint
equations twice with respect to time. This vector is defined by the constraint equation
at the acceleration level C̈ = Cq q̈ + Cs s̈ − Qc = 𝟎, where C is the vector of all constraint
[ ]T
functions and q = qTr qTf eT . Because the ANCF mass matrix Mee is constant, a
Cholesky transformation can be used to obtain an identity inertia matrix associated with
the ANCF Cholesky coordinates (Shabana 2018).
Because the FE catenary model can include a large number of nodal coordinates com-
pared to the number of coordinates used to describe vehicle and pantograph dynamics, it is
necessary to develop an efficient solution procedure for the equations used to form Eq. 43. It
is not always recommended to combine all these equations in one matrix equation unless all
the coordinates are kinematically coupled. When the elastic contact formulation is used to
model pantograph/catenary contact, the catenary system is not kinematically coupled with
the vehicle on which the pantograph is mounted. Therefore, the catenary system can be
considered a structure subjected to pantograph contact, gravity, pretension, and cross-wind
forces. In this case of an elastic contact formulation, Ce = 0 for the catenary, and the solu-
tion for the ANCF accelerations ë can be obtained separately without the need to make
this solution part of Eq. 43. Since the mass matrix Mee is constant and its LU factors can
be determined only once prior to the dynamic simulation, there is no need in this case to
increase the size of the coefficient matrix in Eq. 43, which has to be solved at each time step.
One can simply solve for the ANCF accelerations ë using the equation ë = M−1 ee Qe . There-
fore, once the forces Qe are computed, the solution for the ANCF accelerations involves
only vector–matrix multiplication, and the size of the coefficient matrix in Eq. 43 can be
significantly reduced.

Pantograph/Catenary Contact Forces It is clear that the coefficient matrix in Eq. 43


is a sparse matrix. Therefore, Eq. 43 can be efficiently solved using sparse-matrix techniques
to determine the second time derivatives of the coordinates qr , qf , e, and s as well as the
vector of Lagrange multipliers 𝝀. The vector of Lagrange multipliers can be used to define
the generalized constraint force as Fc = −CTq 𝛌. When the pantograph/catenary constraint
contact formulation is used, Lagrange multipliers can be used to determine the normal
contact forces as constraint forces. Using these normal contact forces, the tangential forces,
which can be defined using Coulomb’s dry friction law or any other friction or damping
model, can be defined and introduced to the dynamic formulation. To avoid using an iter-
ative procedure to determine contact forces, which depend on Lagrange multipliers when
the constraint contact formulation is used, Lagrange multipliers associated with these con-
straints can be stored from the previous time step and used to evaluate pantograph/catenary
contact forces at the current time step. This approach has been found accurate since in most
railroad vehicle system applications, a small time step is used in the numerical integration
of the system equations of motion due to the high speed of rotation of the wheelsets and
the high values of the contact forces. Furthermore, in the case of steady-state rolling and
under normal operating conditions, significant changes in contact forces are not expected.
320 Mathematical Foundation of Railroad Vehicle Systems

By using the values of Lagrange multipliers from the previous time step, the iterative solu-
tion of Eq. 43 can be avoided when the constraint contact formulation is used.
If the elastic contact formulation is used, on the other hand, pantograph/catenary con-
tact forces can be determined using a compliant force model with assumed stiffness and
damping coefficients. In this case, contact is not described using constraints, and conse-
quently, no degrees of freedom are eliminated. When the elastic contact formulation is used
to model pantograph/catenary interaction, the pan-head may lose contact with the catenary
wire, and therefore, a unilateral force model must be used to allow for pantograph/catenary
separation (Kulkarni et al. 2017). In this section, the development of this unilateral contact
force model is demonstrated in the case of longitudinal sliding. A similar procedure can be
used when both longitudinal and lateral sliding are considered.
In the case of longitudinal sliding only, the lateral and normal distances defined
T T ( )
by Eq. 39, dl = jc (rp − rc ) and dn = kc rP − rc , can be used to check whether the
pan-head and catenary contact wire come into contact. In these equations, jc and kc
are the lateral and normal axes of the contact frame, and rp and rc are, respectively,
the global position vectors of the potential contact points on the pan-head and catenary
wire. If pan-head/catenary penetration occurs, one can define the compliant force vector
[ ]T
Fpc = 0 Flpc Fnpc , where Fl = kl dl + cl ḋ l and Fn = kn dn + cn ḋ n are, respectively, the
pc pc

lateral and normal components of the compliant force model, kl and kn are assumed
stiffness coefficients, and cl and cn are assumed damping coefficients. The stiffness and
damping coefficients used in this model can be constant, linear, or nonlinear functions of
the pantograph/catenary coordinates. They can also be defined numerically using spline
functions based on tabulated data obtained from force measurements. The virtual work of
contact force Fpc can then be written as
( pc ( ) pc ( ) )
𝛿W pc = − Fl dl , ḋ l 𝛿dl + Fn dn , ḋ n 𝛿dn (7.44)
T T ( )
Using the distance equations dl = jc (rp − rc ) and dn = kc rP − rc , the virtual change in
these distances can be written as 𝛿dk = ((𝜕dk /𝜕qp )𝛿qp + (𝜕dk /𝜕ec )𝛿ec ), k = l, n. Therefore,
the virtual work of the contact forces can be written as
( ) ( )
pc ( ̇
) 𝜕dl p 𝜕dl c pc ( ̇
) 𝜕dn p 𝜕dn c
𝛿W = −Fl dl , dl
pc
𝛿q + c 𝛿e − Fn dn , dn 𝛿q + c 𝛿e (7.45)
𝜕qp 𝜕e 𝜕qp 𝜕e
Using this equation, the generalized contact forces associated with the pan-head and cate-
nary generalized coordinates can be written as
( ( ) ( )) ⎫
pc ( ) 𝜕dl pc ( ) 𝜕dn
Qp = − Fl dl , ḋ l + F n d , d ̇ ,⎪
𝜕qp n n
𝜕qp ⎪
( ( ) ( )) ⎬ (7.46)
pc ( ̇
) 𝜕dl pc ( ̇
) 𝜕dn ⎪
Q = − Fl dl , dl
c
+ Fn dn , dn ⎪
𝜕ec 𝜕ec ⎭
These unilateral generalized forces, which are assumed to be zero in the case of panto-
graph/catenary separation, can be introduced to the right-hand side of the equations of
motion of Eq. 43 in generalized force vector Qe .
Once the normal force is defined using the constraint or elastic contact formulation, the
tangential forces that oppose relative sliding can be introduced. In the case of dry friction,
pc
for example, tangential friction forces are a function of the resultant normal force Fnr and
Pantograph/Catenary Systems 321

the coefficient of friction 𝜇.


√The resultant normal force can be defined in the case of lon-
pc ( pc )2 ( pc )2
gitudinal sliding as Fnr = Fl + Fn . Using the Coulomb dry friction model, the
pc pc
tangential friction force can be defined as Ft = −𝜇Fnr ic (Kulkarni et al. 2017). A similar
procedure for developing the normal and tangential forces can be developed in the case of
relative longitudinal and lateral sliding. In this case, the normal force has only one compo-
nent since sliding is allowed in two different directions.
It is important to point out that pantograph/catenary contact forces are very small com-
pared to wheel/rail contact forces. Therefore, pantograph/catenary contact force, which is
necessary to avoid interruption of the electric power supply, normally does not have a sig-
nificant effect on vehicle dynamics during steady-state motion. The expected static value
of this force for high-speed trains ranges between 40 and 120 N (Hsiao 2010). Increasing
the pantograph uplift force can contribute to a significant increase in the dynamic value
of pantograph/catenary contact forces, as demonstrated in the literature (Pappalardo et al.
2016). Controlling the values of pantograph/catenary forces, however, is very important
despite their small magnitudes to ensure the continuous, uninterrupted electric power sup-
ply required for train operation.

7.9 PANTOGRAPH/CATENARY CONTACT FORCE


CONTROL

Providing the continuous supply of electric power necessary for the smooth operation of
high-speed trains requires maintaining pantograph/catenary contact and ensuring system
stability. This is a challenging problem due to the disturbances that can result from aero-
dynamics forces, rail car vibrations, and track irregularities. These disturbances can have
an adverse effect on the current-collection system as the result of variation of the contact
force between the pantograph strip and the catenary wire. For this reason, investigations
have been conducted to study the effectiveness of control strategies that can be adopted to
ensure the stability of pantograph/catenary contact and smooth train operation, particu-
larly at high speeds. The design of effective control systems requires accurate modeling of
train dynamics (Poetsch et al. 1997). The stability of the current-collection system depends
on the dynamic behavior of the pantograph/catenary system and its response to distur-
bances. Pan-head/catenary contact is normally maintained using an uplift force exerted
by actuators on the pantograph lower arm. The magnitude of the uplift force should not be
very high to avoid high contact force between the pantograph strips and catenary wire that
can lead to increased wear as a result of high friction forces. Low uplift forces, on the other
hand, can lead to loss of contact, which increases the probability of electrical arcing. Arc-
ing, which is the result of electric current flow in an air gap between the catenary contact
wire and the pantograph contact strip, leads to electric sparks with intense light that can
damage the contact wire and contact strip (Hsiao 2010). One solution to this problem is to
use an active control system by placing an actuator between the pan-head and the plunger of
the pantograph with the goal of reducing contact force variations (Pappalardo et al. 2016).
Pappalardo et al. (2016) presented a review of pantograph/catenary force control
methods and proposed a contact force control strategy based on a continuum-based
322 Mathematical Foundation of Railroad Vehicle Systems

pantograph/catenary dynamic model that employs ANCF finite elements. This dynamic
model, developed using the ANCF gradient-deficient cable elements, was used to test a
method to control pantograph/catenary contact force variations. Because only one ANCF
gradient vector is used in the formulation of pantograph/catenary contact conditions,
as described in this chapter, the proposed control strategy can also be used with fully
parameterized ANCF beam elements. A three-dimensional MBS model of a pantograph
mounted on a train was developed using the nonlinear augmented MBS formulation
defined by Eq. 43. Contact between the pantograph and catenary system was ensured
using the constraint contact formulation, while an elastic contact formulation was used to
predict wheel/rail contact forces. The standard deviation of the contact force was reduced
without affecting its mean value. This was accomplished by using a control actuator placed
between the pan-head and the plunger. Three types of control laws for the control action
were proposed to improve contact quality in the transient and steady-state phases. The first
control law is based on a feedback system, while the second and third control strategies are
based on feedback plus feed-forward systems. To examine the effectiveness of the proposed
control methods, Pappalardo et al. (2016) presented numerical results of train simulations
with and without controllers. The results presented were used to evaluate the proposed
new dynamic damping control method used with the continuum-based ANCF catenary
model. Using this new formulation, three different control strategies were designed to
control pantograph/catenary contact forces. These three strategies – a derivative controller,
a derivative/bang-bang controller, and a derivative/exponential controller – were designed
based on optimal control theory and the theory of virtual passive control. It was found that
all three proposed controllers led to an improvement in the standard deviation of the con-
tact force by more than 35% when compared to the uncontrolled system. The robustness of
the control strategies was tested by using the controllers with a high uplift force. To address
the practicality of the proposed control systems, their physical implementation using
electro-rheological or magneto-rheological devices or general actuators was discussed.

7.10 AERODYNAMIC FORCES


Environmental conditions such as temperature and wind can have a significant impact on
pantograph/catenary interaction forces. For example, high temperatures can alter the static
equilibrium position and pretension in catenary wires, while cold weather conditions can
result in the formation of ice on the wires, leading to undesirable deformations and poor
contact quality. Aerodynamic forces, on the other hand, can result in severe vibrations of
catenary cables as well as undesirable forces acting on the components of the pantograph
that negatively influence its functional operation. Aerodynamic drag and lift forces can
cause variations in contact forces, wear, and loss of contact. Wear can generate asymmetric
drag and lift forces, leading to galloping catenary motion (Stickland Scanlon 2001; Stick-
land et al. 2003), while high cross-wind loads can also cause severe vehicle vibrations that
directly influence pantograph/catenary interaction (Cheli et al. 2010). Aerodynamic drag
and lift force components alter the uplift force exerted on the pantograph mechanism.
For pantographs with multiple pan-heads, aerodynamic forces can cause an imbalance,
resulting in faster wear of one of the collector strips (Carnevale et al. 2016; Pombo et al.
2009). Furthermore, boundary layer turbulence near the car body roof due to vortex
Pantograph/Catenary Systems 323

shedding can excite the pantograph components, generate high frequencies, and increase
the sparking level; this, in turn, can negatively impact the quality of current collection
(Ikeda and Mitsumoji 2009). For these reasons, evaluating the effect of aerodynamic forces
on pantograph/catenary dynamics is an important design consideration.
As pointed out by Kulkarni et al. (2017), in general, two main approaches can be used to
evaluate the effect of the aerodynamic forces on pantograph/catenary systems. In the first
approach, computational fluid dynamics (CFD) is used to calculate aerodynamic forces on
pantograph system components and examine the contribution of these forces to the total
uplift pantograph force (Carnevale et al. 2015). In the second approach, the drag and lift
coefficients are obtained from experimental studies, and the aerodynamic forces are com-
puted using the equation F = (0.5 𝜌Ca |vwp |2 )ia , where F is the drag or lift force, 𝜌 is the
density of the fluid, Ca is the drag or lift pressure coefficient, vwp is the relative velocity
between the wind and pantograph component, and ia is the unit vector in the direction of
the drag or lift force (Pombo et al. 2009).

Aerodynamic Forces and ANCF Catenary Models Kulkarni et al. (2017) proposed
an aerodynamic force formulation that can be used with the ANCF catenary models. It was
shown that a simple aerodynamic force model could be integrated with the ANCF catenary
and rigid body pantograph models to account for the cross-wind effect. The approximate
drag and lift coefficients for pantograph components were obtained from previous investi-
gations reported in the literature (Carnevale et al. 2016). To compute the aerodynamic forces
on each body, a mesh data file, which defines the distribution of points at which aerody-
namic forces are applied, is created for each rigid or flexible body in the model. Using this
mesh data file, the effect of the moments of the aerodynamic forces can still be taken into
consideration. The mesh data file contains 13 parameters for each data point selected: 3
Cartesian position coordinates that define the location of the point with respect to the body
coordinate system, 9 direction cosines defining the orientation of a coordinate system for
the data point with respect to the body coordinate system, and the area assigned to the data
point. In the case of flexible catenaries, nodal locations are used to define the aerodynamic
data points, and aerodynamic force is applied at the nodes of the flexible catenary wires. The
aerodynamic drag and lift forces applied on a body i in the system can be defined as follows
( ( ) | |2 ) ( ( ) | |2 )
1 1
Fid = 𝜌 Cd A |vwi | id , Fil = 𝜌 Cl A |vwi | il (7.47)
2 | | 2 | |
where Fid and Fil are the drag and lift forces, respectively; 𝜌 is the fluid density; Cd and Cl are
the drag and the lift pressure coefficients per unit area, respectively; A is the area on which
the aerodynamic force is applied; vwi is the vector of the relative velocity between the body
and the wind; and id and il are unit vectors in the direction of the drag and the lift forces,
respectively. The generalized forces associated with the generalized coordinates of body i as
a result of aerodynamic drag force Fid and lift force Fil at point P can be defined, respectively,
∑np iT i ∑np iT i
as Qid = i=1 JP FdP and Qil = i=1 JP FlP , where np is the total number of points at which
aerodynamic forces are applied and JiP is the Jacobian matrix of the position field defined
by differentiation with respect to the vector of generalized coordinates of the body. In the
[ ]
case of a rigid body, one has JiP = 𝜕riP ∕𝜕qi = I −AiP ũ iP Gi ; and in the case of an ANCF
body JiP = 𝜕riP ∕𝜕ei = SiP , where riP is the absolute position vector of an arbitrary point P
on the body, qi is the vector of generalized coordinates of body i, AiP is the transformation
324 Mathematical Foundation of Railroad Vehicle Systems

matrix that defines the orientation of a point coordinate system on the body in the global
coordinate system, ũ iP is the skew-symmetric matrix associated with vector u iP that defines
the location of the point with respect to the coordinate system of the body on which the
point is defined, and Gi is the matrix that relates angular velocity vector 𝛚i defined in the
body coordinate system to the time derivatives of the orientation coordinates 𝛉̇ i : that is,
𝛚i = Gi 𝛉̇ i , and SiP is the ANCF shape function matrix defined at point P. By defining the
aerodynamic forces at points on the bodies, the generalized aerodynamic forces associated
with the generalized coordinates can be calculated, as previously discussed in this book. A
more detailed discussion of the formulation of the aerodynamic forces of rigid and flexible
bodies was presented by Kulkarni et al. (2017).
For design purposes, the National Electrical Safety Code (NESC) provides a formula for
evaluating the wind velocity pressure pv acting on an overhead line system. The NESC wind
velocity pressure formula is defined as pv = c𝜌a cve cg ci cf Ap (vw )2 N/m2 , where c𝜌a = 0.613 is a
velocity pressure numerical coefficient that accounts for the air mass density in the standard
atmosphere, cve is the velocity pressure exposure coefficient, cg is the gust response factor,
ci is the importance factor (assumed to be 1 in a pantograph/catenary system), cf is the
force coefficient shape factor, Ap is the projected wind area, and vw is the basic wind speed.
The catenary system should be designed such that contact wire displacement caused by
cross-wind, with respect to the track centerline, should not exceed 0.4 m to avoid contact
problems with the pantograph (Hsiao 2010).

7.11 PANTOGRAPH/CATENARY WEAR

The development of a general computational MBS algorithm for predicting panto-


graph/catenary wear allows incorporating the effects of vehicle vibration, wheel/rail
contact forces, and track irregularities. Such an algorithm also allows for predicting the
wear rate in the case of different motion scenarios that require using nonlinear models,
including curve negotiations, accelerations, and braking (Daocharoenporn et al. 2019). In
these motion scenarios, the effects of contact forces resulting from vehicle dynamics on
the wear rates of the catenary wire can be significant. Severe environmental and operating
conditions may cause arcing, high wear rates, or even failure of the pantograph/catenary
system to provide the electric power necessary for train operation. Arcing, for example, may
significantly increase the wear rate of the pantograph contact strip, leading to variations in
contact forces and deterioration of the pantograph/catenary system performance.
Continuous localized contact between the pan-head and catenary wire can cause signifi-
cant wear to the carbon strip inserted on the top of the pantograph and used for electric cur-
rent collection. Between two overhead line supports, the contact wire segment is straight;
therefore, during curve negotiations, the contact wire sweeps laterally over the entire car-
bon strip, resulting in uniform wear. When the vehicle negotiates a tangent track, it is also
desirable to obtain such a pattern of contact in which the contact wire sweeps laterally over
the surface of the carbon strip to achieve uniform wear instead of localized, more severe
wear. Therefore, in the case of a tangent track, the contact catenary wire is slightly zigzagged
around the centerline of the track to produce a lateral sweep that leads to uniform wear.
Pantograph/Catenary Systems 325

Several important factors can have a direct effect on pantograph/catenary wear, including
the design of the contact wire, the pretension in the catenary cable, and the uplift force of
the pantograph mechanism. For example, using a high uplift force can lead to a significant
increase in the magnitude of the contact force, and this, in turn, can lead to an increase
in the wear rate. Therefore, the wear rate can be reduced by controlling this uplift force of
the pantograph mechanism. The wear rate can also be reduced by properly designing the
pantograph/catenary system to better handle aerodynamic forces, staggering the contact
wire, controlling the intensity of the collection current, using the proper materials for the
contact wire and contact strip to reduce friction, and properly adjusting the pretension of
the catenary cables (Bucca and Collina 2009; Bucca and Collina 2015; Daocharoenporn et al.
2019). A wear model that accounts for electric and mechanical effects such as the electric
arcing effect, the Joule effect of electrical current, and the effect of contact friction forces
can be developed (Bucca and Collina 2015). This section presents a brief summary of the
investigation conducted by Daocharoenporn et al. (2019).

