You are on page 1of 862

Mechanical Engineering

“Overall, this is an excellent book and highly recommended. The coverage of the topics is wide
ranging, which makes it suitable for both undergraduate and graduate courses on dynamics. Baruh

APPLIED
What makes this book truly different from the rest are the applications of the dynamics prin-
ciples to real-world systems, such as vibrating systems and vehicles.”
—Ilhan Tuzcu, California State University, Sacramento, USA

APPLIED DYNAMICS
“The combination of applications with theory without compromising either one is excellently

DYNAMICS
done! Also, the unified and fresh approach to dynamics is excellent. …This book is like a breath
of fresh air…”
—Sorin Siegler, Drexel University, Philadelphia, Pennsylvania, USA

GAIN A GREATER UNDERSTANDING OF HOW KEY COMPONENTS WORK

Using realistic examples from everyday life, including sports (motion of balls in the air or during
impact) and vehicle motions, Applied Dynamics emphasizes the applications of dynamics in en-
gineering without sacrificing the fundamentals or rigor. The text provides a detailed analysis of the
principles of dynamics and vehicle motions analysis. An example included in the topic of collisions
is the famous “Immaculate Reception,” whose 40th anniversary was recently celebrated by the
Pittsburgh Steelers.

COVERS STABILITY AND RESPONSE ANALYSIS IN DEPTH

The book addresses two- and three-dimensional Newtonian mechanics, it covers analytical me-
chanics, and describes Lagrange’s and Kane’s equations. It also examines stability and response
analysis, and vibrations of dynamical systems. In addition, the text highlights a developing interest
in the industry—the dynamics and stability of land vehicles.

CONTAINS LOTS OF ILLUSTRATIVE EXAMPLES

In addition to the detailed coverage of dynamics applications, over 180 examples and nearly 600
problems richly illustrate the concepts developed in the text.

TOPICS COVERED INCLUDE


• General kinematics and kinetics
• Expanded study of two- and three-dimensional motion, as well as of impact dynamics
• Analytical mechanics, including Lagrange’s and Kane’s equations
• The stability and response of dynamical systems, including vibration analysis
• Dynamics and stability of ground vehicles

Designed for classroom instruction appealing to undergraduate and graduate students taking
intermediate and advanced dynamics courses, as well as vibration study and analysis of land ve-
hicles, Applied Dynamics can also be used as an up-to-date reference in engineering dynamics for

Haim Baruh
researchers and professional engineers.
K23786
6000 Broken Sound Parkway, NW
Suite 300, Boca Raton, FL 33487
711 Third Avenue
an informa business New York, NY 10017
2 Park Square, Milton Park
www.taylorandfrancisgroup.com Abingdon, Oxon OX14 4RN, UK
Applied
dynAmics
Applied
dynAmics
Haim Baruh

Boca Raton London New York

CRC Press is an imprint of the


Taylor & Francis Group, an informa business
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2015 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works


Version Date: 20140716

International Standard Book Number-13: 978-1-4822-5079-4 (eBook - PDF)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the valid-
ity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright
holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may
rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or uti-
lized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopy-
ing, microfilming, and recording, or in any information storage or retrieval system, without written permission from the
publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
In honor of my dear wife, Rahel Baruh.
Contents