Brief Literature Review Studies of pantograph/catenary systems have shown that


electromechanical phenomena have a significant effect on contact between the pan-head
and catenary wire. He et al. (1998) used several contact strips with different material
properties to estimate the friction coefficients. It was found that the value of the friction
coefficients is highest in the case of unlubricated copper-to-copper contact. The optimal
value of the friction coefficient was found to be in the range of 0.2–0.24 in the case of
carbon-to-copper contact. He et al. (1998) also concluded that plastic deformation of soft
asperities is the main factor that causes wear of the contact catenary wires in the case of
metal-to-metal sliding. Kubo and Kato (1998, 1999), who investigated the effect of electrical
discharge or arcing on the wear rates of copper-impregnated contact strips, concluded
that the wear rate is proportional to the electrical discharge energy. One of the main
conclusions drawn from the study performed by Kubo and Kato (1998, 1999) is the adverse
effect of heat energy resulting from arcing. This generated heat energy is the cause of
several temperature-related effects such as oxidation of carbon, separation of impregnated
copper particles, melting, and evaporation. Temperatures during pantograph/catenary
system operations can reach a very high level (200–550 ∘ C) at which using grease becomes
less effective. The arc discharge process was further investigated by Chen et al. (2013),
who concluded that arcing has a very adverse effect on current-collection equipment. It
was also found that the pan-head wear rate is approximately proportional to the discharge
energy on a logarithmic scale. A similar conclusion was reported in the investigation
performed by Kubota et al. (2013) for copper alloy–impregnated, carbon fiber–reinforced
carbon composite collector strips. Ding et al. (2011) used a pure carbon contact strip and
a copper contact wire in their investigation. They concluded that the friction coefficient
is significantly higher if the current does not flow in the contact wire, and the friction
coefficient decreases as the electrical current and normal force increase. Ding et al. (2011)
found that the effect of sliding speed on the friction coefficient is insignificant, justifying
the use of the dry friction tangential force model, which does not depend on the magnitude
of the relative velocity. As pointed out by Daocharoenporn et al. (2019), a literature review
revealed that a large number of investigations and research efforts have been devoted to
understanding the wear mechanism of railroad vehicle current-collection systems. Some
326 Mathematical Foundation of Railroad Vehicle Systems

of these investigations were devoted to optimizing current-collection system design, which


is influenced by several interdependent parameters.

Wear Models The wear formulation developed by Bucca and Collina (2015) was adopted
by Daocharoenporn et al. (2019) to study the wear characteristics of a copper contact wire in
a curved-track motion scenario. Daocharoenporn et al. (2019) integrated this wear model
in a MBS computational algorithm that allowed for developing a detailed vehicle model
based on the approaches previously discussed in this book. In the MBS railroad vehicle
model developed by Daocharoenporn et al. (2019), the catenary wires were modeled using
ANCF finite elements, and the pantograph system was modeled as a mechanism consist-
ing of interconnected links. The model also included a detailed vehicle model with two
bogies; each bogie had two wheelsets. A three-dimensional wheel/rail contact formulation
was employed in the detailed MBS model used to study wear.
As previously mentioned, the tribological conditions generated during panto-
graph/catenary contact are the result of electrical and mechanical effects. It has been
shown in the literature that the wear mechanism in a pantograph/catenary system is
mainly governed by the following three factors: a mechanical factor resulting from sliding
friction, a Joule effect resulting from the flow of high-intensity current in catenary wires,
and electrical arcs resulting from pan-head/catenary separations. Because these factors do
not contribute to wear independently, the development of a robust, credible wear model
that can be used in computer simulations may require experimental testing. In such
experimental testing, the relationship between the effects of these wear factors as well as
their separate contributions can be evaluated by varying, for example, the materials of the
pan-head and contact wire, the magnitude of the uplift force, the relative sliding velocity,
and the electric current intensity (Bucca and Collina 2015). The normal wear rate N wr ,
defined as the volume worn out (measured in mm3 ) per unit distance (measured in km)
was computed in the simulation scenarios performed by Daocharoenporn et al. (2019). In
these simulation scenarios, a pure copper contact wire and a Kasperowski type of contact
strip were used. The Kasperowski contact strip has the carbon material part encased in
a copper casing from three sides. The normal wear rate N wr of the copper contact line is
given by (Bucca and Collina 2015)
( ( ))−𝛼 ( )𝛽 ( )
1 Ic Fm Fm Rc Fm •Ic2 V •I
Nwr = k1 1+ + k2 (1 − u) + k3 u a c (7.48)
2 I0 F0 H H •V V •Hm •𝜌
The parameters in this equation are defined in Table 2, which also shows the numerical
values used by Daocharoenporn et al. (2019). Using this equation, the effect of individ-
ual parameters on the normal wear rate N wr can be evaluated separately by varying these
parameters and keeping other parameters constant. Such a parametric study allows for a
better understanding of the wear mechanism. The electrical contribution to the wear for
a given current value is proportional to electrical contact resistance Rc , which depends
on mean contact force F m . Bucca et al. (2011) gave an explicit expression for this resis-
( )
tance based on extensive experimentation. This expression is defined as Rc = Rc Fm =
0.013 + 0.09e−(Fm −14)∕11 .
By examining Eq. 48, it can be shown that the contact wire normal wear rate N wr decreases
as electric current increases at high values of the mean contact force. An opposite trend
Pantograph/Catenary Systems 327

Table 7.2 Catenary wear parameters.

Symbol Description Value

pc
Fm Mean value of contact force Fn (N) —
k1 Weight of the mechanical contribution to the N wr value 22.4
k2 Weight of the electrical contribution to the N wr value 10.3
k3 Weight of the contribution due to electrical arcs to the N wr 0.4
value
𝛼 Coefficient of the dependence of the mechanical 4.5
contribution on the electrical current
𝛽 Coefficient of nonlinear dependence of the mechanical 1.8
contribution on the mean value of the contact force
I0 Reference value of electrical current (A) 500
Ic Nominal electrical current during tests (A) —
F0 Reference value of contact force (N) 90
H Hardness of material (N/mm2 ) 700
Rc Electrical contact resistance between strip and wire (Ω) —
V Sliding speed during tests (m/s) —
u Decimal fraction value of the percentage of contact loss —
Va Electrical arc voltage (V) 50
Hm Latent heat of fusion for copper (kJ/kg) 205
𝜌 Density of copper (kg/m3 ) 8940

Source: Courtesy of Daocharoenporn, S., Mongkolwongrojn, M., Kulkarni, S., and Shabana, A.A.

Table 7.3 Pantograph/catenary interaction criteria.

Pantograph parameter Acceptable values

Mean contact force F m = 0.00097(V)2 + 70N


Standard deviation 𝜎 max < 0.3F m
Maximum contact force F max < 350N
Maximum catenary wire dup ≤ 120 mm
uplift at steady arm
Maximum pantograph △z ≤ 80 mm
vertical amplitude
Percentage of real arcing NQ ≤ 0.2%
328 Mathematical Foundation of Railroad Vehicle Systems

occurs if the value of the mean contact force is below 90 N. For electric current values
above 300 A, wear rates are lower as the contact force increases; this was attributed to the
mechanism of current lubrication (Bucca and Collina 2009). On the other hand, for lower
values of electric current, the wear rate increases as the mean contact force increases. This
type of wear is characterized by an abrasive wear mechanism and dry friction because
the temperatures in the contact region are not high enough to melt the metals in con-
tact. The contribution of arcs is proportional to the power generated by the arcs, which
depends on the electric voltage, the speed of the train, and the gap between the pan-head
and contact wire. These findings demonstrate the interdependency of the electrical and
mechanical contributions to the complex wear mechanism of the catenary contact wire
(Daocharoenporn et al. 2019). The EN50367 and EN50317 standards provide criteria for
electric current-collection systems of railroad vehicles. The limits shown in Table 3 were
used by Daocharoenporn et al. to validate the results obtained using their computer model.
329

Appendix

CONTACT EQUATIONS AND ELLIPTICAL INTEGRALS

Chapter 5 discussed Hertz contact theory, which is widely used in the analysis of wheel/rail
contact. For a given normal load and material properties, this theory can be used to deter-
mine the dimensions of the contact ellipse as a function of the wheel and rail surface
geometries. Hertz theory, which assumes that the dimensions of the contact area are
small in comparison with the dimensions of the two solids in contact, is based on the
elastic half-space assumptions used to determine stress distribution in solids (Boussinesq
1885; Cerruti 1882; Love 1944). The development of Hertz theory requires the evaluation
of the elliptical integrals whose values depend on the dimensions of the contact ellipse.
Section A.1 presents the derivation of some of the equations required for the development
of Hertz theory, and Section A.2 is devoted to elliptical integrals. The equations presented
in this appendix are a summary of the equations presented in the two appendices of a
previous book (Shabana et al. 2008).

A.1 DERIVATION OF THE CONTACT EQUATIONS


In the elastic half-space, shown in Figure A.1, point P is assumed to lie inside the contact
area Ac resulting from load applications. By choosing the proper coordinate system and
without any loss of generality, the coordinates of point P can be assumed to be [xP yP 0]T . If Q
[ ]T
is another arbitrary point inside the solid with coordinates x y z , as shown in Figure A.1,
distance lPQ between points P and Q can be written as

( )2 ( )2
lPQ = xP − x + yP − y + z2 (A.1)
The distributions of normal pressure p and tangential tractions 𝜏 x and 𝜏 y in contact area Ac
are assumed to be functions of xP and yP . Using these assumptions, the following potential
functions can be defined
( ) ⎫
Φx = ∫A 𝜏x xP , yP hP dAc ⎪
c
( ) ⎪
Φy = ∫A 𝜏y xP , yP hP dAc ⎬ (A.2)

c
( )
Φp = ∫A p xP , yP hP dAc ⎪
c ⎭

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
330 Mathematical Foundation of Railroad Vehicle Systems

P(xp,yp,0)
lPQ X

Q(x,y,z)
Z

Figure A.1 Elastic half-space.

where hP = z ln(lPQ + z) − lPQ and dAc = dxP dyP . It follows that

( ) ( ) ⎫
𝜕Φx ⎪
Φx = = 𝜏x xP , yP ln lPQ + z dAc
𝜕z ∫Ac ⎪

𝜕Φy ( ) ( ) ⎪
Φy = = 𝜏y xP , yP ln lPQ + z dAc ⎬ (A.3)
𝜕z ∫Ac ⎪
𝜕Φp ( ) ( ) ⎪
Φp = = p xP , yP ln lPQ + z dAc ⎪
𝜕z ∫Ac ⎪

Using the definitions of the preceding equations, it can be shown that the elastic displace-
ment of an arbitrary point Q can be written as (Love 1944)
[ ]
1 𝜕Φ 𝜕Φz 𝜕Φ 𝜕Φ ⎫
ux = 2 x − + 2𝜈 s − z s ⎪
4𝜋G 𝜕z 𝜕x 𝜕x 𝜕x
[ ] ⎪
1 𝜕Φy 𝜕Φz 𝜕Φ 𝜕Φ ⎪
uy = 2 − + 2𝜈 s − z s ⎬ (A.4)
4𝜋G 𝜕z 𝜕y 𝜕y 𝜕y ⎪
[ ] ⎪
1 𝜕Φ 𝜕Φ
uz = 2 z + (1 − 2𝜈) Φs − z s ⎪
4𝜋G 𝜕z 𝜕z ⎭
( )
where G and 𝜈 are, respectively, the modulus of rigidity and Poisson ratio; Φs = 𝜕Φx ∕𝜕x +
( ) ( )
𝜕Φy ∕𝜕y + 𝜕Φp ∕𝜕z ; and Φs = (𝜕Φx /𝜕x) + (𝜕Φy /𝜕y) + (𝜕Φp /𝜕z) is the Boussinesq function.
If the effect of friction is neglected and the pressure p(xP , yP ) is assumed to be normal to the
contact surface, one has Φx = 0, Φx = 0, Φy = 0 and Φy = 0. Using these assumptions in
Eqs. 2–4, one has
( ) ( )
Φs = Φp = ∫A p xP , yP ln lPQ + z dAc ⎫
c
( ) ⎪
( ) 1 ⎬ (A.5)
Φs = ∫A p xP , yP dAc ⎪
c lPQ ⎭
Appendix: Contact Equations and Elliptical Integrals 331

and
[ ]
1 𝜕Φs 𝜕Φ ⎫
ux = − (1 − 2𝜈) +z s ⎪
4𝜋G 𝜕x 𝜕x
[ ] ⎪
1 𝜕Φs 𝜕Φs ⎪
uy = − (1 − 2𝜈) +z ⎬ (A.6)
4𝜋G 𝜕y 𝜕y ⎪
[ ] ⎪
1 𝜕Φs
uz = 2 (1 − 𝜈) Φs − z ⎪
4𝜋G 𝜕z ⎭

For an arbitrary point on the surface that has a z-coordinate equal to zero – that is, z = 0 – the
preceding equation leads to
( )| ⎫
1 − 2𝜈 𝜕Φs|
ux ||z=0 = − | ⎪
4𝜋G 𝜕x| ⎪
|z=0
( ) ⎪
| 1 − 2𝜈 𝜕Φs || ⎪
uy | = − | ⎬ (A.7)
|z=0 4𝜋G 𝜕y ||z=0 ⎪
( ) ⎪
1 − 𝜈 𝜕Φs || ⎪
uz ||z=0 = |
2𝜋G 𝜕z ||z=0 ⎪

In the case of an elliptical contact area with semi axes that have dimensions a and b, the
applied normal pressure can be written as
[ ( )2 ( )2 ]n
x y
p (x, y) = po 1 − + (A.8)
a b

where n is a given number and the area of contact can be defined by the ellipse equation
(x/a)2 + (y/b)2 = 1. Using potential theory, the Boussinesq function (Eq. A.5) for a general
point in the solid is given by
( )
( 2 2 2
) n+ 12
(x) (y) (z)
∞ 1− − −
Γ (n + 1) Γ (1∕2) (a)2 + w (b)2 + w w
Φs (x, y, z) = p0 ab √( )( 2 ) dw
Γ (n + 3∕2) ∫𝛾1 2
(a) + w (b) + w w
(A.9)

In this equation, Γ is the gamma function and 𝛾 1 is the positive root of the equation
((x)2 /((a)2 + 𝛾)) + ((y)2 /((b)2 + 𝛾)) + (z)2 /𝛾 = 1.
In Hertz theory, n = 1/2. In this special case, Φs of Eq. 9 reduces to
( )
(x)2 (y)2 (z)2
∞ 1− − −
1 (a)2 + w (b)2 + w w
Φs (x, y, z) = 𝜋p0 ab √( ) ( ) dw (A.10)
2 ∫𝛾1
(a)2 + w (b)2 + w w
332 Mathematical Foundation of Railroad Vehicle Systems

For an arbitrary point on the surface, z = 0; and the preceding equation reduces to
( )
(x)2 (y)2
∞ 1 − −
1 (a)2 + w (b)2 + w
Φs (x, y, z) = 𝜋p0 ab √( )( ) dw (A.11)
2 ∫𝛾1
(a)2 + w (b)2 + w w
Using this equation and Eq. 7, the z-displacement can be written as
1 − (𝜈)2 ( )
uz = Le − Me (x)2 − Ne (y)2 (A.12)
𝜋E
where E is Young’s modulus, and

𝜋p0 ab ∞ 𝜋p b [ ] ⎫
dw ⎪
Me = √( = 2 0 2 Ke − E e
2 ∫ )3 ( ) (e) (a) ⎪
0
(a)2 + w (b)2 + w w ⎪
[ ] ⎪
𝜋p0 ab ∞ dw 𝜋p b ( a )2 ⎪
Ne = √( = 20 2 E e − Ke
2 ∫0 )( )3 (e) (a) b ⎬ (A.13)
(a)2 + w (b)2 + w w ⎪

𝜋p0 ab ∞ dw ⎪
Le =
2 ∫0
√( )( ) = 𝜋p0 bK e ⎪
(a) + w (b)2 + w w
2



where Ee and K e are complete elliptical integrals of argument e = 1 − (a∕b)2 for b > a,
as will be explained in the following section. Because the pressure is assumed to have an
elliptical distribution (Eq. 8), the normal force F ns can be defined as F ns = 2p0 𝜋ab/3, as
discussed in Chapter 5.

A.2 ELLIPTICAL INTEGRALS


In Chapter 5, complete elliptical integrals were used in the development of the Hertz theory
of contact. These integrals are defined as
𝜋∕2 ⎫
𝜋∕2 cos2 w cos2 w ⎪
Be = ∫0 √ dw = √ dw,
1 − (e)2 sin2 w ∫0 cos2 w − (g)2 sin2 w ⎪
( )− 3

𝜋∕2
Ce = ∫0 sin2 wcos2 w 1 − (e)2 sin2 w 2 dw, ⎪

𝜋∕2 √ ⎬ (A.14)
𝜋∕2 sin2 w
De = ∫0 √ dw, Ee = 1 − (e) sin w dw, ⎪
2 2

1 − (e)2 sin2 w ∫0 ⎪

𝜋∕2 1 ⎪
Ke = ∫ 0 √ dw
1 − (e)2 sin2 w ⎪

Appendix: Contact Equations and Elliptical Integrals 333

Table A.1 Elliptical integrals.

g = b/a Be Ce De Ee Ke (e)2

0 1.0 −2 + ln(4/g) −1 + ln(4/g) 1.0 + ln(4/g) 1.00


0.1 0.9889 1.7351 2.7067 1.0160 3.6956 0.99
0.2 0.9686 1.1239 2.0475 1.0505 3.0161 0.96
0.3 0.9451 0.8107 1.6827 1.0965 2.6278 0.91
0.4 0.9205 0.6171 1.4388 1.1507 2.3593 0.84
0.5 0.8959 0.4863 1.2606 1.2111 2.1565 0.75
0.6 0.8719 0.3929 1.1234 1.2763 1.9953 0.64
0.7 0.8488 0.3235 1.0138 1.3456 1.8626 0.51
0.8 0.8267 0.27060 0.9241 1.4181 1.7508 0.36
0.9 0.8055 0.22925 0.8491 1.4933 1.6545 0.19
1.0 𝜋/4 = 0.7864 𝜋/16 = 0.19635 𝜋/4 = 0.7864 𝜋/2 = 1.5708 𝜋/2 = 1.5708 0.00

Source: Courtesy of Kalker, J.J.


where e = 1 − (a∕b)2 , b > a and g = a/b. The elliptical integrals presented in the preced-
ing equations are not totally independent because they are related by the equations
( ) ⎫
Ke = 2De − (e)2 Ce , Ee = 2 − (e)2 De − (e)2 Ce , Be = De − (e)2 Ce , ⎪
( ) ( ) ⎬ (A.15)
De = Ke − Ce ∕(e)2 , Be = Ke − De , Ce = De − Be ∕(e)2 ⎪

The values of the elliptical integral as a function of the ratio of the contact ellipse semi axes
g = a/b are shown in Table A.1.
335