Preface ix

Acknowledgments xiii

1 Introductory Concepts 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 What Is a System? . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Particles, Rigid Bodies, and Deformable Solids . . . . . . . . . . . . . . . . 2
1.3 Degrees of Freedom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3.1 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.2 Generalized Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Types of Forces and Motions . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.1 Vibratory Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.5 Systems of Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.6 Linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 Differential Equations and the Principle of Superposition . . . . . . . . . . 15
1.8 Dimensional Analysis and Nondimensionalization . . . . . . . . . . . . . . 17
1.9 Numerical Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.10 What Is a Vehicle? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.11 Cause-and-Effect Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.12 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.13 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2 Kinematics Fundamentals 25
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Position, Velocity, and Acceleration . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Reference Frames: Single Rotation in a Plane . . . . . . . . . . . . . . . . 26
2.4 Column Vector Representation . . . . . . . . . . . . . . . . . . . . . . . . 29
2.5 Commonly Used Coordinate Systems . . . . . . . . . . . . . . . . . . . . . 30
2.5.1 Rectilinear Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5.2 Normal-Tangential Coordinates . . . . . . . . . . . . . . . . . . . . 33
2.5.3 Cylindrical Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 38
2.5.4 Spherical Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.6 Moving Reference Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.7 Selection of Rotation Parameters . . . . . . . . . . . . . . . . . . . . . . . 48
2.7.1 Transformation by Three Rotation Angles . . . . . . . . . . . . . . 49
2.7.2 Resolving a Rotated Vector . . . . . . . . . . . . . . . . . . . . . . 51
2.7.3 Single Rotation about a Specified Axis . . . . . . . . . . . . . . . . 52
2.7.4 Finite Rotations Do Not Commute . . . . . . . . . . . . . . . . . . 54
2.8 Rate of Change of a Vector, Angular Velocity . . . . . . . . . . . . . . . . 58
2.8.1 Angular Velocity for Plane Motion . . . . . . . . . . . . . . . . . . 58
2.8.2 Angular Velocity for Three-Dimensional Motion . . . . . . . . . . . 59

vii
viii Contents

2.8.3 Other Definitions of Angular Velocity . . . . . . . . . . . . . . . . . 62


2.8.4 Additive Properties of Angular Velocity . . . . . . . . . . . . . . . 62
2.9 Angular Acceleration and Second Derivatives . . . . . . . . . . . . . . . . 64
2.9.1 Angular Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.9.2 Second Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.10 Relative Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.11 Instantaneous Center of Zero Velocity . . . . . . . . . . . . . . . . . . . . 73
2.12 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
2.13 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

3 Kinematics Applications 83
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.2 Motion with Respect to the Rotating Earth . . . . . . . . . . . . . . . . . 83
3.3 Contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.4 Rolling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4.1 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.4.2 Rolling Constraints, Wheel on an Axle . . . . . . . . . . . . . . . . 88
3.5 Bicycle Model of a Car . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.5.1 Where Is the Instant Center of a Car? . . . . . . . . . . . . . . . . 95
3.6 Kinematic Differential Equations . . . . . . . . . . . . . . . . . . . . . . . 98
3.7 Topspin and Backspin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.8 Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.8.1 Links and Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
3.8.2 Degrees of Freedom: Gruebler’s Equation . . . . . . . . . . . . . . . 102
3.8.3 Four-Bar Linkage and Slider-Crank Mechanism . . . . . . . . . . . 104
3.9 Instant Center Analysis for Linkages . . . . . . . . . . . . . . . . . . . . . 105
3.9.1 Locating Instant Centers in Linkages . . . . . . . . . . . . . . . . . 105
3.9.2 Velocity Analysis Using Instant Centers . . . . . . . . . . . . . . . 108
3.10 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.11 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

4 Kinetics Fundamentals 115


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.2 Rigid Body Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.2.1 Center of Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.2.2 Mass Moment of Inertia . . . . . . . . . . . . . . . . . . . . . . . . 118
4.3 Linear Momentum and Angular Momentum . . . . . . . . . . . . . . . . . 120
4.4 Resultant Force and Moment . . . . . . . . . . . . . . . . . . . . . . . . . 124
4.5 Laws of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.5.1 First Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.5.2 Second Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.5.3 Third Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
4.5.4 Inertia Forces and Inertia Moments . . . . . . . . . . . . . . . . . . 129
4.6 Forces and Moments Acting on Bodies . . . . . . . . . . . . . . . . . . . . 131
4.7 Force of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.8 Contact and Reaction Forces . . . . . . . . . . . . . . . . . . . . . . . . . . 133
4.9 Dry Friction Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.10 Aerodynamic Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.10.1 Lift Force and Drag Force . . . . . . . . . . . . . . . . . . . . . . . 140
4.10.2 Aerodynamic Coefficients . . . . . . . . . . . . . . . . . . . . . . . 141
4.10.3 Flow Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Contents ix