Bibliography

1. ACC. (2017). Competitive switching: making the switch to a more competitive,


healthy freight rail system. Freight Rail Reform. https://www.freightrailreform.com/
making-the-switch-to-a-more-competitive-healthy-freight-rail-system.
2. Andersson, C. and Abrahamsson, T. (2002). Simulation of interaction between a train
in general motion and a track. Vehicle System Dynamics 38: 433–455.
3. APTA Press Task Force. (2007). Standard for wheel flange angle for passenger
equipment. The American Public Transportation Association, Report # APTA
PR-M-S-015-06.
4. Arnold, M. (1996). Numerical problems in the dynamical simulation of wheel-rail
systems. Zeitschrift Fur Angewandte Mathematik Und Mechanik 76: 151–154.
5. Atkinson, K.E. (1978). An Introduction to Numerical Analysis. Wiley.
6. Bailey, J.R. and Wormley, D.N. (1992). A comparison of analytical and experimen-
tal performance data for a two-axle freight car. ASME Journal of Dynamic Systems,
Measurement, and Control 114: 141–147.
7. Baumgarte, J. (1972). Stabilization of constraints and integrals of motion. Computer
Methods in Applied Mechanics and Engineering 1: 1–16.
8. Bavinck, H. and Dieterman, H.H. (1996). Wave propagation in a finite cascade of
masses and springs representing a train. Vehicle System Dynamics 26: 45–60.
9. Berg, M. (1998). A non-linear rubber spring model for rail vehicle dynamics analysis.
Vehicle System Dynamics 30: 197–212.
10. Berghuvud, A. (2002). Freight car curving performance in braked conditions.
Proceedings of the Institution of Mechanical Engineers, Part F: Journal of Rail and
Rapid Transit 216: 23–29.
11. Berzeri, M., Sany, J.R., and Shabana, A.A. (2000). Curved track modeling using the
absolute nodal coordinate formulation. Technical report MBS00-4-UIC. Department of
Mechanical Engineering, University of Illinois at Chicago.
12. Bing, A.J., Shaun, R.B., and Henderson, B. (1996). Design data on suspension systems
of selected rail passenger cars. FRA report DOT/FRA/ORD-96-01.
13. Blader, F. and Klauser, P. (1989). User’s manual for NUCARS, version 1.0. Technical
report R-734. Association of American Railroads.
14. Blader, F.B. (1989). A review of literature and methodologies in the study of derail-
ments caused by excessive forces at the wheel/rail interface. AAR technical report
R-717. AAR Technical Center, Chicago.
Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,
First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
336 Bibliography

15. Bocciolone, M., Caprioli, A., Cigada, A., and Collina, A. (2007). A measurement
system for quick rail inspection and effective track maintenance strategy. Mechanical
Systems and Signal Processing 21: 1242–1254.
16. Bocciolone, M., Resta, F., Rocchi, D., and Collina, A. (2006a). Pantograph aerodynamic
effects on the pantograph–catenary interaction. Vehicle System Dynamics 44 (1):
560–570.
17. Bocciolone, M., Caprioli, A., Cigada, A., and Collina, A. (2007). A measurement
system for quick rail inspection and effective track maintenance strategy. Mechanical
Systems and Signal Processing 21: 1242–1254.
18. Bocciolone, M., Cheli, F., Corradi, R., and Collina, A. (2008a). Crosswind action on
rail vehicles: Wind tunnel experimental analyses. Journal of Wind Engineering and
Industrial Aerodynamics 96: 584–610.
19. Bocciolone, M., Resta, F., Rocchi, D., Tosi, A., and Collina, A. (2006b). Pantograph
aerodynamic effects on the pantograph–catenary interaction. Vehicle System Dynamics
44 (1): 560–570.
20. Bocciolone, M., Cheli, F., Corradi, R., Muggiasca, S., and Tomasini, G. (2008b). Cross-
wind action on rail vehicles: wind tunnel experimental analyses. Journal of Wind
Engineering and Industrial Aerodynamics 96 (5): 584–610.
21. Bonet, J. and Wood, R.D. (1997). Nonlinear Continuum Mechanics for Finite Element
Analysis. Cambridge University Press.
22. Boocock, D. (1969). The steady-state motion of railway vehicles on curved track.
Journal of Mechanical Engineering Science 2: 556–566.
23. Boresi, A.P. and Chong, K.P. (2000). Elasticity in Engineering Mechanics, 2e. Wiley.
24. Bosso, N., Gugliotta, A., and Zampieri, N. (2015). Strategies to simulate wheel-rail
adhesion in degraded conditions using a roller-rig. Vehicle System Dynamic 53 (5):
619–634.
25. Bosso, N. and Zampieri, N. (2013). Real-time implementation of a traction control
algorithm on a scaled roller rig. Vehicle System Dynamics 51 (4): 517–541.
26. Boussinesq, J. (1885). Application des Potentials ‵ al’étude de l’équilibre et du mouvement
des solides élastiques. Paris: Gauthier-Villars.
27. Bracciali, A., Cascini, G., and Ciuffi, R. (1998). Time domain model of the vertical
dynamics of a railway track up to 5 kHz. Vehicle System Dynamics 30: 1–15.
28. British Railway Board. (1996). Wheelset tread standards & gauging. MT/288 report,
issue 2, revision A.
29. Bruni, S., Facchinetti, A., Kolbe, M., and Massat, J-P., (2011). Hardware-in-the-loop
testing of pantograph for homologation. Proceedings 9th World Congress on Railway
Research, 22–26 May 2011. Lille, France, PB003802.
30. Bruni, S., Ambrosio, J., Carnicero, A., Cho, Y.H., Finner, L., Ikeda, M., and Zhang, W.
(2015). The results of the pantograph–catenary interaction benchmark. Vehicle System
Dynamics 53 (3): 412–435.
31. Bucca, G. and Collina, A. (2009). A procedure for the wear prediction of collector strip
and contact wire in pantograph–catenary system. Wear 266 (1): 46–59.
32. Bucca, G. and Collina, A. (2015). Electromechanical interaction between carbon-based
pantograph strip and copper contact wire: a heuristic wear model. Tribology Interna-
tional 92: 47–56.
Bibliography 337

33. Bucca, G., Collina, A., Manigrasso, R., Mapelli, F., and Tarsitano, D. (2011).
Analysis of electrical interferences related to the current collection quality in
pantograph–catenary interaction. Proceedings of the Institution of Mechanical
Engineers, Part F: Journal of Rail and Rapid Transit 225 (5): 483–500.
34. Business, Wire (2017). Axalta launches new high performance, protective industrial
rail car coatings. https://www.businesswire.com/news/home/20171018005982/en/
Axalta-Launches-New-High-Performance-Protective-Industrial.
35. Carnevale, M., Facchinetti, A., Maggiori, L., and Rocchi, D. (2016). Computational
fluid dynamics as a means of assessing the influence of aerodynamic forces on the
mean contact force acting on a pantograph. Proceedings of the Institution of Mechani-
cal Engineers, Part F: Journal of Rail and Rapid Transit 230 (7): 1698–1713.
36. Carter, F.W. (1926). On the action of a locomotive driving wheel. Proceedings of the
Royal Society of London Series A 112: 151–157.
37. Cerruti, V. (1882). Roma, Acc. Lincei, Mem. Fis. Mat.
38. Cheli, F., Ripamonti, F., Rocchi, D., and Tomasini, G. (2010). Aerodynamic behaviour
investigation of the new EMUV250 train to cross wind. Journal of Wind Engineering
and Industrial Aerodynamics 98 (4): 189–201.
39. Chen, G.X., Yang, H.J., Zhang, W.H., Wang, X., Zhang, S.D., and Zhou, Z.R. (2013).
Experimental study on arc ablation occurring in a contact strip rubbing against a
contact wire with electrical current. Tribology International 61: 88–94.
40. Cole, C. and Sun, Y.Q. (2006). Simulated comparisons of wagon coupler systems
in heavy haul trains. Proceedings of the Institution of Mechanical Engineers, Part F:
Journal of Rail and Rapid Transit 220: 247–256.
41. Cole, C., Spiryagin, M., and Sun, Y.Q. (2013). Assessing wagon stability in complex
train systems. International Journal of Rail Transportation 1 (4): 193–217.
42. Cole, C., Spiryagin, M., Wu, Q., and Sun, Y.Q. (2017). Modelling, simulation and appli-
cations of longitudinal train dynamics. Vehicle System Dynamics 55 (10): 1498–1571.
43. Collina, A. and Bruni, S. (2002). Numerical simulation of pantograph-overhead
equipment interaction. Vehicle System Dynamics 38 (4): 261–291.
44. Cooperrider, N.K. (1991). Ride quality assessment for a 6-axle locomotive and a heavy
truck. ASME Paper No. RTD-Vol.4: 153-160.
45. Cooperrider, N.K., Law, E.H., Hull, R., Kadala, P.S., and Tuten, P.S. (1975). Analytical
and experimental determination of nonlinear wheel/rail geometric constraints. Report
FRA-OR&D 76-244 (PB-252 290).
46. Cooperrider, N.K., Hedrick, J.K., Law, E.H., and Malstrom, C.W. (1976). The applica-
tion of quasi-linearization techniques to the prediction of nonlinear railway vehicle
response. In: Proceedings of the IUTAM Symposium, 314–325.
47. Coulomb, C.A. (1785). Theorie des Machines Simples. In: Memoire de Mathematique
et de Physique de l’Academie Royale, 161–342.
48. Cummings, S. (2018). A new wheel profile for North American freight railroads:
AAR-2A. Heavy Haul Seminar, May 2–3. http://www.wheel-rail-seminars.com/
archives/2018/hh-papers/presentations/HH03.pdf.
49. Daocharoenporn, S., Mongkolwongrojn, M., Kulkarni, S., and Shabana, A.A. (2019).
Prediction of the pantograph/catenary wear using nonlinear multibody system
dynamic algorithms. ASME Journal of Tribology, accepted for publication.
338 Bibliography

50. Datoussaid, S., Verlinden, O., Wenderloot, L., and Conti, C. (1998). Computer-aided
analysis of urban railway vehicles. Vehicle System Dynamics 30: 213–227.
51. De Pater, A.D. (1974). The propagation of waves in a semi-infinite chain of material
points and springs representing a long train. Vehicle System Dynamics 3: 123–140.
52. De Pater, A.D. (1988). The geometrical contact between track and wheel-set. Vehicle
System Dynamics 17: 127–140.
53. Diana, G., Fossati, F., and Resta, F. (1998). High speed railway: collecting pantographs
active control and overhead lines diagnostic solutions. Vehicle System Dynamics 30:
69–84.
54. Dietz, S., Netter, H., and Sachau, D. (1988). Fatigue life prediction of a railway bogie
under dynamic loads through simulation. Vehicle System Dynamics 29: 385–402.
55. Ding, T., Chen, G.X., Wang, X., Zhu, M.H., Zhang, W.H., and Zhou, W.X. (2011).
Friction and wear behavior of pure carbon strip sliding against copper contact wire
under AC passage at high speeds. Tribology International 44 (4): 437–444.
56. Diomin, Y.V. (1994). Stabilization of high-speed railway vehicles. Vehicle System
Dynamics 23: 107–114.
57. Do Carmo, M.P. (1976). Differential Geometry of Curves and Surfaces. Prentice Hall.
58. Duff, I.S., Erisman, A.M., and Reid, J.K. (1986). Direct Methods for Sparse Matrices.
Oxford: Clarendon Press.
59. Duffek, W. (1982). Contact geometry in wheel rail vehicles. In: Proceedings of Con-
tact Mechanics and Wear of Rail/Wheel Systems, 161–181. Vancouver: University of
WaterLoo.
60. Dukkipati, R.V. (1998). Dynamics of wheelset on roller rig. Vehicle System Dynamics
30 (6): 409–430.
61. Dukkipati, R.V. (2000). Vehicle Dynamics. New Delhi: CRC Press.
62. Dukkipati, R.V. and Amyot, J.R. (1988). Computer-Aided Simulation in Railway
Dynamics. New York: Mercel-Dekker.
63. Dukkipati, R.V. and Dong, R. (1999). Impact loads due to wheel flats and shells.
Vehicle System Dynamics 31: 1–22.
64. Dukkipati, R.V., Swamy, S.N., and Osman, M.O.M. (1991). Independently rotating
wheel systems for railway vehicle – a survey of the state of the art. The Archives of
Transport III: 297–330.
65. Durali, M. and Shadmehri, B. (2003). Nonlinear analysis of train derailment in severe
braking. ASME Journal of Dynamic Systems, Measurement, and Control 125: 48–53.
66. Eich-Soellner, E. and Fhuhrer, C. (1998). Numerical Methods in Multibody Dynamics.
Teubner, B. G. GmbH.
67. Elkins, J.A. (1992). Prediction of wheel rail interaction – the state-of-the-art. Vehicle
System Dynamics 20: 1–27.
68. Elkins, J.A. and Carter, A. (1993). Testing and analysis techniques for safety assess-
ment of rail vehicles: the state-of-the-art. Vehicle System Dynamics 22: 185–208.
69. Elkins, J.A. and Eikhoff, B.M. (1982). Advances in nonlinear wheel/rail force predic-
tion methods and their validation. ASME Journal of Dynamic Systems, Measurement,
and Control 104: 133–142.
70. Elkins, J.A. and Gostling, R.J. (1977). A general quasi-static curving theory for railway
vehicles. In: Proceedings of 2nd IUTAM Symposium, Vienna, 388–406.
Bibliography 339

71. Elkins, J. and Wu, H. (2000). New criteria for flange climb derailment. IEEE/ASME
Joint Rail Conference, Newark, New Jersey, April 4–6.
72. CENELEC. (2007). Railway applications. Supply voltages of traction systems. EN
50163.
73. Endlicher, K.O. and Lugner, P. (1990). Computer simulation of the dynamical curving
behavior of a railway bogie. Vehicle System Dynamics 19: 71–95.
74. Escolona, J.L. (2002). Elastic Contact Personal Communications.
75. Fisette, P. and Samin, C. (1991). Lateral dynamics of a light railway vehicle with inde-
pendent wheels. In: Proceedings of the 12th IAVSD Symposium on the Dynamics of
Vehicles on Road and Tracks, Lyon, 157–171.
76. Fries, R.H., Cooperrider, N.K., and Law, E.H. (1981). Experimental investigation of
freight car lateral dynamics. ASME Journal of Dynamic Systems, Measurement, and
Control 103: 201–210.
77. Garcia de Jalon, J. and Bayo, E. (1994). Kinematic and Dynamic Simulation of
Multibody Systems: The Real-Time Challenge. Springer-Verlag.
78. Garg, V.K. and Dukkipati, R.V. (1988). Dynamics of Railway Vehicle Systems.
New York: Academic Press.
79. Gere, J.M. and Weaver, W. (1965). Analysis of Framed Structures. New York: Van
Nostrand.
80. Gerstmayr, J. and Shabana, A.A. (2006). Analysis of thin beams and cables using the
absolute nodal co-ordinate formulation. Nonlinear Dynamics 45 (1–2): 109–130.
81. Gilchrist, A.O. (1998). The long road to solution of the railway hunting and curving
problems. Proceedings of the Institution of Mechanical Engineers, Part F: Journal of Rail
and Rapid Transit 212: 219–226.
82. Gilchrist, A.O. and Brickle, B.V. (1976). A re-examination of the proneness to derail-
ment of a railway wheel-set. Journal of Mechanical Engineering Science 18: 131–141.
83. Gioboata, D., Constantin, G., Abalaru, A., Stanciu, D., Logofatu, C., and Ghionea, I.
(2017). Method and system for measurementof railway wheels rolling surface in
re-shaping process. Proceedings in Manufacturing Systems 12: 113–118.
84. Gladwell, G.M.L. (1980). Contact Problem in the Classical Theory of Elasticity. Alphen
aan den Rijn: Sijthogg and Noordhof.
85. Goetz, A. (1970). Introduction to Differential Geometry. Addison Wesley.
86. Goldsmith, W. (1960). Impact – The Theory and Physical Behaviour of Colliding Solids.
London, UK: Edward Arnold LTD.
87. Grassie, S.L. (1993). Dynamic modeling of the track. In: Rail Quality and Mainte-
nance for Modern Railway Operation (eds. J.J. Kalker, D. Cannon and O. Orringer).
Dordrecht: Kluwer.
88. Greenwood, D.T. (1988). Principle of Dynamics, 2e. Prentice Hall.
89. Guida, D. and Pappalardo, C.M. (2014). Development of a closed-chain multibody
model of a high-speed railway pantograph for hybrid motion/force control of the pan-
tograph/catenary interaction. International Journal of Mechanical Engineering and
Industrial Design 3 (5): 45–85.
90. Hamid, A., Rasmussen, K., Baluja, M., and Yang, T-L. (1983). Analytical description of
track geometry variations. FRA report DOT/FRA/ORD-83/03.1.
340 Bibliography

91. Handoko, Y.A. (2006). Investigation of the dynamics of railway bogies subjected to
traction/braking torque. Ph.D. dissertation. Central Queensland University, Australia.
92. Handoko, Y. and Dhanasekar, M. (2006). An inertial reference frame method for the
simulation of the effect of longitudinal force to the dynamics of railway wheelsets.
Nonlinear Dynamics 45: 399–425.
93. Handoko, Y., Xia, F., and Dhanasekar, M. (2004). Effect of asymmetric brake shoe
force application on wagon curving performance. Vehicle System Dynamics Supplement
41: 113–122.
94. Haque, I., Law, E.H., and Cooperrider, N.K. (1979). User’s manual for program for
calculation of Kalker’s linear creep coefficients. FRA report FRA/ORD-78/71.
95. He, D.H., Manory, R.R., and Grady, N. (1998). Wear of railway contact wires against
current collector materials. Wear 215 (1–2): 146–155.
96. Hedrick, J.K. and Arslan, A.V. (1979). Nonlinear analysis of rail vehicle forced lateral
response and stability. ASME Journal of Dynamic Systems, Measurement, and Control
101: 230–237.
97. Heller, R. and Cooperrider, N.K. (1977). Users’ manual for asymmetric wheel/rail con-
tact characterization program. FRA report FRA/ORD-78/05.
98. Hertz, H. (1882). Über die berührung fester elastische Körper und über die Harte.
Verhandlungen des Vereins zur Beförderung des Gewerbefleisses, Leipzig.
99. Hesser, K. (1998). Progress in railway mechanical engineering: passenger and com-
muter rail vehicles and components. ASME International Mechanical Engineering
Congress & Exposition (Rail Transportation), Anaheim, California.
100. Hsiao, M. (2010). OCS requirements. TM 3.2.1, prepared by PB for the California High
Speed Rail Authority.
101. Huan, R.H., Pan, G.F., and Zhu, W.Q. (2012). Dynamics of pantograph–catenary
system considering local singularities of contact wire with critical wavelengths. In:
Proceedings of the 1st International Workshop on High-Speed and Intercity Railways,
319–333. Berlin, Heidelberg: Springer.
102. Huang, C., Zeng, J., Luo, G., and Shi, H. (2018). Numerical and experimental studies
on the car body flexible vibration reduction due to the effect of car body-mounted
equipment. Proceedings of the Institution of Mechanical Engineers, Part F: Journal of
Rail and Rapid Transit 232: 103–120.
103. Huston, R.L. (1990). Multibody Dynamics. Butterworth-Heinemann.
104. Huston, R.L. (1993). Flexibility effects on multibody systems. NATO-Advanced Study
Institute on Computer Aided Analysis of Rigid and Flexible Mechanical Systems,
Troia, Portugal.
105. IEC. (2007). Railway applications – Supply voltages of traction systems. IEC 60850, 3e.
106. Ikeda, M. and Mitsumoji, T. (2009). Numerical estimation of aerodynamic interference
between panhead and articulated frame. Quarterly Report of RTRI 50 (4): 227–232.
107. Ikeda, M., Mitsumoji, T., Sueki, T. and Takaishi, T. (2012). Aerodynamic noise reduc-
tion of a pantograph by shape-smoothing of panhead and its support and by the
surface covering with porous material. In: Noise and Vibration Mitigation for Rail
Transportation Systems: Notes on Numerical Fluid Mechanics and Multidisciplinary
Design 118 (eds. T. Maeda, P.-E. Gautier, C. Hanson, B. Hemsworth, J. Nelson, B.
Schulte-Werning, D. Thompson and P. Vos), 419–426. Tokyo: Springer.
Bibliography 341

108. Institut Mines-Telecom. (2019). A mini revolution in railway catenaries.


https://blogrecherche.wp.imt.fr/en/2019/10/03/a-minor-revolution-in-railway-catenary.
109. Iwnicki, S. (1999). The Manchester Benchmarks for Rail Vehicle Simulation, Supplement
to Vehicle System Dynamics. Taylor & Francis.
110. Iwnicki, S.D. (1998). Manchester benchmarks for rail vehicle simulation. Vehicle
System Dynamics 30 (3-4): 295–313.
111. Iwnicki, S.D. (Ed.), (1999). The Manchester Benchmarks for Rail Vehicle Simulation.
Swets & Zeitlinger. Lisse, the Netherlands.
112. Iwnicki, S.D. and Wickens, A.H. (1998). Validation of a MATLAB railway vehicle sim-
ulation using a scale roller rig. Vehicle System Dynamics 30: 257–270.
113. Jacobson, B. and Kalker, J.J. (2001). Rolling Contact Phenomena. Springer.
114. Jahnke, F.E. (1943). Tables of Functions. Dover Publications.
115. Jaschinski, A. (1990). On the application of similarity laws to a scaled railway bogie
model. Doctoral thesis. Delft University of Technology.
116. Jeambey, J. (1998). Improving high speed stability of freight cars with hydraulic
dampers. ASME International Mechanical Engineering Congress & Exposition (Rail
Transportation), Anaheim, California.
117. Johnson, K.L. (1958a). The effect of spin upon the rolling motion of an elastic sphere
upon a plane. ASME Journal of Applied Mechanics 25: 332–338.
118. Johnson, K.L. (1958b). The effect of a tangential contact force upon the rolling motion
of an elastic sphere on a plane. ASME Journal of Applied Mechanics 25: 339–346.
119. Johnson, K.L. (1985). Contact Mechanics. Cambridge University Press.
120. Jung, S.P., Kim, Y.G., Paik, J.S., and Park, T.W. (2012). Estimation of dynamic con-
tact force between a pantograph and catenary using the finite element method.
ASME Journal of Computational and Nonlinear Dynamics 7 https://doi.org/10.1115/1
.4006733.
121. Kalker, J.J. (1967). On the rolling contact of two elastic bodies in the presence of dry
friction. PhD thesis. Delft University, Netherlands.
122. Kalker, J.J. (1973). Simplified theory of rolling contact. Delft Progress Report 1: 1–10.
123. Kalker, J.J. (1979). Survey of wheel-rail rolling contact theory. Vehicle System Dynamics
8 (4): 317–358.
124. Kalker, J.J. (1980). Review of wheel-rail rolling contact theories. In: The General Prob-
lem of Rolling Contact (eds. L. Browne and N.T. Tsai). The 1989 ASME Winter Annual
Meeting, Chicago, IL, AMD-Vol.40, 77–92.
125. Kalker, J.J. (1982). A fast algorithm for the simplified theory of rolling contact. Vehicle
System Dynamics 11: 1–13.
126. Kalker, J.J. (1986). Wheel-rail wear calculations with the program contact. In: Contact
Mechanics and Wear of Rail/Wheel Systems II: Proceedings of the International Sympo-
sium Held at the University of Rhode Island, Kingston, RI, July 8–11, 1986 (eds. G.M.L.
Gladwell, H. Ghonem and J. Kalousek), 3–26. Waterloo, ON: University of Waterloo
Press.
127. Kalker, J.J. (1990). Three-Dimensional Elastic Bodies in Rolling Contact. Dordrecht:
Kluwer.
128. Kalker, J.J. (1991). Wheel-rail rolling contact theory. Wear 144: 243–261.
342 Bibliography

129. Kalker, J.J. (1994). Consideration of rail corrugation. Vehicle System Dynamics 23:
3–28.
130. Kalker, J.J. (1996). Discretely supported rails subjected to transient loads. Vehicle
System Dynamics 25: 71–88.
131. Karnopp, D. (2004). Vehicle Stability. Marcel Dekker.
132. Kassa, E., Andersson, C., and Nielsen, J. (2006). Simulation of dynamic interaction
between train and railway turnout. Vehicle System Dynamics 44 (3): 247–258.
133. Kerr, A.D. (2000). On the determination of the rail support modulus k. International
Journal of Solids and Structures 37: 4335–4351.
134. Kerr, A.D. and El-Sibaie, M.A. (1987). On the new equations for the lateral dynamics
of rail-tie structure. ASME Journal of Dynamic Systems, Measurement, and Control
107: 180–185.
135. Khulief, Y.A. and Shabana, A. (1986). Dynamics of multi-body systems with vari-
able kinematic structure, ASME Paper No. 85-DET-83. ASME Journal of Mechanisms,
Transmissions, and Automation in Design 108 (2): 167–175.
136. Khulief, Y.A. and Shabana, A.A. (1987). A continuous force model for the impact
analysis of flexible multi-body systems. Mechanism and Machine Theory 22 (3):
213–224.
137. Kik, W. (1992). Comparison of the behavior of different wheelset-track models. Vehicle
System Dynamics 20: 325–339.
138. Kik, W. and Piotrowski, J. (1996). A fast approximation method to calculate normal
load at contact between wheel and rail and creep forces during rolling. 2nd Mini
Conference on Contact Mechanics and Wear of Rail/Wheel System.
139. Klapas, D., Benson, F.A., Hackam, R., and Evison, P.R. (1988). Wear in simulated
railway overhead current collection systems. Wear 126 (2): 167–190.
140. Klingel, W. (1883). Über den Lauf von Eisenbahnwagen auf Gerarder Bahn. Organ für
die Fortschritte des Eisenbahnwesens 20: 113–123, Table XXI.
141. Knothe, K. and Bohm, F. (1999). History of stability of railway and road vehicle.
Vehicle System Dynamics 31: 283–323.
142. Knothe, K.L. and Grassie, S.L. (1993). Modeling of railway track and vehicle/track
interaction at high frequencies. Vehicle System Dynamics 22: 209–262.
143. Knothe, K. and Stichel, S. (1994). Direct covariance analysis for the calculation of
creepages and creep-forces for various bogie on straight track with random irregulari-
ties. Vehicle System Dynamics 23: 237–251.
144. Knudsen, C., Feldberg, R., and Jaschinski, A. (1991). Non-linear dynamic phe-
nomenon in the behavior of a railway wheelset model. Nonlinear Dynamics 2:
389–404.
145. Kono, H., Suda, Y., Yamaguchi, M., Yamashita, H., Yanobu, Y., and Tsuda, K. (2005).
Dynamic analysis of the vehicle running on turnout at high speed considering longi-
tudinal variation of rail profiles. IDETC/CIE 2005, ASME 2005 International Design
Engineering Technical Conference & Computers and Information in Engineering
Conference, September 24–28, Long Beach, California.
146. Kreyszig, E. (1991). Differential Geometry. Dover Publications.
147. Kubo, S. and Kato, K. (1998). Effect of arc discharge on wear rate of Cu-impregnated
carbon strip in unlubricated sliding against Cu trolley under electric current. Wear
216 (2): 172–178.
Bibliography 343

148. Kubo, S. and Kato, K. (1999). Effect of arc discharge on the wear rate and wear mode
transition of a copper-impregnated metallized carbon contact strip sliding against a
copper disk. Tribology International 32 (7): 367–378.
149. Kubota, Y., Nagasaka, S., Miyauchi, T., Yamashita, C., and Kakishima, H. (2013). Slid-
ing wear behavior of copper alloy impregnated C/C composites under an electrical
current. Wear 302 (1–2): 1492–1498.
150. Kulkarni, S., Pappalardo, C.M., and Shabana, A.A. (2017). Pantograph/catenary con-
tact formulations. ASME Journal of Vibration and Acoustics 139: 011010-1–011010-12.
151. Kumaniecka, A. and Jacek, S. (2008). Dynamics of the catenary modelled by a periodi-
cal structure. Journal of Theoretical and Applied Mechanics 46 (4): 869–878.
152. Kumaniecka, A. and Snamina, J. (2008). Dynamics of the catenary modelled by a peri-
odical structure. Journal of Theoretical and Applied Mechanics 46: 869–878.
153. Kwak, B.M. (1991). Complementarity problem formulation of three-dimensional
friction contact. ASME Journal of Applied Mechanics 58: 134–140.
154. Law, E.H. and Cooperrider, N.K. (1974). A survey of railway vehicle dynamics
research. ASME Journal of Dynamic Systems, Measurement, and Control 96 (2):
132–146.
155. Lee, J.H. and Park, T.W. (2012). Development of a three-dimensional catenary model
using the cable elements based on absolute nodal coordinate formulation. Journal of
Mechanical Science and Technology 26 (12): 3933–3941.
156. Lee, T.W. and Wang, A.C. (1983). On the dynamics of intermittent motion mecha-
nisms, part I: dynamic model and response. ASME Journal of Mechanisms, Transmis-
sions, and Automation in Design 105: 534–540.
157. Lee, H.W., Kim, K.C., and Lee, J. (2006). Review of maglev train technology. IEEE
Transactions on Magnetics 42: 1917–1925.
158. Liao, W.H. and Wang, D.H. (2003). Semiactive vibration control of train suspension
systems via magnetorheological dampers. Journal of Intelligent Material Systems and
Structures 14 (161): 161–172.
159. Lieh, J. and Haque, I. (1991). A study of the parametrically excited behavior of pas-
senger and freight railway vehicles using linear models. ASME Journal of Dynamic
Systems, Measurement, and Control 113: 336–338.
160. Ling, H. and Shabana, A.A. (2020). Euler angles and numerical representation of the
railroad track geometry. Acta Mechanica (in press).
161. Litvin, F.L. (1994). Gear Geometry and Applied Theory. Englewood Cliffs, NJ: Prentice
Hall.
162. Liu, Z. (2015). Numerical study on multi-pantograph railway operation at high
speed. Licentiate thesis. KTH Royal Institute of Technology, Stockholm, Sweden.
https://www.diva-portal.org/smash/get/diva2:856402/FULLTEXT01.pdf.
163. Lodec Jinshu. (2016). Aluminium innovation in high-speed trains.
http://lodecjinshu.com/en/innovation-aluminium-trains.
164. Love, A.E.H. (1944). A Treatise on the Mathematical Theory of Elasticity, 4e. Dover.
165. Luo, Y., Yin, H., and Hua, C. (1996). The dynamic response of railway ballast to the
action of trains moving at different speeds. Proceedings of the Institution of Mechanical
Engineers, Part F: Journal of Rail and Rapid Transit 210: 95–101.
344 Bibliography

166. Mace, S., Pena, R., Wilson, N., and DiBrito, D. (1996). Effects of wheel-rail contact
geometry on wheelset steering forces. Wear 191: 204–209.
167. Malvezzi, M., Presciani, P., Allotta, B., and Toni, P. (2003). Probabilistic analysis
of braking performance in railways. Proceedings of the Institution of Mechanical
Engineers, Part F: Journal of Rail and Rapid Transit 217: 149–165.
168. Manory, R. and Sinkis, H. (2000). A sliding wear tester for overhead wires and current
collectors in light rail systems. Wear 239 (1): 10–20.
169. Marquis, B. and Grief, R. (2011). Application of Nadal limit in the prediction of wheel
climb derailment. Paper JRC2011-56064. ASME/ASCE/IEEE Joint Rail Conference,
Pueblo, Colorado, March 16–18.
170. Massat, J.P., Laine, J.P., and Bobillot, A. (2006). Pantograph–catenary dynamics simu-
lation. Vehicle System Dynamics 44 (1): 551–559.
171. Matsudaira, T. (1960). Paper awarded prize in the competition sponsored by Office
of Research and Experiment (ORE) of the International Union of Railways (UIC),
Utrecht. ORE-Report RP2/SVA-C9 UIC.
172. Matsumoto, A., Sato, A., Ohno, Y., Suda, H., Nishimura, Y., Tanimoto, R., and Oka,
M. (1999). Compatibility of curving performance and hunting stability of railway
bogie. Vehicle System Dynamics Supplement 33: 740–748.
173. Mayville, R., Rancatone, R., and Tegeler, L. (1998). Investigation and simulation of
lateral bucling in trains. ASME International Mechanical Engineering Congress &
Exposition (Rail Transportation), Anaheim, California.
174. McClanachan, M., Handoko, Y., Dhanasekar, M., Skerman, D., and Davey, J. (2004).
Modeling freight wagon dynamics. Vehicle System Dynamics Supplement 41: 438–447.
175. McGonigal, R.S. (2006). Grades and curves. Trains. http://trn.trains.com/railroads/
abcs-of-railroading/2006/05/grades-and-curves.
176. McPhee, J.J. and Anderson, R.J. (1996). A model reduction procedure for the dynamic
analysis of rail vehicles subjected to linear creep forces. Vehicle System Dynamics 25:
349–367.
177. Meijaard J.P. (1991). Dynamics of mechanical systems; algorithms for a numeri-
cal investigation of the behaviour of non-linear discrete models. PhD thesis. Delft
University of Technology.
178. Meijaard, J.P. and De Pater, A.D. (1989). Railway vehicle systems dynamics and
chaotic vibrations. International Journal of Non-Linear Mechanics 24 (1): 1–17.
179. Melzer, F. (1994). Symbolic computations in flexible multibody systems. In: Pro-
ceedings of NATO-Advanced Study Institute on Computer Aided Analysis of Rigid and
Flexible Mechanical Systems, vol. II, 365–381. Troia, Portugal.
180. Meng, Q., Heineman, J., and Shabana, A.A. (2005). A longitudinal force model for
multibody railroad vehicle system applications. DETC2005/MSNDC-84058). ASME
International Design Engineering Technical Conferences and Computer and Informa-
tion in Engineering Conference, Long Beach, California, September 24–28.
181. Miyamoto, T., Suda, Y., and Ishida, H. (2006). Dynamics simulation of railway vehicles
on vibrated track upon seismicity. Third Asian Conference on Multibody Dynamics,
Tokyo, Japan, August 1–4.
Bibliography 345

182. Murase, S., Tanaka, A., Takahashi, Y., and Sekine, Y. (1988). Rail vehicle dynamics of
pitch caused by phase difference inputs. In: Proceedings of the ASME Rail Transporta-
tion Winter Conference, Chicago, Illinois, 45–51.
183. Nadal, M.J. (1908). Locomotives a vapeur. In: Collection Encyclopédie Scientifique,
vol. 186. Paris: Bibliothèque de Mécanique Appliquée et Génie.
184. Nagase, K., Wakabayashi, Y., and Sakahara, H. (2002). A study of the phenomenon of
wheel climb derailment: results of basic experiment using model bogies. Proceedings of
the Institution of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 216:
237–247.
185. Nakanishi, T., Yin, X.G., and Shabana, A.A. (1996). Dynamics of multibody tracked
vehicles using experimentally identified modal parameters. ASME Journal of Dynamic
Systems, Measurement, and Control 118: 449–507.
186. Netter, H., Schupp, G., Rulka, W., and Schroeder, K. (1998). New aspects of contact
modeling and validation within multibody system simulation of railway vehicles.
Vehicle System Dynamics Supplement 28: 246–269.
187. Newland, D.E. and Cassidy, R.J. (1975). Suspension and structure: some fundamental
design considerations for railway vehicles. Railway Engineering Journal 4 (2): 4–26.
188. Nielsen, J.C.O. (1994). Dynamic interaction between wheel and track – a paramet-
ric search towards an optimal design of rail structures. Vehicle System Dynamics 23:
115–132.
189. Nikravesh, P.E. (1988). Computer-Aided Analysis of Mechanical Systems. New Jersey:
Prentice Hall.
190. Nilsson, C.M., Jones, C.J.C., Thompson, D.J., and Ryue, J. (2009). A waveguide finite
element and boundary element approach to calculating the sound radiated by railway
and tram rails. Journal of Sound and Vibration 321: 813–836.
191. O’Connor, D.N., Eppinger, S.D., Seering, W.P., and Wormley, D.N. (1997). Active con-
trol of a high-speed pantograph. ASME Journal of Dynamic Systems, Measurement,
and Control 119: 1–4.
192. O’Donnell, W. (1993). Experimental and numerical investigation of the pitch and
bounce response of a railroad vehicle. M.S. thesis. Department of Mechanical
Engineering, University of Illinois at Chicago.
193. Ogden, R.W. (1984). Non-linear Elastic Deformations. Dover Publications.
194. Oura, Y., Yoshifumi, M., and Hiroki, N. (2008). Railway electric power feeding
systems. Railway Technology Today 16 (3): 48–58.
195. Ozaki, T. and Shabana, A.A. (2003a). Treatment of constraints in complex multibody
systems, part I: methods of constrained dynamics. International Journal for Multiscale
Computational Engineering 1 (2&3): 235–252.
196. Ozaki, T. and Shabana, A.A. (2003b, 2003). Treatment of constraints in complex
multibody systems, part II: application to tracked vehicles. International Journal for
Multiscale Computational Engineering 1 (2&3): 253–276.
197. Pappalardo, C.M., Patel, M.D., Tinsley, B., and Shabana, A.A. (2016). Contact force
control in multibody pantograph/catenary systems. Proceedings of the Institution of
Mechanical Engineers, Part K: Journal of Multi-body Dynamics 230 (4): 307–328.
198. Park, T.J., Han, C.S., and Jang, J.H. (2003). Dynamic sensitivity analysis for the panto-
graph of a high-speed rail vehicle. Journal of Sound and Vibration 266 (2): 235–260.
346 Bibliography

199. Pascal, J.P. (1993). About multi-Hertzian-contact hypothesis and equivalent conicity in
the case of S1002 and UIC60 analytical wheel/rail profiles. Vehicle System Dynamics
22: 57–78.
200. Pascal, J.P. and Sauvage, G. (1993). The available methods to calculate the wheel/rail
forces in non Hertizian contact patches and rail damaging. Vehicle System Dynamics
22: 263–275.
201. Pereira, M. and Ambrosio, J. (eds.) (1994). Computer-Aided Analysis of Rigid and
Flexible Mechanical Systems. Dordrecht: Kluwer.
202. Pfeiffer, F. and Glocker, C. (1996). Multibody Dynamics with Unilateral Contacts.
New York: Wiley.
203. Poetsch, G., Evans, J., Meisinger, R., Kortum, W., Baldauf, W., Veitl, A., and
Wallaschek, J. (1997). Pantograph/catenary dynamics and control. Vehicle System
Dynamics 28: 159–195.
204. Polach, O. (1999). A fast wheel-rail forces calculation computer code. Vehicle System
Dynamics Supplement 33: 728–739.
205. Polach, O. (2001). Influence of locomotive tractive effort on the forces between wheel
and rail. Vehicle System Dynamics Supplement 35: 7–22.
206. Polach, O. (2005). Creep forces in simulation of traction vehicles running on adhesion
limit. Wear 258: 992–1000.
207. Pombo, J. and Ambrosio, J. (2003). A wheel/rail contact model for rail guided vehicles
dynamics. ECCOMAS Thematic Conference on Advances in Computational Multibody
Dynamics, Lisboa, Portugal, July 1–4.
208. Pombo, J. and Ambrósio, J. (2012). Influence of pantograph suspension characteris-
tics on the contact quality with the catenary for high speed trains. Computers and
Structures 110 (111): 32–42.
209. Pombo, J., Ambrósio, J., Pereira, M., Rauter, F., Collina, A., and Facchinetti, A.
(2009). Influence of the aerodynamic forces on the pantograph–catenary system for
high-speed trains. Vehicle System Dynamics 47 (11): 1327–1347.
210. Pop, K. and Schiehlen, W. (1993). Fahrzeugdynamik. Stuttgart: Teubner.
211. Press, W.H., Teukolsky, S.A., Vetterling, W.T., and Flannery, B.P. (1992). Numerical
Recipes in Fortran, 2e. Cambridge University Press.
212. Profilidis, V.A. (2000). Railway Engineering. Cambridge University Press.
213. Rabinowicz, E. (1995). Friction and Wear of Material. Wiley.
214. Railsystem. (2015). High-speed rail. www.railsystem.net/high-speed-rail.
215. Rathod, C. and Shabana, A.A. (2006a). Geometry and differentiability requirements in
multibody railroad vehicle dynamic formulations. Nonlinear Dynamics 47 (1): 249–261.
216. Rathod, C. and Shabana, A.A. (2006b). Rail geometry and Euler angles. ASME Journal
of Computational and Nonlinear Dynamics 1 (3): 264–268.
217. Raymond, G.P. and Cai, Z. (2005). Dynamic track support loading from heavier and
faster train sets. Transportation Research Record 1381: 53–59.
218. Riches, E. (1988). Will maglev lift off? IEE Review 34: 427–430.
219. RifTek Sensors & Instruments. (2008). Laser wheel profilometer, user’s manual, IKP-5,
IKP-5R series. www.riftek.com.
220. Ripke, B. and Knote, K. (1995). Simulation of high frequency wagon-track interaction.
Vehicle System Dynamics Supplement 24: 72–85.
Bibliography 347