4.10.4 Drag Approximation for Very Low Reynolds Numbers . . . . . . . 145


4.11 Spring Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.11.1 Modeling of Springs . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.11.2 Equivalent Spring Constants . . . . . . . . . . . . . . . . . . . . . . 149
4.11.3 Stiffness Generating Components . . . . . . . . . . . . . . . . . . . 151
4.12 Dampers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.13 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

5 Kinetics Applications 165


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
5.2 Rolling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
5.3 Mechanical Trail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.4 Impulse and Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.4.1 Impulsive Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.4.2 Idealized Model of an Impulsive Force . . . . . . . . . . . . . . . . 174
5.5 Work, Energy, and Power . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
5.5.1 Kinetic Energy and Power . . . . . . . . . . . . . . . . . . . . . . . 175
5.5.2 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
5.5.3 Gravitational Potential Energy . . . . . . . . . . . . . . . . . . . . 179
5.5.4 Potential Energy of Springs . . . . . . . . . . . . . . . . . . . . . . 181
5.5.5 Work-Energy Relations . . . . . . . . . . . . . . . . . . . . . . . . . 182
5.5.6 Forces That Do No Work . . . . . . . . . . . . . . . . . . . . . . . 182
5.5.7 Hysteresis and Energy Loss . . . . . . . . . . . . . . . . . . . . . . 183
5.6 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
5.6.1 Deriving the Equation of Motion of Conservative One-Degree-of-
Freedom Systems Using Energy . . . . . . . . . . . . . . . . . . . . 188
5.7 Solution of the Equations of Motion . . . . . . . . . . . . . . . . . . . . . 189
5.8 Linearization, Equilibrium, and Stability . . . . . . . . . . . . . . . . . . . 190
5.8.1 Calculating the Equilibrium Position(s) . . . . . . . . . . . . . . . 191
5.8.2 Motion in the Vicinity of Equilibrium . . . . . . . . . . . . . . . . 191
5.8.3 Nature of the Response of the Linearized Equations . . . . . . . . . 193
5.8.4 Equations of Motion of Linear Systems about Equilibrium . . . . . 193
5.9 Motion in the Vicinity of the Earth . . . . . . . . . . . . . . . . . . . . . . 200
5.10 Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
5.10.1 Collisions of Particles . . . . . . . . . . . . . . . . . . . . . . . . . . 203
5.11 Impact of Rigid Bodies: Simple Solution . . . . . . . . . . . . . . . . . . . 208
5.12 A More Accurate Model of Rigid Body Impact . . . . . . . . . . . . . . . 214
5.12.1 Compression Stage . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
5.12.2 Restitution Stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
5.13 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
5.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221

6 Response of Dynamical Systems 231


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
6.2 The Unit Impulse and Unit Step Functions . . . . . . . . . . . . . . . . . . 231
6.2.1 The Unit Impulse Function . . . . . . . . . . . . . . . . . . . . . . 231
6.2.2 The Unit Step Function . . . . . . . . . . . . . . . . . . . . . . . . 232
6.3 Homogeneous Plus Particular Solution Approach . . . . . . . . . . . . . . 233
6.4 Laplace Transform Solution . . . . . . . . . . . . . . . . . . . . . . . . . . 235
6.4.1 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 235
x Contents

6.4.2 Solving Differential Equations Using the Laplace Transform . . . . 237


6.5 Response of First-Order Systems . . . . . . . . . . . . . . . . . . . . . . . 239
6.6 Review of Complex Variables . . . . . . . . . . . . . . . . . . . . . . . . . 243
6.7 Second-Order Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
6.8 Free Response of Undamped Second-Order Systems . . . . . . . . . . . . . 246
6.9 Free Response of Damped Second-Order Systems . . . . . . . . . . . . . . 250
6.10 Underdamped Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
6.11 Damping Estimation by Logarithmic Decrement . . . . . . . . . . . . . . . 255
6.12 Response to an Impulsive Force . . . . . . . . . . . . . . . . . . . . . . . . 258
6.12.1 Response to Multiple Impulses . . . . . . . . . . . . . . . . . . . . 261
6.13 Step Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
6.14 Response to General Excitations—Convolution Integral . . . . . . . . . . . 271
6.15 Time-Domain vs. Frequency-Domain Analysis . . . . . . . . . . . . . . . . 274
6.16 Response to Harmonic Excitation . . . . . . . . . . . . . . . . . . . . . . . 278
6.16.1 General Formulation for Second-Order Systems . . . . . . . . . . . 278
6.16.2 Quality Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
6.16.3 Phase Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
6.17 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
6.18 Transmitted Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
6.19 Base Excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
6.20 Harmonic Excitation Due to Imbalances and Eccentricity . . . . . . . . . . 290
6.20.1 Rotating Imbalances . . . . . . . . . . . . . . . . . . . . . . . . . . 290
6.20.2 Whirling of Rotating Shafts . . . . . . . . . . . . . . . . . . . . . . 291
6.21 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
6.22 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294

7 Response of Multi-Degrees-of-Freedom Systems 303


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
7.2 Modeling of Multi-Degrees-of-Freedom Systems . . . . . . . . . . . . . . . 303
7.2.1 Definition of Sign Definiteness . . . . . . . . . . . . . . . . . . . . . 304
7.3 Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
7.4 Free Motion of Undamped Multi-Degrees-of-Freedom Systems . . . . . . . 312
7.4.1 Natural Frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . 314
7.4.2 Modal Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
7.4.3 Normalization of Modal Vectors . . . . . . . . . . . . . . . . . . . . 318
7.4.4 General Form of the Free Response . . . . . . . . . . . . . . . . . . 320
7.5 Solving for the Natural Frequencies and Modal Vectors . . . . . . . . . . . 321
7.6 Beat Phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
7.7 Unrestrained Motion and Rigid Body Modes . . . . . . . . . . . . . . . . . 327
7.8 Orthogonality of the Modal Vectors . . . . . . . . . . . . . . . . . . . . . . 331
7.9 Expansion Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
7.9.1 Expansion by Sets of Functions . . . . . . . . . . . . . . . . . . . . 333
7.9.2 Expansion by Geometric Vectors . . . . . . . . . . . . . . . . . . . 334
7.9.3 Expansion by Modal Vectors . . . . . . . . . . . . . . . . . . . . . . 335
7.10 Modal Equations of Motion and Response . . . . . . . . . . . . . . . . . . 336
7.10.1 Modal Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
7.10.2 Modal Response Using Homogeneous and Particular Solution Ap-
proach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
7.10.3 Response of Modes with Zero Eigenvalues . . . . . . . . . . . . . . 340
7.10.4 Response to Impulsive Loading . . . . . . . . . . . . . . . . . . . . 341
7.11 Mode Participation and Isolation . . . . . . . . . . . . . . . . . . . . . . . 345
Contents xi

7.11.1 Mode Participation . . . . . . . . . . . . . . . . . . . . . . . . . . . 345


7.11.2 Mode Isolation—Mode Control . . . . . . . . . . . . . . . . . . . . 347
7.12 Approximate Approach for Damped Systems . . . . . . . . . . . . . . . . . 348
7.13 Response to Harmonic Excitation . . . . . . . . . . . . . . . . . . . . . . . 352
7.13.1 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 352
7.13.2 Special Case: Two-Degrees-of-Freedom Systems . . . . . . . . . . . 353
7.13.3 Modal Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
7.14 Vibration Reducing Devices . . . . . . . . . . . . . . . . . . . . . . . . . . 356
7.14.1 Motion Amplitude before Vibration Reducer Is Added . . . . . . . 357
7.14.2 Undamped Vibration Absorbers . . . . . . . . . . . . . . . . . . . . 358
7.14.3 Tuned Mass Dampers . . . . . . . . . . . . . . . . . . . . . . . . . 362
7.15 First-Order Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
7.16 Numerical Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
7.17 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
7.18 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368