221. Roberson, R.E. and Schwertassek, R. (1988). Dynamics of Multibody Systems.


Springer-Verlag.
222. Rosenberg, R.M. (1977). Analytical Dynamics of Discrete Systems. Plenum Press.
223. Samin, J.C. and Neuve, L. (1984). A multibody approach for dynamic investigation of
rolling systems. Ingenieur Archiv 54: 1–15.
224. Sanborn, G., Heineman, J., and Shabana, A.A. (2007). A low computational cost non-
linear formulation for multibody railroad vehicle systems. Paper DETC2007-34522.
ASME Design Engineering Technical Conferences, Las Vegas, Nevada, September 4–7.
225. Schmid, R., Endlicher, K.O., and Lugner, P. (1994). Computer simulation of the
dynamical behavior of a railway bogie passing a switch. Vehicle System Dynamics
23: 481–499.
226. Schupp, G. (1996). Different contact models for wheel-rail systems: a compari-
son. Internal report IB 515-96-22. Institute for Robotics and System Dynamics,
DLR-Oberpfaffenhofen, Germany.
227. Schupp, G. (2003). Simulation of railway vehicles: necessities and application. Mechan-
ics Based Design of Structures and Machines 31: 297–314.
228. Schupp, G., Weidemann, C., and Mauer, L. (2004). Modelling the contact between
wheel and rail within multibody system simulation. Vehicle System Dynamics 41 (5):
349–364.
229. Schwab, A.L. (2002). Dynamics of flexible multibody systems. PhD thesis. Delft
University of Technology.
230. Schwab, A.L. and Meijaard, J.P. (2002). Two special finite elements for modelling
rolling contact in a multibody environment. In: Proceedings of the First Asian Confer-
ence on Multibody Dynamics, ACMD’02 (31 July to 2 August 2002), Iwaki, Fukushima,
Japan (eds. N. Shimizu et al.), 386–391. The Japan Society of Mechanical Engineering.
231. Schwab, A.L. and Meijaard, J.P. (2003). Dynamics of flexible multibody systems with
non-holonomic constraints: a finite element approach. Multibody System Dynamics 10:
107–123.
232. Schwartz, B. (1988). An analytical study of the bounce motion of a freight car model
in response to profile irregularities. ASME Rail Transportation Conference, Chicago.
233. Schwarz, B. (1991). Wheel climb and rollover potential due to excess elevation and
curvature, ASME Paper No. RTD. 4: 83–91.
234. Schwertassek, R. and Wallrap, O. (1999). Dynamik flexibler Mehrkorper Systeme.
Germany: Vieweg.
235. Seering, W., Armbruster, K., Vesely, C., and Wormley, D. (1991). Experimental and
analytical study of pantograph dynamics. ASME Journal of Dynamic Systems, Measure-
ment, and Control 113: 591–605.
236. Seo, J.H., Sugiyama, H., and Shabana, A.A. (2005). Three dimensional large deforma-
tion analysis of the multibody pantograph/catenary systems. Nonlinear Dynamics 42:
199–215.
237. Seo, J.H., Kim, S., Jung, I., Park, T., Mok, J., Kim, Y., and Chai, J. (2006). Dynamic
analysis of a pantograph–catenary system using absolute nodal coordinates. Vehicle
System Dynamics 44 (8): 615–630.
348 Bibliography

238. Shabana, A.A. (1986). Dynamics of inertia-variant flexible systems using experi-
mentally identified parameters. ASME Journal of Mechanisms, Transmissions, and
Automation in Design 108: 358–366.
239. Shabana, A.A. (1996a). Theory of Vibration: An Introduction, 2e. Springer Verlag.
240. Shabana, A.A. (1996b). An absolute nodal coordinate formulation for the large rota-
tion and deformation analysis of flexible bodies. Technical report MBS96-1-UIC.
Department of Mechanical Engineering, University of Illinois at Chicago.
241. Shabana, A.A. (1997). Flexible multibody dynamics: review of past and recent devel-
opment. Multibody System Dynamics 1: 189–222.
242. Shabana, A.A. (2005). Dynamics of Multibody Systems, 3e. Cambridge, UK: Cambridge
University Press.
243. Shabana, A.A. (2010). Computational Dynamics, 3e. New York: Wiley.
244. Shabana, A.A. (2018). Computational Continuum Mechanics, 3e. Chichester, UK:
Wiley.
245. Shabana, A.A. (2019). Dynamics of Multibody Systems, 5e. Cambridge, UK: Cambridge
University Press.
246. Shabana, A.A. and Ling, H. (2019). Noncommutativity of finite rotations and def-
initions of curvature and torsion. ASME Journal of Computational and Nonlinear
Dynamics 14: 091005-1–091005-10.
247. Shabana, A.A. and Rathod, C. (2007). Geometric coupling in the wheel/rail con-
tact formulations: a comparative study. Proceedings of the Institution of Mechanical
Engineers, Part K: Journal of Multi-body Dynamics 221: 147–160.
248. Shabana, A.A. and Sany, J.R. (2001). An augmented formulation for mechanical sys-
tems with non-generalized coordinates: application to rigid body contact problems.
Nonlinear Dynamics 24: 183–204.
249. Shabana, A.A., Hussien, H.A., and Escalona, J.L. (1998). Application of the absolute
nodal coordinate formulation to large rotation and large deformation problems. ASME
Journal of Mechanical Design 120: 188–195.
250. Shabana, A.A., Berzeri, M., and Sany, J.R. (2001). Numerical procedure for the simula-
tion of wheel/rail contact dynamics. ASME Journal of Dynamic Systems, Measurement,
and Control 123 (2): 168–178.
251. Shabana, A.A., Zaazaa, K.E., Escalona, L.J., and Sany, J.R. (2004). Development of
elastic force model for wheel/rail contact problems. Journal of Sound and Vibration
269: 295–325.
252. Shabana, A.A., Tobaa, M., Sugiyama, H., and Zaazaa, K.E. (2005). On the computer
formulations of the wheel/rail contact. Nonlinear Dynamics 40: 169–193.
253. Shabana, A.A., Tobaa, M., Marquis, B., and El-Sibaie, M. (2006). Effect of the lin-
earization of the kinematic equations in railroad vehicle system dynamics. ASME
Journal of Computational and Nonlinear Dynamics 1: 25–34.
254. Shabana, A.A., Zaazaa, K.E., and Sugiyama, H. (2008). Railroad Vehicle Dynamics:
A Computational Approach. Boca Raton, FL: CRC Press, Taylor & Francis Group.
255. Shabana, A.A., Hamper, M.B., and O’Shea, J.J. (2013). Rolling condition and gyro-
scopic moments in curve negotiations. ASME Journal of Computational and Nonlinear
Dynamics 8: 0111015-1–011015-10.
Bibliography 349

256. Shampine, L. and Gordon, M. (1975). Computer Solution of ODE: The Initial Value
Problem. San Francisco, CA: Freeman.
257. Sharma, V., Sneed, W., and Punwani, S. (1984). Freight equipment environmental
sampling test-description and results. ASME Rail Transportation Spring Conference,
Chicago.
258. Shen, G. and Goodall, R. (1997). Active yaw relaxation for improved bogie perfor-
mance. Vehicle System Dynamics 28: 273–289.
259. Shen, G. and Pratt, I. (2001). The development of railway dynamics modelling and
simulation package to cater for current industrial trends. Proceedings of the Institution
of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 215: 167–178.
260. Shen, Z.Y., Hedrick, J.K., and Elkins, J.A. (1983). A comparison of alternative
creep-force models for rail vehicle dynamic analysis. Vehicle System Dynamics 12:
79–87.
261. Shen, G., Ayasse, J.B., Chollet, H., and Pratt, I. (2003). A unique design method for
wheel profiles by considering the contact angle function. Proceedings of the Institution
of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 217, Part F: 25–30.
262. Shi, H. and Wu, P. (2016a). A nonlinear rubber spring model containing fractional
derivatives for use in railroad vehicle dynamic analysis. Proceedings of the Institution
of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 230: 1745–1759.
263. Shi, H. and Wu, P. (2016b). Flexible vibration analysis for car body of high-speed
EMU. Journal of Mechanical Science and Technology 30: 55–66.
264. Shi, J., Wei, Q.C., and Zhao, Y. (2007). Analysis of dynamic response of the high-speed
EMS maglev vehicle/guideway coupling system with random irregularity. Vehicle
System Dynamics 45 (12): 1077–1095.
265. Shi, H., Wang, J., Wu, P., Song, C., and Teng, W. (2018). Field measurements of the
evolution of wheel wear and vehicle dynamics for high-speed trains. Vehicle System
Dynamics 56: 1187–1206.
266. Shikin, E.V. and Plis, A.I. (1995). Handbook on Splines for User. Boca Raton, FL: CRC
Press.
267. Shust, W.C., Elkins, J.A., Kalay, J.A., and El-Sibaie, M. (1997). Wheel climb derail-
ment tests using AAR’s track loading vehicle. Report R-910. Association of American
Railroads.
268. Simeon, B., Fuhrer, C., and Rentrop, P. (1991). Differential algebraic equations in
vehicle system dynamics. Surveys on Mathematics for Industry 1: 1–37.
269. Sinha, P.K. (1987). Electromagnetic Suspension: Dynamics & Control. London:
P. Peregrinus.
270. Sinokrot, T., Nakhaeinejad, M., and Shabana, A.A. (2008). A velocity transformation
method for the nonlinear dynamic simulation of railroad vehicle systems. Nonlinear
Dynamics 51: 289–307.
271. Smith, C.A.M. and Bowler, E.H. (1981). A curved track simulator. XV Pan American
Railway Congress, Mexico.
272. Smith, R.. (2019). A railroad switch on the Schynige Platte-Bahn rack railway
in Switzerland. https://www.reddit.com/r/InfrastructurePorn/comments/8xbvq3/
a_railroad_switch_on_the_schynige_plattebahn_rack.
350 Bibliography

273. Spiryagin, M., Wu, Q., and Cole, C. (2017). International benchmarking of longitudi-
nal train dynamics simulators: Benchmarking questions. Vehicle System Dynamics 55
(4): 450–463.
274. Stichel, S. (1999). On freight wagon dynamics and track deterioration. Proceedings of
the Institution of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 213:
243–254.
275. Stickland, M.T. and Scanlon, T.J. (2001). An investigation into the aerodynamic char-
acteristics of catenary contact wires in a cross-wind. Proceedings of the Institution of
Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 215 (4): 311–318.
276. Stickland, M.T., Scanlon, T.J., Craighead, I.A., and Fernandez, J. (2003). An investiga-
tion into the mechanical damping characteristics of catenary contact wires and their
effect on aerodynamic galloping instability. Proceedings of the Institution of Mechanical
Engineers, Part F: Journal of Rail and Rapid Transit 217 (2): 63–71.
277. Suda, Y. and Anderson, R.J. (1994). High speed stability and curving performance of
longitudinal unsymmetric trucks with semi-active control. Vehicle System Dynamics
23: 29–52.
278. Sugiyama, H. and Shabana, A.A. (2006). Trajectory coordinate constraints in multi-
body railroad vehicle systems. Third Asian Conference on Multibody Dynamics,
Tokyo, Japan, August 1–4.
279. Sugiyama, H., Escalona, J.L., and Shabana, A.A. (2003). Formulation of three-
dimensional joint constraints using the absolute nodal coordinates. Journal of Non-
linear Dynamics 31: 167–195.
280. Sun, Y.Q. and Dhanasekar, M. (2001). A dynamic model for the vertical interaction
of the rail track and wagon system. International Journal of Solids and Structures 39:
1337–1359.
281. Swanson, J. (2019). Railroad workers need to fight for workers control over trains,
safety. The Militant 83 (25) https://themilitant.com/2019/07/06/railroad-workers-
needto-fight-for-workers-control-over-trains-safety.
282. Sweet, L.M. and Garrison-Phelan, P. (1980). Identification of wheel/rail creep coeffi-
cients from steady state and dynamic wheelset experiments. In: The General Problem
of Rolling Contact (eds. L. Browne and N.T. Tsai), The ASME 1980 Winter Annual
Meeting, Chicago, IL, AMD-Vol. 40, 77–92.
283. Sweet, L.M. and Karmel, A. (1981). Evaluation of time-duration dependent wheel load
criteria for wheel climb derailment. ASME Journal of Dynamic Systems, Measurement,
and Control 103: 219–227.
284. Szolc, T. (1998). Simulation of bending-torsional-lateral vibrations of the railway
wheelset-track system in the medium frequency range. Vehicle System Dynamics 30:
473–508.
285. Thompson, D. (2009). Railway noise and vibration: Mechanisms, modelling and means
of control. Elsevier. Oxford, UK, 2009.
286. Thompson, D.J. and Jones, C.J.C. (2000). A review of the modelling of wheel/rail
noise generation. Journal of Sound and Vibration 231: 519–536.
287. Tse, Y. and Martin, G. (1975). Flexible body railroad freight car, vol. I. Technical
report R-199. Association of American Railroads.
Bibliography 351

288. Tur, M., García, E., Baeza, L., and Fuenmayor, F.J. (2014). A 3D absolute nodal coor-
dinate finite element model to compute the initial configuration of a railway catenary.
Engineering Structures 71: 234–243.
289. Tuten, J.M., Law, E.H., and Coperrider, N.K. (1979). Lateral stability of freight cars
with axles having different wheel profiles and asymmetric loading. ASME Journal of
Engineering for Industry 101: 1–16.
290. US Department of Transportation, Federal Railroad Administration, Office of Safety.
(2004). Track safety standards, part 213, subpart G. Distributed by the Railway Educa-
tional Bureau.
291. US Department of Transportation, Federal Railroad Administration, Office of Safety.
(2005). Track safety standards, part 213, subpart A to F. Distributed by the Railway
Educational Bureau.
292. Vaccaro, A. (2019). Take the E-train? MBTA mulling electric locomotives.
Boston Globe. https://www.bostonglobe.com/metro/2019/03/21/take-train-mbta-
mullingelectric-locomotives/kWekh87ZPI7IKKr3EjmIUJ/story.html.
293. Valtorta, D., Zaazaa, K.E., Shabana, A.A., and Sany, J.R. (2001). A study of the lateral
stability of railroad vehicles using a nonlinear constrained multibody formulation.
ASME International Mechanical Engineering Congress and Exposition, New York.
294. Vermeulen, P.J. and Johnson, K.L. (1964). Contact of nonspherical bodies transmitting
tangential forces. ASME Journal of Applied Mechanics 31: 338–340.
295. Wang, W., Suda, Y., Komine, H., Sato, Y., Nakai, T., and Shimokawai, Y. (2006). Tran-
sition curve negotiating simulation of railway vehicle for air suspension control to
prevent wheel load reduction. Third Asian Conference on Multibody Dynamics,
Tokyo, Japan, August 1–4.
296. Warren, P., Samavedam, G., Tsai, T., and Gomes, J. (1998). Passenger vehicle dynamic
performance – tests and analysis correlation. ASME International Mechanical Engi-
neering Congress & Exposition (Rail Transportation), Anaheim, California.
297. Wehage, R.A. (1980). Generalized coordinate partitioning in dynamic analysis of
mechanical systems. PhD thesis. University of Iowa, Iowa City.
298. Weidemann, C., Mauer, L., and Arnold, M. (2006). Improving the calculation speed
for time domain integration of complex railway vehicles. Third Asian Conference on
Multibody Dynamics, Tokyo, Japan, August 1–4.
299. Weinstock, H. (1984). Wheel climb derailment criteria for evaluation of rail vehicle
safety. Paper 84-WA/RT-1. ASME Winter Annual Meeting.
300. White, R.C., Limbert, J.K., Hedrick, K., and Cooperrider, N.K. (1978). Guidewaysus-
pension tradeoffs in rail vehicle systems. Report DOT-OS-50107. US Department of
Transportation, Washington, DC.
301. Wickens, A.H. (1965). The dynamic stability of railway vehicle wheelsets and bogies
having profiled wheels. International Journal of Solids and Structures 1: 319–341.
302. Wickens, A.H. (1986). Non-linear dynamics of railway vehicles. Vehicle System
Dynamics 15: 289–301.
303. Wickens, A.H. (1988). Stability optimization of multi-axle railway vehicles possessing
perfect steering. ASME Journal of Dynamic Systems, Measurement, and Control 110:
1–7.
352 Bibliography

304. Wickens, A.H. (1991). Steering and stability of bogie: vehicle dynamics and suspension
design. Proceedings of the Institution of Mechanical Engineers, Part F: Journal of Rail
and Rapid Transit 205: 109–122.
305. Wickens, A.H. (1996). Static and dynamic instabilities of bogie railway vehicles with
linkage steered wheelsets. Vehicle System Dynamics 26: 1–16.
306. Wickens, A. (1998). Key note address. Vehicle System Dynamics 30 (3–4): 193–195.
307. Wickens, A.H. (2003). Fundamental of Rail Vehicle Dynamics. Swets & Zeitlinger.
308. Wilson, N., Shu, X., and Kramp, K. (2004). Effect of independently rolling wheels on
flange climb derailment. ASME International Mechanical Engineering Congress.
309. Wilson, N., Wu, H., Tournay, H., and Urban, C. (2009). Effects of wheel/rail contact
patterns and vehicle parameters on lateral stability. Vehicle System Dynamics 48 (S1):
487–503.
310. Wilson, N.G., Fries, R., Witte, M., Haigermoser, A., Wrang, M., Evans, J., and Orlova,
A. (2011). Assessment of safety against derailment using simulations and vehicle
acceptance tests: A worldwide comparison of the state-of-the-art assessment methods.
Proceedings 22nd IAVSD Symposium, 14–19 August 2011, Manchester, UK, CD-ROM.
311. Wu, T.X. and Brennan, M.J. (1998). Basic analytical study of pantograph–catenary sys-
tem dynamics. Vehicle System Dynamics 30: 443–456.
312. Wu, T.X. and Thompson, D.J. (2002). A hybrid model for the noise generation due to
railway wheel flats. Journal of Sound and Vibration 251: 115–139.
313. Wu, H. and Elkins, J. (1999). Investigation of wheel flange climb derailment criteria.
Report R-931. Association of American Railroads.
314. Wu, W.X., Brickle, B.V., Smith, J.H., and Luo, R.K. (1998). An investigation into
stick-slip vibrations on vehicle/track systems. Vehicle System Dynamics 30: 229–236.
315. Wu, Q., Cole, C., Luo, S., and Spiryagin, M. (2014). A review of dynamics modelling
of friction draft gear. Vehicle System Dynamics 52: 733–758.
316. Wu, X., Cai, W., Chi, M., Lai, W., Shi, H., and Zhu, M. (2015). Investigation of the
effects of sleeper-passing impacts on the high-speed train. Vehicle System Dynamics 53:
1902–1917.
317. Wu, Q., Spiryagin, M., and Cole, C. (2016). Longitudinal train dynamics: an overview.
Vehicle System Dynamics 54: 1688–1714.
318. Yabuno, H., Okamoto, T., and Aoshima, N. (2001). Stabilization control for the hunt-
ing motion. Vehicle System Dynamics Supplement 35: 41–55.
319. Yakoub, R.Y. and Shabana, A.A. (2001). Three dimensional absolute nodal coordinate
formulation for beam elements: implementation and applications. ASME Journal of
Mechanical Design 123 (4): 614–621.
320. Yang, G. (1994). Aspects in modeling a railway vehicle on an arbitrary track, ASME
Rail Transportation, RTD. 8: 31–36.
321. Yang, Y.B. and Yau, J.D. (2011). An iterative interacting method for dynamic analysis
of the maglev train guideway/foundation soil system. Engineering Structures 33 (3):
1013–1024.
322. Yau, J.D. (2009). Response of a maglev vehicle moving on a series of guideways with
differential settlement. Journal of Sound and Vibration 324 (3–5): 816–831.
Bibliography 353