8 Analytical Mechanics 375


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
8.2 Generalized Coordinates and Constraints . . . . . . . . . . . . . . . . . . . 375
8.2.1 Generalized Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 376
8.2.2 Constraints and Constraint Forces . . . . . . . . . . . . . . . . . . 377
8.3 Velocity Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
8.3.1 Representation by Generalized Velocities . . . . . . . . . . . . . . . 381
8.3.2 Representation by Generalized Speeds (Quasi-Velocities) . . . . . . 382
8.3.3 Relationship between Generalized Velocities and Generalized Speeds 382
8.4 Virtual Displacements and Virtual Work . . . . . . . . . . . . . . . . . . . 388
8.5 Virtual Displacements and Virtual Work for Rigid Bodies . . . . . . . . . 390
8.6 Generalized Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 394
8.7 Principle of Virtual Work for Static Equilibrium . . . . . . . . . . . . . . . 399
8.8 D’Alembert’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
8.8.1 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 401
8.8.2 Extension to Rigid Bodies . . . . . . . . . . . . . . . . . . . . . . . 402
8.8.3 Using D’Alembert’s Principle to Obtain Equations of Motion . . . 404
8.9 Hamilton’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
8.10 Lagrange’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
8.11 Constrained Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
8.11.1 Lagrange Multiplier Method . . . . . . . . . . . . . . . . . . . . . . 419
8.11.2 Constraint Relaxation Method . . . . . . . . . . . . . . . . . . . . 421
8.12 Kane’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
8.13 Natural and Nonnatural Systems, Equilibrium . . . . . . . . . . . . . . . . 426
8.14 Small Motions about Equilibrium . . . . . . . . . . . . . . . . . . . . . . . 431
8.15 Rayleigh’s Dissipation Function . . . . . . . . . . . . . . . . . . . . . . . . 437
8.16 Generalized Momentum, First Integrals . . . . . . . . . . . . . . . . . . . . 439
8.17 Impulsive Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
8.17.1 Impulsive Excitation in Lagrangian Mechanics . . . . . . . . . . . . 440
8.17.2 Impulse-Momentum Relationships for Kane’s Equations . . . . . . 441
8.18 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
8.19 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
xii Contents

9 Three-Dimensional Kinematics of Rigid Bodies 459


9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
9.2 Basic Kinematics of Rigid Bodies . . . . . . . . . . . . . . . . . . . . . . . 459
9.2.1 Pure Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
9.2.2 Pure Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
9.2.3 Combined Translation and Rotation . . . . . . . . . . . . . . . . . 465
9.3 Euler Angles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
9.3.1 Euler Angle Sequences . . . . . . . . . . . . . . . . . . . . . . . . . 467
9.3.2 Angular Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
9.3.3 Angular Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . 472
9.4 Axisymmetric Bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
9.5 Rolling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
9.5.1 Disk Originally Lying Flat . . . . . . . . . . . . . . . . . . . . . . . 480
9.5.2 Beginning with the Disk as Vertical . . . . . . . . . . . . . . . . . . 483
9.5.3 Steady Precession . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
9.6 Orientation Change by Successive Rotations . . . . . . . . . . . . . . . . 488
9.7 Interconnections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 490
9.7.1 Basic Types of Joints . . . . . . . . . . . . . . . . . . . . . . . . . . 490
9.7.2 Combined Sliding and Rotation . . . . . . . . . . . . . . . . . . . . 492
9.7.3 Universal Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
9.8 4×4 Matrix Description of a General Transformation . . . . . . . . . . . . 496
9.9 Euler Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
9.9.1 Transformation Matrix in Terms of Euler Parameters . . . . . . . . 501
9.9.2 Relating the Euler Parameters to Angular Velocities . . . . . . . . 502
9.9.3 Relating Euler Parameters to the Transformation Matrix . . . . . . 503
9.9.4 Relating Euler Parameters to the Euler Angles . . . . . . . . . . . 504
9.10 Rodrigues Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
9.10.1 Screw Algebra and Rodrigues Parameters . . . . . . . . . . . . . . 508
9.10.2 Rotation of a Body . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
9.10.3 Finding the Screw Axis and Rotation Angle . . . . . . . . . . . . . 513
9.11 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
9.12 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517

10 Mass Moments of Inertia 527


10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
10.2 Center of Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
10.3 Mass Moment of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . 528
10.3.1 Calculation of the Mass Moments and Products of Inertia . . . . . 530
10.4 Transformation Properties of the Inertia Matrix . . . . . . . . . . . . . . . 536
10.4.1 Translation of Coordinates . . . . . . . . . . . . . . . . . . . . . . . 536
10.4.2 Rotation of Coordinate Axes . . . . . . . . . . . . . . . . . . . . . 538
10.5 Principal Moments of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . 542
10.5.1 Special Case: Repeated Principal Moments of Inertia . . . . . . . . 543
10.6 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 546
10.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 546

11 Dynamics of Three-Dimensional Rigid Body Motion 553


11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553
11.2 Linear and Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . 553
11.3 Transformation Properties of Angular Momentum . . . . . . . . . . . . . . 557
11.3.1 Translation of Coordinates . . . . . . . . . . . . . . . . . . . . . . . 557
Contents xiii

11.3.2 Rotation of Coordinates . . . . . . . . . . . . . . . . . . . . . . . . 558


11.4 General Describing Equations . . . . . . . . . . . . . . . . . . . . . . . . . 562
11.4.1 Resultant Force and Moment . . . . . . . . . . . . . . . . . . . . . 563
11.4.2 Force and Moment Balances . . . . . . . . . . . . . . . . . . . . . . 564
11.4.3 Moment Balance about an Arbitrary Different Point . . . . . . . . 564
11.5 Description in Terms of Body-Fixed Coordinates . . . . . . . . . . . . . . 565
11.6 Angular Momentum Balance for Axisymmetric Bodies . . . . . . . . . . . 569
11.6.1 Modified Euler’s Equations . . . . . . . . . . . . . . . . . . . . . . 569
11.6.2 Expressing the Rotational Equations Using Components of Shape
Frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
11.7 Stability Analysis of Rotational Motion . . . . . . . . . . . . . . . . . . . . 577
11.8 Steady Precession of a Rolling Disk . . . . . . . . . . . . . . . . . . . . . . 579
11.9 Rotation about a Fixed Axis . . . . . . . . . . . . . . . . . . . . . . . . . . 584
11.10 Impulse and Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 588
11.11 Energy and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 590
11.11.1 Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 590
11.11.2 Work and Conservation of Energy . . . . . . . . . . . . . . . . . . . 592
11.12 Analytical Equations for Rigid Bodies . . . . . . . . . . . . . . . . . . . . 594
11.12.1 Euler Angles as Generalized Coordinates . . . . . . . . . . . . . . . 594
11.12.2 Lagrange’s and Kane’s Equations . . . . . . . . . . . . . . . . . . . 595
11.13 Torque-Free Motion of Axisymmetric Bodies . . . . . . . . . . . . . . . . . 602
11.13.1 Integrals of the Motion . . . . . . . . . . . . . . . . . . . . . . . . . 602
11.13.2 Body and Space Cones . . . . . . . . . . . . . . . . . . . . . . . . . 604
11.13.3 Direct and Retrograde Precession . . . . . . . . . . . . . . . . . . . 606
11.13.4 Energy Dissipation and Nutational Instability . . . . . . . . . . . . 607
11.14 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611
11.15 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611