323. Yau, J.D. (2010). Aerodynamic vibrations of a maglev vehicle running on flexible
guideways under oncoming wind actions. Journal of Sound and Vibration 329 (10):
1743–1759.
324. Yokoyama, N. (2009). Research and development toward wear reduction of current
collecting system. JR East Technical Review 13: 50–54.
325. Zaazaa, K.E. (2003) Elastic force model for wheel/rail contact in multibody railroad
vehicle systems. PhD thesis. Department of Mechanical Engineering, University of
Illinois at Chicago.
326. Zboiński, K. (1998). Dynamical investigation of railway vehicles on a curved track.
European Journal of Mechanics - A/Solids 17 (6): 1001–1020.
327. Zboinski, K. (1999). The importance of imaginary forces and kinematic-type nonlin-
earities for the description of railway vehicle dynamics. Proceedings of the Institution of
Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 213: 199–210.
328. Zboinski, K. and Dusza, M. (2006). Development of the method and analysis for non-
linear lateral stability of railway vehicles in a curved track. Vehicle System Dynamics
44 (sup 1): 147–157.
329. Zhang, Y., El-Sibaie, M., and Lee, S. (2004). FRA track quality indices and distribution
characteristics. AREMA Annual Conference, Nashville.
330. Zhai, W. (2019). Vehicle–track coupled dynamics: Theory and application. Springer,
Singapore.
331. Zhai, W.M., Cai, C.B., and Guo, S.Z. (1996). Coupling model of vertical and lateral
vehicle/track interactions. Vehicle System Dynamics 26 (1): 61–79.
332. Zheng, X.J., Wu, J.J., and Zhou, Y.H. (2000). Numerical analyses on dynamic control
of five-degree-of-freedom maglev vehicle moving on flexible guideways. Journal of
Sound and Vibration 235 (1): 43–61.
333. Zhou, L. and Shen, Z.Y. (2013). Dynamic analysis of a high-speed train operating on
a curved track with failed fasteners. Journal of Zhejiang University Science A 14 (6):
447–458.
334. Zhou, D., Tian, H.Q., Zhang, J., and Yang, M.Z. (2014). Pressure transients induced
by a high speed train passing through a station. Journal of Wind Engineering and
Industrial Aerodynamics 135: 1–9.
335. Zobory, I. (1997). Prediction of wheel/rail profile wear. Vehicle System Dynamics 28:
221–259.
355

Index

a Acceleration transformation, see Acceleration,


AAR, see Association of American Railroads transformation
(AAR) Acceleration vector, see Acceleration, vector
AAR wheel profiles 128 Active control system 39, 221, 321
Absolute acceleration, see Acceleration, absolute Adhesion region 178, 206, 210
Absolute coordinates 311 Aerodynamic force 1, 37, 39, 40, 41, 119, 292,
Absolute nodal coordinate formulation, see 293, 296, 298, 307, 321, 322–324, 325
ANCF Algebraic constraint equations, see
Absolute velocity, see Velocity, absolute Constraint(s), equations
Acceleration 27, 30, 40, 119, 120, 182, 183, 186, Alignment deviations 21, 22, 104–105, 138,
226, 231, 232, 234, 235, 238, 242, 244, 164, 166–169
245–251, 253, 256, 258, 261, 269–273, Analytical geometry-variation functions
278–279, 290, 301, 308, 316, 319, 324 164–166. See also Deviations
absolute 24, 25, 100 ANCF 9, 38, 46, 60, 78, 292, 169–173
analysis 243, 259 accelerations 319
angular 96, 99, 123, 181, 235, 240–242, 253, beam element 79, 145, 299, 301–303, 305,
270 322
Cartesian 123 cable element 299, 304, 308
centrifugal 26 catenary model 38, 298–304, 309, 313, 322,
centripetal component, see Acceleration, 323–324
normal component Cholesky accelerations 308
dependent 119, 251, 253 Cholesky coordinates 319
equation 23, 95–97, 122, 225, 234 coordinates 80, 318
independent 119, 251, 253, 254 displacement field 143, 172, 301–304, 307
normal component 96, 238 elastic force 305
tangential component 96 equations 143
transformation 242, 270–271 finite elements 9, 38, 40, 46, 79, 139, 143,
vector 11, 23, 26, 31, 97, 235, 238, 269–270, 145, 148, 156, 172, 296, 299, 308, 310,
304 311, 318, 326
Acceleration analysis, see Acceleration, analysis fully-parameterized elements 79, 143, 299,
Acceleration equations, see Acceleration, 301–303, 312
equation geometry 82, 141, 142–148, 214
Acceleration kinematic equation 27, 183, 184, gradient coordinates 9, 46, 79, 145, 172, 300
186, 226, 244–245, 249, 253–254, 261, gradient deficient elements 322
273, 278, 316, 319 inertia force 304–307

Mathematical Foundation of Railroad Vehicle Systems: Geometry and Mechanics,


First Edition. Ahmed A. Shabana.
© 2021 John Wiley & Sons Ltd. Published 2021 by John Wiley & Sons Ltd.
356 Index

ANCF (contd.) B-splines, see Computational geometry methods,


interpolation 22, 126, 135, 140, 141, B-splines
145–146, 148, 157, 161, 166, 169–173, 183 Bullhead rail, see Rail, bullhead
kinematics 79–81 Bump deviation, see Deviations, bump
mass matrix 304, 319 Bushing 261
mesh 143 axes 267
position field 145, 322 coordinate system 266–267
position gradients 9, 46, 143–145, 148, 299, damping 266
309, 311, 315 element 262, 266–267
rail geometry 142–148 forces 267
shape function matrix 80, 303, 324 material 266–267
Angle of attack 33–35, 180 stiffness 266–267
Angular acceleration, see Acceleration, angular
Angular velocity, see Velocity, angular c
Applied forces 100, 102, 119, 226, 228–229, Cant angle 21, 104–105
232, 235, 239–241, 251, 254, 256, 260, Cant deficiency 25
261, 267, 271, 285, 307, 308, 313 Cant excess 25
APTA 120-wheel profile 130 Carter’s creep theory 204–205
Arc length 7, 13, 23, 25, 33, 45, 49–53, 58, 63, creep forces 204–205
67–71, 84, 89, 104, 106, 107, 110–113, creepage coefficient 205
120–122, 136, 144–146, 150–155, 157, creepages 199
159–161, 166, 169, 170, 215–217, 268, law 205
269, 271, 274, 277, 306, 311–315 Cartesian coordinates 10, 18, 32, 45, 46, 48, 55,
Association of American Railroads (AAR) 34 59, 69, 83, 104, 115, 116, 119–122, 127,
Augmented contact constraint formulation 257 189, 215–217, 230, 245, 264, 268,
Augmented formulation 31, 119, 226, 232, 271–272, 317
244–254, 312–313, 318 Catenary 1, 4, 5, 30, 36–41, 220, 291–328
Axle load 2, 3, 36, 131, 163, 220, 223, 224 aerodynamic forces 322–324
ANCF model, see ANCF, catenary model
b cables 4, 36–41
Balance speed 3, 22–26, 54, 126, 154, 238 contact wire 37–41, 293, 295–298, 309, 313,
Ball joint 115, 230 315, 316, 320, 321, 324–328
Bank angle 21, 24–26, 106, 110, 113, 149, contact 28, 36–41, 176, 219, 292, 293, 313
152–158 data 293–294
Beam element, see ANCF, beam element deformation 30
Bearing element 29, 30, 131, 219, 261 design 292, 298, 316
Binormal vector 51–53, 108–109 droppers 37, 295, 296
Body coordinate system, see Coordinate system, dynamic equations, see Dynamic equations,
body catenary
Body reference point 10, 31, 83–85, 96, 114, elastic forces 305–306
122, 225, 232, 238 elastic waves 38
Body trajectory coordinate system, see equations 36–41, 298–304, 306–308, 317–321
Coordinate system, body-trajectory finite elements, see ANCF, catenary model
Bogie 28–29, 219, 261, 326 force control 321–322
bolster 29 forces 309
components 29 galloping motion 40, 322
frame 29, 261 geometry 79, 143, 296, 299
structure 219 inertia forces 304–305
Bolster, see Bogie, bolster mathematical formulation 38
Index 357

MBS formulation 317–321 Conic surface 56, 57, 61, 64, 68, 125, 128, 198,
messenger wire 37, 295, 296 284
nodal coordinates 308, 312 Constant forward velocity 23, 285
pretension 296–297 constraint 271
structure 295–296 Constrained dynamic equations, see Dynamic
voltage specifications 297–298 equations, constrained
wear 40–41 Constrained dynamics 227–232
wire 1, 37, 36–41, 292, 293, 308–309, 316, Constraint(s) 5, 6, 30, 39, 227
324–328 approach 254–256
Center of mass 14, 15, 23–25, 31, 85, 127, 216, contact formulation, see Contact formulation,
225, 232, 233, 238, 247, 311 constraint
Centerline equations 7, 11, 27, 30, 32, 43, 83, 87, 90, 91,
ANCF element 308, 311 114–119, 226, 229, 231, 277–279
track-section 137 Euler-parameter 94, 95, 243
track space curve 13, 14, 26, 32, 41, 84, 106, forces 30, 31, 39, 44, 119, 184, 185, 187, 219,
149, 156, 157, 159, 160, 161, 271, 272, 226, 228, 232, 245, 246, 250
274, 324 formulation 176, 182–186, 310–314
Centrifugal force 3, 10, 22–26, 30, 126, 235, 238 functions 43, 44, 229, 244, 250
Centripetal acceleration 96 holonomic 250
Chain rule of differentiation 43 Jacobian matrix 32, 181, 187, 226, 229, 231,
Characteristic equation 72 244, 246, 249–250, 277–279
Cholesky accelerations, see ANCF, Cholesky Joint 114–119
accelerations kinematic 30, 100, 244
Cholesky coordinates, see ANCF, Cholesky motion 1, 30, 33, 114–119, 135, 182, 227
coordinates specified motion trajectories 1, 30, 33, 43,
Chord definition of curvature, see Curvature, 100, 115, 118, 227, 229, 256
chord definition trajectory-coordinate 271–274
Circular profile, see Profile, circular wheel/rail 254–259
Complete cubic spline, see Cubic splines, Contact angle 34, 201–202, 284
complete Contact area 57, 125, 175–179, 181, 188, 190,
Computational algorithms 1, 4, 20, 27, 34, 38, 194, 199, 204–206, 208, 210–212, 214,
176, 184, 188, 208, 210, 298, 301, 317, 318 215, 261, 284, 296, 329, 331
Computational approach 9, 30–31, 38, 143, 291 Contact constraint(s) 181, 256, 257, 259
Computational fluid dynamics 40 Contact constraint conditions 180, 256
Computational geometry methods 79, 143 Contact constraint formulation 257, 260
B-splines 143 Contact ellipse 28, 176, 181, 188, 193–199, 203,
NURBS 143 205–214, 329, 333
Computer-aided design (CAD) 79, 143, 173 ratio 208
Computer modeling 17, 18, 125, 168, 292, 296 semi-axes 195, 205, 213, 333
Computer program(s) Contact equations 57, 308, 329–333
CONTACT 208 Contact force 4, 9, 17, 26–29, 38–41, 45, 46, 57,
FASTSIM 212, 213 61, 103, 125, 132, 140, 160, 175–188,
USETAB 213 194–199, 203, 204, 206, 209, 210, 213,
Cone geometry 56 218, 219, 255–261, 282, 288, 295, 298,
Conformal contact 26, 27, 175–178, 183, 186, 307–310, 313, 316, 319–322, 324–328
188, 213, 255 Contact force control, see Pantograph/
Conical profile 14, 128, 197 catenary, contact force control
Conicity 3, 14, 16–18, 125, 128, 283, 284, 287, Contact formulation(s) 26–28, 33, 36, 200
288 augmented 257
358 Index

Contact formulation(s) (contd.) passive 293


conformal 178 derivative 322
constraint 39, 176, 180, 183–194, 187, 201, derivative/bang-bang 322
203, 213–219, 259, 310, 310–314, 318, derivative/exponential 322
319, 320, 322 Coordinate(s)
creep 178 absolute, see Absolute coordinates
elastic 39, 176, 180, 181, 183–186, 187, 194, ANCF, see ANCF, coordinates
196, 197, 201, 203, 213–219, 259–261, body 85
308, 310, 314–317, 319, 320, 322 Cartesian, see Cartesian coordinates
pantograph/catenary 39–40, 310–317, 319 catenary 320
planar 216–219 Cholesky, see ANCF, Cholesky coordinates
wheel/rail 28, 33, 176, 213, 254, 259–261, dependent 30, 95, 115, 119, 187, 226, 227,
317, 326 231, 246, 249, 251, 252, 257, 272, 273
Contact frame 213, 284 elastic 317
pantograph/catenary 308–310, 311, 312, 320 generalized 11, 13, 31, 43, 83, 84, 89, 97–100,
Contact point 18, 57, 61, 127, 140, 169, 172, 114, 119, 120, 179, 182–184, 187, 188,
178, 181–184, 187, 188, 200–204, 201, 217, 218, 223, 225–229, 231,
210–215, 216, 219, 254, 255, 257, 280, 255–259, 263, 265, 267, 270, 289, 290,
282, 283 310–315, 317–320, 323, 324
coordinates 180, 181, 203, 214 gradient, see ANCF, gradient coordinates
location of 27, 34, 57, 132, 152, 176, 179, 180, independent 7, 10, 30, 31, 83, 84, 107, 115,
183, 185, 186, 188, 203, 214–215, 254, 119, 225–227, 231, 233, 246, 249–252,
255, 260 257, 277, 279. See also Degrees of
pantograph/catenary 298, 302, 309–316, freedom
318, 320 nodal 9, 38, 46, 60, 79–82, 126, 143, 145, 146,
surface parameters 156, 172, 181, 185, 186, 148, 166, 172, 214, 292, 299–308, 311,
196 312, 315, 318, 319
velocity 175, 178, 179, 181, 200, 201, 202, pantograph 320
214, 247, 260, 283 partitioning 231
wheel/rail 4, 22, 27, 34, 45, 82, 126, 139, 148, position 5, 22, 113, 114, 126, 145, 148, 149,
152, 169, 176, 180, 184, 203, 216, 219, 151, 152, 154–156, 159–161, 169–173,
247, 248, 260, 282 227, 250, 323
Contact pressure 26, 175, 179, 194, 198–199 rail 182, 258
Contact region 17, 26, 125, 175, 177, 178, 181, redundant 119, 226, 231, 232, 245, 251
190, 194, 199, 203, 328 selection 31
adhesion 175, 178–179 trajectory 31, 32–33, 119–123, 215, 216, 225,
slip 175, 178–179 268–274, 277, 278
Contact strip, see Pantograph, contact strip transformation 190, 223, 268
Contact surface 26, 76, 175, 180, 184, 185, Coordinate system
187–194, 199, 200, 204, 208, 211–213, body 10, 11, 31, 33, 43, 84–86, 88–90, 96–99,
216, 218, 219, 254–255, 257, 330 120–123, 156, 157, 160, 180–182, 217,
Contact theory 212–214, 261, 282, 329 218, 225, 230, 232–234, 236, 239, 240,
creep, see Creep, contact theory 245, 255, 268–272, 274, 289, 309, 311,
Hertz, see Hertz, contact theory 317, 318, 323, 324
nonlinear 213, 282 body-track 31, 225
wheel/rail 212–213 bushing, see Bushing, coordinate system
Contact wire, see Catenary, contact wire profile, see Profile curve, frame
Controllers rail 20, 105, 136, 139, 140–145, 147, 161, 169,
active system, see Active control system 182, 183, 254
Index 359

track 31, 110, 135, 225, 274 road 138


trajectory-body 31, 33, 120–122, 225, section 138, 139
268–272, 274 Cubic interpolation 76–78, 79, 145, 148, 156,
wheel 18, 19, 64, 68, 127, 133, 182, 216, 217 172, 173, 301, 302
wheelset 127, 128, 132, 140, 201, 202, 280, Cubic polynomial 237, 300
282, 283 Cubic splines 76–78, 81, 128, 146
Coriolis force 30 complete 77
Co-simulation approaches 38, 318 natural 77
Coulomb friction 27, 175, 178–179, 199, 205, Current collection 1, 37, 39–41, 294, 296, 298,
319, 321 315, 316, 321, 323–328
Crane rail, see Rail, crane Curvature 5, 7, 10, 11, 12, 19, 21, 23, 25, 26,
Creep 50–54, 57, 67, 69–75, 78, 84, 103, 108,
coefficients 210, 284 110–114, 125, 126, 142, 144, 152, 154,
contact formulation 178, 203–213 157, 160–162, 168, 169, 170, 301
contact theory 205 ANCF element curvature 306
force 26–28, 61, 175–176, 179, 181, 187, 194, chord definition 112
199, 200, 203–213, 219, 257, 261, 282, continuity 54
284–286 curve 10, 19, 26, 45, 50, 103, 106, 110, 125,
linear force model 284 126
phenomenon 27, 175, 179, 199, 200 Gaussian 73, 74
spin moment 26–28, 175, 176, 179, 181, 187, helix 53
194, 199, 200, 203, 207, 208, 212, 213, 285 horizontal 12, 84, 104, 106, 112–114,
Creep-force formulations 203–213 149–158, 162
Carter’s theory 204–205 mean 73, 74
heuristic nonlinear model 208–209 normal 69–75, 193
Johnson and Vermeulen’s theory 205–206 principal 9, 17, 28, 46, 57, 62, 72–75, 82, 125,
Kalker’s linear theory 206–208 148, 176, 181, 188, 190, 193, 196, 203, 214
Polach nonlinear model 210 radius of 5, 10, 13, 19, 23, 25, 26, 50, 59, 70,
simplified theory 210–212 104, 112, 125, 126, 150, 159, 188, 190,
Creep spin moment, see Creep, spin moment 197, 247
Creepage 27, 28, 57, 175, 176, 179, 181, rail space curve 79, 106
207–210, 212, 213, 282, 284 spiral 10, 54, 103, 106
coefficients 205, 207, 208, 213, 214 surface 62, 190, 261
definition 199–203, 282–284 track space curve 106
discontinuity 214 vector 22, 25, 50–54, 66, 67, 69, 70, 71, 89,
force relationship 205, 206 107–109, 111, 112, 126, 132–135, 141,
lateral 179, 199–203, 206, 208, 210, 211 142, 144, 147, 148, 150, 153, 156, 157,
longitudinal 179, 199–206 161, 169–171, 192
normalized-velocity 179, 200, 203 vertical 111, 112, 153, 154, 162, 171, 172
spin 179, 199–203, 205, 207–210, 213, wheel 132–135
286 Curve
Critical speed 187, 261, 286–288 arc length, see Arc length
Cross-level variation 166 binormal vector, see Binormal vector
Cross product 41–42, 51, 59, 70, 71, 87, 108 circular 5, 7, 19, 23–25, 50, 52, 54, 59, 67, 69,
Crossing 138, 139, 165 125, 131, 247
grade 165, 166 curvature, see Curvature
level 138 geometry 6, 13, 46–54, 79, 84, 89, 107–108,
panel 139 110, 113, 114, 143, 152, 166
rail 163 helix 48, 52–53
360 Index

Curve (contd.) identification of 249–250


implicit form 7, 45–48 Dependent coordinates, see Coordinates,
normal vector 51, 52, 69, 70, 108, 109 dependent
parameterization 6–8, 60 Derailment 3, 4, 10, 17, 18, 21, 22, 33–36, 45,
parametric form 7, 8, 45–47, 59, 63, 64, 67, 54, 103, 125, 126, 128, 132, 138, 139, 162,
68, 74, 107, 110, 113, 114, 126, 135, 164, 185, 187, 219, 238, 260
152 Derailment criteria 33–36
radius of curvature, see Curvature, radius Association of American Railroads (AAR)
segment 12, 19, 50, 84, 111, 113, 125, 151, Chapter11, 50-ms time limit 34
152 Electro-Motive Diesel (EMD) L/V time
tangent vector 9, 48–52, 61, 70, 107, 108, duration criterion 34
111, 300 Federal Railroad Administration (FRA)
theory of 1, 3, 4–6, 33, 45–46, 53–54 high-speed passenger distance limit 34
torsion 7, 45, 51–53, 108, 110 Japanese National Railway (JNR) L/V time
Curving performance 128 duration criterion 34
Cusp deviations, see Deviations, cusp Nadal single-wheel L/V limit criterion 34
Cylindrical joints, see Joints, cylindrical Transportation Technology Center, Inc.
Cylindrical profile, see Profile, cylindrical (TTCI) wheel climb distance criterion
Cylindrical surface 7, 48, 56–58, 61, 63, 66, 69, 34
71, 73, 177, 204, 253 Weinstock axle-sum L/V limit criterion
Cylindrical wheels 16, 68 34
Derivatives of Euler angles, see Euler angles,
d derivatives
D’Alembert’s principle 32, 119, 226, 251, 254, Development angle 110, 112, 113, 149, 152–158
256, 259 Deviations 163, 164–169. See also Rail
Damped sinusoidal deviation, see Deviations, irregularities
damped sinusoidal alignment (lateral), see Alignment deviations
Damping allowable 166
coefficient 28, 39, 176, 181, 185, 187, 188, bump 165–167
216, 254, 260–262, 264–267, 286, 287, cusp 165–167
296, 310, 316, 317, 320 damped sinusoidal 165–167
matrix 267 functions 165, 166
model 319 geometry 296
Deformation 189, 193, 194, 299, 301, 304, 305, jog 165–167
317 measured data 219
analysis 302 occurrence 165
bending 299 plateau 165–167
large 318 profile (vertical), see Profile deviations
local 179 shapes 165
modes 302, 303, 304, 306 sinusoidal 165–167
plastic 325 standard 164, 322, 327
rail 9, 46, 143, 145 track 21, 22, 162, 164–169
relative 178, 179, 199, 297 trough 165–166
shear 299, 301, 303 types 165
Degrees of freedom 27, 30, 31, 44, 115, 116, Differential/algebraic equations (DAE) 30,
119, 180, 184, 226, 229, 230, 249–253, 119, 232
255, 257, 260, 268, 273, 274, 279, 299, Differential geometry 3–9, 20, 26, 45–82,
301, 311–314, 316, 317, 320. See also 188
Coordinates, independent Differentiation of vector function 43–44
Index 361

Direction cosines 11, 84, 86–88, 90, 323 Elliptical integrals 194, 206, 329–333
Discontinuities 138, 178, 179, 214, 215, Embedding technique 31, 119, 226, 232,
301 244–254, 273, 278, 312–313
Drag and lift force(s) 40, 322, 323 Equations of motion, see Dynamic equations;
coefficients 40 Dynamic formulation(s)
Driving constraints, see Trajectory constraint(s) Euler angles 10–13, 31, 83, 84, 86, 88–91, 94,
Droppers, see Catenary, droppers 95, 97–110, 113–115, 120, 136, 140,
Dry friction, see Coulomb friction 143–145, 148–161, 168–172, 201, 202,
Dynamic curving 27, 31–32 216, 217, 225, 230, 231, 233, 234, 242,
Dynamic equations 1, 4, 11, 22, 31, 38, 41, 83, 246, 268, 269, 280, 281
100, 126, 152, 187, 223, 225–290 derivatives 101, 161, 170
catenary 299 singular configuration 90–91, 100, 101, 233
constrained 31, 34, 225–290 Euler equation, see Newton-Euler equations
linearized 284–286 Euler parameter(s) 11, 31, 83, 91–95, 97–102,
of motion 1, 9, 19, 31, 42, 125, 179, 184, 216, 114, 115, 118, 225, 230, 231, 233, 234,
225, 225–290 243–244, 246
nonlinear 1, 4, 9, 19, 125, 149, 179, 182 constraint 94
Dynamic formulation 17, 44, 319 identities of 94, 99
augmented formulation, see Augmented Euler-parameter constraint, see Constraint(s),
formulation Euler-parameter
constrained 292 Exact differential 11
embedding technique, see Embedding External force 228, 229, 247, 253, 307, 318
technique
f
e FASTSIM program, see Computer program(s),
Elastic contact formulation. See also Contact FASTSIM
formulation(s), elastic Federal Railroad Administration (FRA) 34, 163
algebraic equations 184–186 Office of Safety 163
lookup tables 180, 185, 214, 260 First fundamental form of surfaces 9, 45, 62,
nodal search method 185, 214–216, 218, 260 62–64, 71
Elastic coordinates, see Coordinate(s), elastic coefficients 62–64, 73–75, 82, 148, 181, 190,
Elastic deformation 296, 317, 318. See also 192
Deformation Flange, see Wheel, flange
Elastic force 143, 187, 194, 195 Flange angle 34, 36, 128, 129
analysis 194 Flange contact 3, 14, 34, 178, 199, 208
catenary 304–308 Flange geometric details and dimensions 129,
formulation 260 130
Elastic half-space 188, 194, 205, 329, 330 Flexible bodies 33, 119, 143, 268, 299, 301, 306,
Electric arcing 325 307, 317, 318, 323, 324
Electrification system 36, 291, 298 Force
Electrodynamic suspension (EDS) 220–221 aerodynamic, see Aerodynamic force
Electromagnetic suspension (EMS) 220–222, centrifugal, see Centrifugal force
289, 290 contact, see Contact force
Electro-mechanical phenomenon 325 Coriolis, see Coriolis force
Electromechanical systems 288–290 creep, see Creep, force
Elevation angle, see Development angle drag and lift, see Drag and lift force(s)
Elimination of surface parameters 257–259, elastic, see Elastic force
313–314 friction 26, 34, 39, 41, 61, 175, 178, 179, 199,
Elliptical contact area 208, 331 200, 209, 247, 293, 320, 321, 325
362 Index

Force (contd.) Generalized coordinates, see Coordinate(s),


generalized 11, 31, 83, 181, 223, 225, generalized
226–228, 239–241, 264, 265, 284, 288, Generalized coordinate partitioning, see
289, 307, 319, 320, 323 Coordinate(s), partitioning
gravity 10, 23, 25, 238, 285, 286 Generalized external force, see External force
inertia, see Inertia force Generalized force vectors, see Force, generalized
levitation 220–224, 288–290 Generalized mass matrix 308
Maglev 219–224. See also Force of attraction, Generalized Newton–Euler equations 240–241
Maglev force Generalized trajectory coordinates 121
magnetic 28, 176, 219–221, 224, 288 Geometry
suspension 221–223, 261, 285 ANCF, see ANCF, geometry
uplift, see Uplift force computational, see Computational geometry
Force balance methods
centrifugal force 24, 25 differential, see Differential geometry
Nadal 36 integration with mechanics, see Geometry and
Force of attraction mechanics, integration of
Maglev force 221 numerical representation 76–82
Force element(s) rail 4, 17, 20, 22, 45, 46, 60, 61, 80, 83, 125,
126, 135–148, 157, 171, 257
bushing 266–268
railroad 57–60, 81, 125–173
rotational spring-damper-motor 263–264
track 1, 9–13, 19–23, 30, 84, 102–107,
series spring damper 265
111–114, 120, 121, 125, 136, 138, 144,
spring-damper-actuator 261–263
148–155, 156, 160, 164, 166, 168, 268,
Forward velocity 14–17, 23, 33, 118, 200–201,
269, 275, 277
203, 211, 271, 272, 277, 285, 287
wheel 3, 4, 17, 18, 45, 83, 125, 128, 129–135,
Free body diagram 31, 32, 226, 232, 248, 249,
142, 257
254
Geometry and Mechanics 1–3, 9, 23, 179, 180,
Freight trains 1–3, 28, 36, 83, 163, 177, 219,
182
220, 223, 224, 277, 291
integration of 9–13, 179. See also I-CAD-A
Frenet frame 51–53, 107–109
Global position
Friction force, see Force, friction
of ANCF node 300, 302
Full saturation 27, 175, 199
of contact point 181, 183, 186, 200, 202, 204,
Fully-parameterized ANCF finite elements, see 254, 280, 282, 311, 320
ANCF, fully-parameterized elements of joint definition point 230, 245
of particle 307
g of point 10, 83, 84, 85, 95, 114, 122, 161, 169,
Gage 20, 21, 24, 25, 103, 104, 160, 162, 233, 239, 311
167–169, 284 of reference point 10, 31, 83, 120, 122, 133,
limits 163, 166 180, 201, 217, 225, 230, 245, 269, 280,
point 34, 130, 131 311, 317
value 104, 158, 161, 163, 169, 287 vector 10, 84, 85, 122, 161, 181, 230
variation 135, 166, 169 Grade 12, 21, 84, 104, 106, 108, 110, 112–114,
widening 35, 164 142, 144, 149–152, 156–158, 162, 165,
Galloping motion, see Catenary, galloping 166, 169, 275
motion Gradient-deficient ANCF finite elements, see
Gaussian curvature, see Curvature, Gaussian ANCF, gradient-deficient elements
Gaussian elimination 249 Guard rail, see Rail, guard
Gaussian procedure 250 Gyroscopic moment 235, 238, 243, 276
General displacement 10–11, 32, 268 Gyroscopic motion 89, 90, 98
Index 363

h effect 203, 204


Half-space assumption, see Elastic half-space mass moments of 236, 237, 247, 253, 287,
Helix, see Curve, helix 294
curve 48 matrix 119, 319
pitch 48 moment 210, 254
Hertz principal moments of 236–237
coefficients 196, 197, 213 products of 236, 282
contact force 261 properties 297
contact pressure, see Contact pressure tensor 236, 237
contact theory 27–28, 176, 177, 181, Inertia coupling 31, 85, 225, 232, 238, 318
188–199, 203, 210, 213–215, 261, 329, 331 Inertia force 26, 31, 100, 226, 228–229, 232,
elliptical integrals 329–333 235–239, 241, 243, 244, 254, 256
force model 197 ANCF-catenary 298, 313. See also ANCF
stiffness coefficients 188, 261 inertia
Heuristic nonlinear creep-force model, see quadratic-velocity 237–238
Creep-force formulations, heuristic virtual work of 97, 228–229, 232, 235, 239,
nonlinear model 241, 251
High rail 23, 26, 166 Instability
High-speed trains 1–4, 30, 34, 36–41, 139, 163, Maglev suspension 289
219, 221, 291, 294, 296, 298, 321 magnetic levitation 223
Horizontal curvature 15 motion 17
Horizontal projection, see Projection, horizontal
source of 221
Hunting 14–17
unsuspended wheelset 286, 288
analysis 227
vehicle 132
frequency 14, 16, 17, 286
Integration of computer-aided design and
instability 16
analysis, see I-CAD-A
oscillation 3, 14–17, 24, 31, 225
Integration of geometry and analysis 1–2
phenomenon 14–17, 227
Integration of geometry and mechanics,
speed 128
see Geometry and mechanics, integration
stability 16, 227, 280–288
of
two-degree-of-freedom model 227–279
Intermediate wheel coordinate system 216, 217
wheelset 227
Hyperbolic at point 71, 74–75
Hyperbolic surfaces 66, 74–75 j
Jacobian matrix 44, 55, 100, 186, 187, 258, 271,
i 324
I-CAD-A 173 constraint equations, see Constraint(s),
Implicit form, see Curve, implicit form Jacobian matrix
Incremental-rotation approaches 38, 318 Japanese National Railway (JNR) L/V time
Independent coordinates, see Coordinate(s), duration derailment criterion, see
independent; Degrees of freedom Derailment criteria, Japanese National
identification of, see Degrees of freedom, Railway (JNR) L/V time duration
identification criterion
Independent velocities 247, 251, 273 Jog deviation, see Deviations, jog
Inertia Johnson and Vermeulen’s theory, see
car 28 Creep-force formulations, Johnson and
centrifugal force, see Centrifugal force Vermeulen’s theory
Coriolis, see Coriolis force Joints
distributed 28, 29, 38, 114 cylindrical joints 115, 116–118, 230
364 Index

Joints (contd.) Lagrangian formulation 44, 225, 318


mechanical 1, 30, 32, 43, 100, 114, 115, 118, Lagrangian mechanics 248
226, 227, 229, 246 Lateral contact force, see Contact force
prismatic (translational) joints 115, 117–118 Lateral creepage, see Creepage, lateral
revolute (pin) joints 29, 115, 116–117, 293 Lateral displacement 3, 14–16, 128, 277, 285,
sliding, see Sliding, joint 311, 312, 315
spherical joints 115–117, 230, 245, 293, 294 Lateral motion 17, 23, 26, 312
universal joints 115 Lateral oscillations 25
Lateral sliding, see Sliding, lateral
k Lead rail 138, 139
Kalker’s coefficient, see Creep, coefficients; Length of the rails 159–160
Creepage, coefficients Levitation force, see Force, levitation
Kalker’s linear theory, see Creep-force Linear algebra 41–44
formulations, Kalker’s linear theory Linear interpolation 78, 148, 152–155, 157,
Kalker’s USETAB, see Computer program(s), 169–171, 173
USETAB Linearization 1, 28
Kinematic(s) 1, 13, 14 Linearized approaches 28
acceleration level 27, 183, 184, 186, 226, 244, Linearized equations 35, 284, 285
245, 247, 249, 253, 254, 261, 273, 278, Linearized models 33
316, 319
Local deformation, see Deformation, local
ANCF element 79–82
Locomotive 130, 112
body 84–85, 135, 232–234
diesel 36, 112, 291
catenary 299
Longitudinal contact force, see Contact force;
position level 27, 115, 183, 184, 227, 247,
and Creep, force
251, 252, 260, 314, 316
Longitudinal creep forces, see Contact force; and
velocity level 27, 183, 184, 186, 245, 247, 251,
Creep, force
261, 277, 313, 316
Longitudinal shift 218, 220
Kinematic constraint equations, see
Longitudinal sliding, see Sliding, longitudinal
Constraint(s), equations
Longitudinal tangent 33, 108, 136, 149, 161,
linearization of 1
202, 204, 215, 218, 315
Kinematic coupling 216, 219
Kinematic singularity 90, 100–102 Lookup tables, see Elastic contact formulation,
Kinetic(s) lookup tables
energy 32, 226 LU factorization 290, 319
Klingel’s formula 17, 286
m
l Maglev forces, see Force, Maglev
L/V ratio 33–36. See also Derailment Criteria Maglev suspension 220–221, 289
Lagrange-D’Alembert equation 228 instability, see Instability, Maglev suspension
Lagrange-D’Alembert principle 119, 226, 251, stability, see Stability, Maglev suspension
256, 259 Maglev trains 28
Lagrange multipliers 32, 181, 226, 231–232, Magnetic levitation (Maglev)
244, 249, 251, 257, 259, 261, 271, 313, electrodynamic suspension (EDS), see
319, 320 Electrodynamic suspension
Lagrangian approach 31–32, 187, 225–226, electromagnetic suspension (EMS), see
228, 248 Electromagnetic suspension
Lagrangian dynamics 229, 248 forces, see Force, levitation
Lagrangian form of the equations of motion instability, see Instability, Maglev suspension,
249 and magnetic levitation
Index 365

modeling of electromagnetic suspensions n


237–240 Nadal’s criterion 34
MBS modeling of electromechanical systems Nadal’s formula 34, 36
288–290 Natural cubic spline, see Cubic splines, natural
Magnetic levitation forces, see Force, levitation Newton differences 186, 218, 246, 247, 250,
Mass matrix 30, 119, 123, 235–236, 238, 244, 258, 260, 313, 315
252, 256, 270, 271, 276, 299, 304, 308, Newton–Euler assumption 238
318–319 Newton–Euler equations 31, 85, 122, 123, 181,
Mass moments of inertia, see Inertia, mass 182, 225, 232, 234, 235, 239–244, 248,
moments of 271, 275, 280, 318
MBS, see Multibody systems Newtonian approach 11, 31, 226–227, 228, 248
Measured data 132, 163, 169 Newton–Raphson algorithm 132, 141, 180,
Measured track data 140, 162–169 184, 185, 186, 218, 246, 250, 252, 258,
Messenger wire, see Catenary, messenger wire 260, 313, 314, 315, 316
MiniProf 60, 76, 81, 128, 146 Nodal points 22, 104, 126, 143, 145, 148,
Modulus of rigidity 206, 208, 210, 330 153–156, 161, 172
Moment Nodal search method, see Elastic contact
gyroscopic, see Gyroscopic moment formulation, nodal search method
inertia, see Inertia, moment Nonconformal contact 258
Motion Non-generalized coordinates 318. See also
amplitude 114 Surface parameters
constraint 135, 182, 277, 242, 260, 271, 317, Nonlinear dynamic analysis 210
318 Nonlinear equations of motion. See also
gyroscopic, see Gyroscopic motion Dynamic equations, nonlinear
instability, see Instability, motion Normal component of the acceleration, see
scenarios 4, 24, 34, 39, 40, 45, 123, 160, 185, Acceleration, normal component
187, 204, 208, 214, 235, 271, 316, 324, Normal contact force 27, 28, 39, 61, 176, 178,
326 179, 181, 185, 187–188, 194–199, 203,
trajectory curve 23–25 204, 206, 210, 213, 218, 219, 254, 257,
Motion trajectories 261, 310, 316, 319
specified, see Constraint(s), specified motion Normal curvature, see Curvature, normal
trajectories Normal plane 51
Multibody system(s) (MBS) Normal stress 194
algorithms 10, 11, 31, 32, 38, 40, 83, 91, 101, Normal vector, see Curve, normal vector
119, 142, 226, 233, 254, 268, 293, 298, Normal wear rate 326
301, 317–321, 325, 326 Normalized relative velocities, see Creepage,
approach 4, 28–34, 38, 118 normalized-velocity
computer program(s) 149, 299 Numerical methods 119, 132
constrained dynamics 31, 289, 293 Numerical representation of geometry, see
dynamics 232 Geometry, numerical representation
dynamic equations 223, 226, 290, 318 Numerical solution 19, 125
formulation 318, 322 NURBS, see Computational geometry methods,
literature 226, 251 NURBS
modeling of electromechanical systems
289–290 o
model(s) 34, 322, 326 Orientation angles 122. See also Euler angles
railroad vehicle algorithms 244, 289 Orthogonal matrix 42–43, 51, 52, 87, 95, 237,
railroad vehicle model 30, 326 266
366 Index

Orthogonal vectors 51, 57, 63, 69, 71, 117, 237, normal contact force 39
309 penetration, see Penetration,
Oscillations, see Hunting, oscillation pantograph/catenary
performance 40, 324
p relative sliding 311, 312, 315–316
Pan head 293, 295, 310, 312, 315, 322, 324 separation 39, 310
catenary contact 297, 299, 309, 311, 314, stability 39
315, 316, 320, 321 technology 36, 291
catenary separation 326, 328 voltage specifications 397–398
coordinates 312–313 Pantograph/catenary vibration control, see
lateral displacement 315, 316 Controllers
wear 296 Pantograph/catenary wear 40–41, 296, 317,
Pantograph 292–294 324–328
balance arm 294 Joule effect 326
components 38, 40, 293–294, 323 lubrication effect 328
contact strip 324, 326 mechanical factor 326
data 293–294 models 326
lower arm 293–294 normal wear rate 326
lower link 294 Parabolic curve 46–47
plunger 294 Parabolic point 71
upper arm 293–294 Parabolic surfaces 66
upper link 294 Parametric form, see Curve, parametric form
Pantograph/catenary system 1, 4, 5, 36–41, Parameterization
220, 291–328 ANCF beam element 304
aerodynamic force 40, 322–324 curve, see Curve, parameterization
augmented formulation 312–313 rail surface 33, 254
constraint contact 316–317 surface 8, 81, 146
contact 28, 37 –38, 39–40, 176, 219, 292, 295, wheel surface 18, 33, 57, 58, 127, 254
297, 298, 308, 314, 317–322 Passenger train 2, 3, 36, 163, 219, 291
contact force control 39–40, 321–322 Passive controllers 293
contact formulations, see Contact Penalty method, see Elastic contact formulation
formulation(s), pantograph/catenary Penetration
contact frame 210–312 pantograph/catenary 39, 310, 314, 320
contact points, see Contact point, wheel/rail 181, 185, 186–188, 216, 218, 219,
pantograph/catenary 260–261
control 39 Pitch angle 33, 35, 89, 90, 106, 121, 201, 216,
coordinates 317–318 217, 233, 268, 272–273, 277, 283
design 292 Pitch angular velocity 285
dynamics 40, 323 Planar contact 214, 216–219
elastic contact formulation 314–317 Planar surface 66, 69, 71
elimination of geometric parameters Plateau deviation, see Deviations, plateau
313–314 Poisson’s ratio 206
embedding technique 312–313 Polach nonlinear creep-force model, see
equations 317– 321 Creep-force formulations, Polach
forces 309, 310, 313, 319–322 nonlinear model
interaction 38–40, 296, 299, 308, 322 Position analysis 187, 196, 247
interaction criteria 327 Position coordinates, see Coordinate(s), position
mathematical formulation 38 Pressure distribution, see Contact pressure
MBS algorithms 317–321 Pretension, see Catenary, pretension
model 30, 296, 322 Primary suspension 29, 261, 286–288
Index 367

Principal axes 237 Quadratic transformation matrix 94


Principal curvatures, see Curvature, principal Quadratic velocity inertia forces 235, 236,
Principal directions 9, 17, 57, 62, 73–75, 82, 237–238, 244, 304
125, 181, 188, 196, 214, 237 Quadratic velocity terms 123, 123, 235, 246
Principal moments of inertia, see Inertia, Quadratic velocity vector 253, 313, 314, 319
principal moments of Quasi-static creep force model 203
Principal radii of curvature 190
Prismatic joints, see Joints, prismatic r
(translational) joints Radius of curvature, see Curvature, radius of
Products of inertia, see Inertia, products of Rail
Profile curve 18, 21, 53, 57, 59, 60, 63, 67, 70, arc length 89, 144, 215, 216, 218. See also
71, 76, 125–132, 135–140 Arc length
AAR 128 base 137
circular 67 basic geometry and dimensions 137–138
conical, see Conical profile bullhead 138
conicity 128, 287 crane 138
cylindrical 59, 60, 64, 68, 128 flat-bottom 137
design 128–130 geometry, see Geometry, rail
deviations 21, 22, 104, 105, 164–168 guard 19, 138
dimensions 128–130 head 137
flange 130 irregularities 14, 104. See also Deviations
frame 18, 19, 57, 58–60, 63, 64, 68, 127, 131, lateral surface parameter 214, 215, 254
135–140, 161, 162, 182–183 length 159–160
function 18–20, 58, 63, 64, 68, 125, 127, 128, longitudinal direction 17, 125, 137, 139
131–132, 133, 135–140, 182–183, 285 longitudinal surface parameter 22, 126, 139,
geometry 9, 46, 58, 60, 76, 77, 79, 83, 128, 140, 145, 156, 172, 215, 254
131, 136, 143, 204, 277 profile, see Profile curve, rail
irregularities 4, 45, 53, 164–168 rollover 35
limits 168 segments 19, 141, 144–146, 149, 152–153,
measurement, see MiniProf 155, 160, 166, 169, 170, 172
rail 9, 17, 20, 34, 46, 53, 54, 60, 62, 67, 76, 81, semi-analytical approach 22, 126, 135–143,
83, 125, 135–140, 146, 168, 214–216 145 –148, 161, 166, 169–173, 188
segment 77 space curve 12, 13, 13, 21, 22, 54, 58, 59, 63,
shape 58 79, 84, 102, 104, 106, 107, 111, 126, 136,
slope 130 140, 144, 145, 148, 149, 152, 155–157,
terminologies 128, 136–138 159, 160, 166, 217, 218
tolerance 132 stock rail, see Stock rail
wheel 9, 14, 18, 46, 53, 54, 59, 60, 63, 68, 76, tongue rail, see Tongue rail
83, 125–132, 214–216, 218, 277 web 137
wear allowance 132 Railroad geometry, see Geometry, railroad
worn 4, 45, 128, 131 Railroad vehicle
Projection, horizontal 111–112, 149, 151 components 28
Pure rolling 199, 200, 202, 203, 285 constrained dynamics, see Constrained
Pure sliding, see Sliding, pure dynamics. See also Dynamic equations,
constrained
q dynamics 9–11, 20, 46, 57, 61, 76, 83, 84, 88,
Quadratic characteristic equation 72 89, 101, 178, 181, 188, 199, 208, 233
Quadratic function 190, 194, 223 equations, see Dynamic equations; Dynamic
Quadratic interpolation 190 formulation(s)
Quadratic polynomial 191, 192 formulations 32
368 Index

Railroad vehicle (contd.) resistance 128


geometry, see Geometry, railroad stock 130
linearized models, see Linearized models surface 130
motion instability, see Instability, motion Root loci 287
motion scenarios, see Motion scenarios Rotation matrix 88, 99
Rail surface 1, 7, 9, 17, 20, 22, 26, 33, 45, 53, Rodrigues’ formula 92
57–59, 61, 63, 67, 81, 125, 126, 131, simple rotations 88
135–146, 175, 179–181, 183–188, 199, in terms of direction cosines 86
203, 214, 215, 224–225, 302, 329 in terms of Euler angles 90
ANCF geometry 142–146 in terms of Euler parameters 94
equation(s) 9, 22, 126, 141, 182–183 Rotation parameters 88, 91, 180, 225
geometry 9, 17, 135–142, 142–146, 204, 301 Rotational spring-damper-motor element, see
higher derivatives 141 Force element(s), rotational
lateral parameter 62. See also Rail, lateral spring-damper-motor
surface parameter
normal vector 140, 142, 148, 186, 187, 200 s
parameter(s) 136, 140–142, 147, 161, 169, Secondary suspension 29, 220, 261
170, 172, 182 Second fundamental form of surfaces 9, 45, 62,
parameterization 146 65–68, 71, 73
tangent vector(s) 140, 200, 215 coefficients 65–68, 73, 74, 75, 82, 148, 181,
Rectifying plane 51 190, 192
Redundant coordinates, see Coordinate(s), Second point of contact, see Flange contact
redundant Semi-analytical approach, see Rail,
Reference point, see Body reference point semi-analytical approach
Relative angular velocity 181, 200 Series spring-damper element, see Force
Relative principal radii of curvature 190 element(s), series spring-damper
Relative radii of curvature 188 Serret–Frenet equations 50–53
Relative sliding, see Sliding, relative Shear stresses 194
Revolute joints, see Joints, revolute (pin) joints Simple rotation(s) 86–88, 90, 93, 94, 108
Rigid bodies 10, 84–85, 96, 227, 232–235, matrices 89, 90, 93, 98, 99, 106, 107
238–240, 310, 317, 318, 323 Simplified creep-force theory, see Creep-force
coordinates 85, 87, 317 formulations, simplified theory
displacement 83, 141, 145, 301 Single-degree-of-freedom joints
dynamics 84, 85, 232, 235 prismatic joint, see Joints, prismatic
general displacement 32, 268, 305 (translational) joints
kinematics 13, 84–85, 232 –234 revolute joint, see Joints, revolute (pin) joints
models 30, 323 Single-degree-of-freedom model 277
motion 32, 145, 189, 225, 243, 317, 318 Singular configuration(s) 233, 250, 309, 310
Rodrigues’ formula 91–95 of Euler angles, see Euler angles, singular
Roll angle 33, 59, 89, 90, 101, 106, 121, 201, configurations
268, 272, 277, 285 kinematic singularity 90, 100–102
Rolling Singular point 49
circle 130, 233 Sinusoidal deviation, see Deviations, sinusoidal
conditions 247, 248 deviation
constraints 248 Skew-symmetric matrix 41–43
contact 144, 175, 199, 202, 204, 211 Sliding 27, 144, 175, 178, 179, 199, 200, 210,
direction 205–207, 210, 212 247, 297, 321, 326
motion 179, 199, 200, 202, 203, 247, 285, 319 bar 297
radii 14, 17, 197, 198, 201, 205, 283, 285, 287 lateral 310–313, 315–316, 320, 321
Index 369

joint 311–313 equilibrium 40, 322


longitudinal 311–313, 315–316, 320, 321 force 321
pure 178, 199 quasi-, see Quasi-static creep force model
relative 26, 175, 199, 311, 313, 320, 321, Stock rail 138, 139
326 Stress distribution 188, 210, 329
speed 327 Successive rotations 89
Slip region, see Contact region, slip Super-elevation 12, 20, 21, 23–25, 84, 104, 106,
Space curves, see Curve 108, 110, 113, 114, 142, 144, 149–152,
Specialized railroad vehicle algorithms 119, 154, 156–158, 162, 168, 169, 238, 275
268 Surface(s)
Specialized railroad vehicle formulations 32 curvature, see Curvature, surface
Specialized software 119, 268 first fundamental form, see First fundamental
Spherical joints, see Joints, spherical joints form of surfaces
Spin creep moment 285 Gaussian curvature, see Curvature, Gaussian
Spin creepage, see Creepage, spin geometry 1, 4, 8, 9, 17, 19, 54–75, 78–82,
Spin moment, see Creep, spin moment 126–148, 177, 181, 188–194, 302
Spiral(s) 10–12, 19, 22, 23, 35, 54, 84, 103, 104, mean curvature, see Curvature, mean
114, 125, 126, 139, 143, 149, 152–154, normal vector 65, 69, 70, 135, 140, 142, 148
164–166 principal curvatures, see Curvature, principal
curve-to-curve 153, 157 principal directions, see Principal directions
curve-to-tangent 152, 157 rail, see Rail surface
curvature 54, 103 second fundamental form, see Second
design 125 fundamental form of surfaces
geometry 126, 154 tangent plane 60–62
negotiation 22, 54, 126 tangent vectors 61, 140
segment 12, 19, 54, 84, 103, 106, 125, theory of 4, 7–9, 45–46
154 wheel, see Wheel, surface
tangent-to-curve 54, 104, 157 Surface geometry, see Surface(s), geometry
Spline representation Surface parameters 7, 55, 57–61, 63, 65, 68, 69,
B-spline, see Computational geometry 73, 127, 189, 196, 318
methods, B-splines elimination of, see Elimination of surface
cubic, see Cubic splines parameters
Spring-damper-actuator element, see Force rail 19, 20, 23, 126, 135–136, 140–142,
element(s), spring-damper-actuator 147, 148, 156, 161, 169, 170, 172,
Stability 187, 238 180–187, 203, 214–216, 218, 219,
catenary 38, 296 254–260
current collection system 321 wheel 19, 23, 126, 127, 131–133, 135, 156,
hunting 277, 280–288 180–187, 203, 211, 214–216, 218, 219,
maglev suspension 221, 223, 289 254–260
pantograph/catenary 39, 293, 321 Suspended wheelset 285–288
vehicle 4, 9, 14, 17, 18, 21, 26, 27, 30, 36, 45, Suspension 24, 114, 219, 261, 286
61, 76, 115, 125, 132, 139, 164, 168, 176, characteristics 3, 286
181, 199, 235, 238, 261 circuit 290
wheelset 17, 128, 277, 280–288 damping 286
Static electrodynamic, see Electrodynamic
analysis 194 suspension
assumptions 189 electromagnetic, see Electromagnetic
considerations 188 suspension
equations 41 elements 261
370 Index

Suspension (contd.) Track data 155–162


forces 261, 285 file 160–162
Maglev train suspension, see Maglev train measured, see Measured track data
suspension Track deviations, see Deviations
primary, see Primary suspension Track geometry, see Geometry, track
secondary, see Secondary suspension Track irregularities 9, 21, 39, 40, 105, 157, 163,
stiffness 286 164, 166–168, 321, 324
Sway 297 Track preprocessor 13, 22, 84, 89, 104–106,
Switches 19, 138, 139, 166, 224 148, 153–172, 269
panel 139 data file 160–162
deviations 166–168
t input 89, 148–150, 156–159
Tangent plane, see Surface(s), tangent plane length of the rails 159–160
Tangent rail segment 11, 19, 54, 84, 125, 152, measured data, see Measured track data
163, 166 output 13, 84, 104, 126, 144, 153–154, 156
Tangent-to-curve entry spiral, see Spiral(s), Track quality 2, 83, 163
tangent-to-curve Track segments, see Rail, segments
Tangent vector(s), see Surface(s), tangent Track super-elevation, see Super-elevation
vectors. See also Curve, tangent vector Trajectory body coordinate system, see
Tangential component of the acceleration, see Coordinate system, trajectory-body
Acceleration, tangential component Trajectory constraint(s), see Constraint(s),
Tangential force(s) specified motion trajectories. See also
contact, see Contact force Constraint(s), trajectory-coordinate
creep, see Creep, force Trajectory coordinates, see Coordinate(s),
friction, see Force, friction
trajectory
Tangential stress 210
Trajectory coordinate constraints, see
Theory of curves, see Curve, theory of
Constraint(s), trajectory-coordinate
Theory of elasticity 204, 213, 305
Transformation matrix, see Rotation matrix
Theory of surfaces, see Surface(s), theory of
Translational spring-damper-actuator element,
Three-dimensional contact 180, 185, 214
see Force element(s),
Three-dimensional rotations 91, 93
spring-damper-actuator
Time derivatives
Tread 3, 17, 129, 132, 198
angular velocity 96
Chamfer 129
constraint equations 259, 314
contact 3
electric current 290
Euler angles 98, 281 cross section 128
Euler parameters 102 datum 128, 129
generalized coordinates 256, 319 profile 131
orientation parameters 11, 83, 95, 97, 181, run-out 132
230, 234, 235, 241, 263, 324 taper 128
penetration 181, 185, 187 Tread contact, see Tread, contact
surface parameters (non-generalized Trough deviation, see Deviations, trough
coordinates) 187, 256, 314 Turnouts 19, 138, 139, 165
trajectory coordinates 122, 123, 270 Twist 166
Tongue rail 138, 139 Two-degree-of-freedom
Torsion, see Curve, torsion cylindrical joint 116, 231
Track centerline, see Centerline, track space rail-car model 277–279
curve wheelset 285
Track classes 163–164, 166–168 Two-dimensional contact, see Planar contact
Index 371

u principle 31, 225, 227–232, 239–241, 251,


Unit vector(s) 10, 11, 26, 42, 49–51, 56, 61, 69, 306, 307
71, 85–87, 90, 91, 93, 94, 133 stress (elastic) forces 305–307
Unsuspended wheelset
w
instability, see Instability, unsuspended
Warp, see Twist
wheelset
Wear 132
Uplift force 37, 39–42, 293, 294, 321–323,
allowance 132
325–327
catenary 39, 295, 296, 327
control 322, 325
characteristics 326
USETAB software, see Computer program(s),
electric effect 325
USETAB
electro-mechanical phenomenon 325
v formulation 326
Vehicle high rail 166
dynamic behavior 115 Joule effect 325
dynamic equations 289 mechanical effect 325
dynamics 9–11, 20, 27, 30, 40, 46, 57, 58, 61, mechanical factor 326
76, 83, 84, 88, 89, 101, 126, 139, 175, 182, mechanism 325, 326, 328
188, 199, 208, 233, 238, 321, 324 model 325, 326
dynamics algorithms 178 normal wear rate 326, 327
instability, see Instability, vehicle pan-head carbon strip 296, 316, 324
Velocity pantograph/catenary 39–41, 292, 293, 316,
absolute 11, 31, 96, 100, 120, 181, 200–202, 321, 322, 324–328
215, 225, 234, 247, 269, 274, 283, 301, profile 54, 132
304 rate 324, 325, 326, 328
angular 11, 14, 41–43, 83, 90, 95–97, resistance 128
100–102, 122, 18, 200–203, 215, 230, 234, wheel/rail 112, 128, 219, 220
239, 242, 245, 247, 263, 270, 272, 274, wheel tire roll 130
281, 283, 285, 324 Wheel
relative 27, 61, 175, 178, 179, 181, 187, 199, angular surface parameter 118, 59, 60, 63,
200, 203, 211, 214, 265, 323, 325 127, 132, 215, 254
relative angular, see Relative angular velocity basic dimensions 129, 130
Velocity transformation matrix 11, 32, 83, 119, conicity, see Conicity
122, 226, 251–253, 273, 279 disc 129
Vertical bump, see Deviations, bump flange 34, 36, 128–132
Vertical-development angle, see Development geometric details 129
angle geometry, see Geometry, wheel
Vertical displacement 33, 121, 268 hub 129
Vertical force 33, 34, 35 position and derivative equations 134
Vertical motion 24 profile, see Profile, wheel
Virtual work 31, 32, 225, 226–232, 239, 263, radii of curvature, see Curvature, radius of
264, 305 rolling radii, see Rolling, radii
applied forces 239, 241, 252, 307 structure 129
constraint forces 228, 251 surface 17–19, 53, 59, 63, 64, 68, 126–134,
contact forces 320 182
in dynamics 228, 306, 318 surface parameters, see Surface parameters,
inertia forces 97, 228, 232, 235, 239, 241, wheel
251, 304, 306, 307 tread, see Tread
372 Index

Wheel (contd.) hunting, see Hunting, wheelset


worn 18, 19, 59, 63, 125–128, 131–133, 178. instability, see Instability, wheelset
See also Profile curve, worn lateral displacement 3, 14, 201
Wheel/rail stability, see Stability, wheelset
constraints, see Constraint(s), wheel/rail suspended, see Suspended wheelset
contact formulations, see Contact two-degree-of-freedom model, see
formulation(s), wheel/rail Two-degree-of-freedom, wheelset
contact frame, see Contact frame, wheel/rail unsuspended, see Unsuspended
contact mechanism 177–183 wheelset
contact points, see Contact point, wheel/rail Worn wheels, see Wheel, worn
contact theories, see Contact theory,
wheel/rail y
creep forces, see Creep, force Yaw
flange contact, see Flange contact angle 3, 14–16, 33, 35, 89, 90, 101, 106, 120,
normal contact force, see Normal contact force 121, 198, 201, 233, 268, 272, 277, 284,
penetration, see Penetration, wheel/rail 285
separation 27, 35, 176, 180, 181, 183, 184, angular velocity 202
187, 216, 254, 255, 259–261 displacement 16, 31, 225, 277
Wheelset moment 284
Frame (coordinate system), see Coordinate motion 272
system, wheelset Young’s modulus of elasticity 296, 332

You might also like