12 Vehicle Dynamics—Basic Loads and Longitudinal Motions 623


12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 623
12.2 Vehicle Coordinate Systems and Nomenclature . . . . . . . . . . . . . . . 623
12.3 Loads on Vehicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
12.4 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
12.4.1 Acceleration Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 627
12.4.2 Maximum Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . 628
12.5 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
12.5.1 Constant Power Approximation . . . . . . . . . . . . . . . . . . . . 632
12.6 More Advanced Model Including Wheel Inertia . . . . . . . . . . . . . . . 634
12.7 Braking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637
12.7.1 Brake Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637
12.7.2 Force Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
12.7.3 Brake Proportioning . . . . . . . . . . . . . . . . . . . . . . . . . . 641
12.7.4 Anti-Lock Brake Systems . . . . . . . . . . . . . . . . . . . . . . . 644
12.8 Rollover and Lateral Instability . . . . . . . . . . . . . . . . . . . . . . . . 647
12.8.1 Simple Combined Lateral and Roll Analysis . . . . . . . . . . . . . 650
12.8.2 Critical Sliding Velocity . . . . . . . . . . . . . . . . . . . . . . . . 650
12.9 Weight Shift and Statical Indeterminacy . . . . . . . . . . . . . . . . . . . 656
12.9.1 Statically Indeterminate Systems . . . . . . . . . . . . . . . . . . . 656
12.9.2 An Approximate Method for Calculating Wheel Loads . . . . . . . 657
12.9.3 Calculation of Center of Mass Location . . . . . . . . . . . . . . . . 662
12.10 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
xiv Contents

12.11 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665

13 Vehicle Dynamics—Tire and Aerodynamic Forces 671


13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
13.2 Tires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
13.2.1 Tire Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
13.2.2 Tire Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
13.3 Tire Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 674
13.3.1 Resultant Tire Forces and Moments . . . . . . . . . . . . . . . . . 677
13.4 Lateral Forces and Tire Slip . . . . . . . . . . . . . . . . . . . . . . . . . . 678
13.5 Tire Torques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
13.5.1 Aligning Torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
13.5.2 Overturning Moment . . . . . . . . . . . . . . . . . . . . . . . . . . 685
13.6 Slip Ratio and Longitudinal Tire Forces . . . . . . . . . . . . . . . . . . . 685
13.6.1 Wheel Slip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
13.6.2 Tractive and Lateral Tire Forces . . . . . . . . . . . . . . . . . . . 686
13.6.3 More Comprehensive Tire Models . . . . . . . . . . . . . . . . . . . 687
13.7 Rolling Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
13.7.1 Factors Affecting Rolling Resistance . . . . . . . . . . . . . . . . . 691
13.7.2 Induced Drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692
13.7.3 Rolling Resistance Models . . . . . . . . . . . . . . . . . . . . . . . 693
13.8 Camber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694
13.9 Other Tire Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
13.9.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
13.9.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
13.9.3 Speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
13.9.4 Conicity and Ply Steer . . . . . . . . . . . . . . . . . . . . . . . . . 697
13.10 Summary of Tire Force Effects . . . . . . . . . . . . . . . . . . . . . . . . 698
13.11 Nondimensional Analysis of Tire Behavior . . . . . . . . . . . . . . . . . . 699
13.12 Aerodynamic Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700
13.12.1 Calculation of Aerodynamic Coefficients from Test Data . . . . . . 704
13.13 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 706
13.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707

14 Vehicle Dynamics—Lateral Stability 711


14.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 711
14.2 Kinematics—Steer Angle Definitions . . . . . . . . . . . . . . . . . . . . . 711
14.3 Wheel Loads and Slip Angles . . . . . . . . . . . . . . . . . . . . . . . . . 717
14.3.1 Free-Body Diagram and Slip Angles . . . . . . . . . . . . . . . . . 717
14.3.2 Understeer Gradient and Critical Speed . . . . . . . . . . . . . . . 720
14.3.3 Bundorf Compliances . . . . . . . . . . . . . . . . . . . . . . . . . . 723
14.3.4 Lateral Acceleration Gain . . . . . . . . . . . . . . . . . . . . . . . 723
14.4 Slip Angle Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725
14.5 Transient Motion Equations . . . . . . . . . . . . . . . . . . . . . . . . . . 728
14.6 Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731
14.6.1 Numerical Integration of the Stability Equations . . . . . . . . . . 732
14.7 Eigenvalue Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735
14.8 Mass-Spring-Damper Analogy . . . . . . . . . . . . . . . . . . . . . . . . . 739
14.9 Steady-State Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
14.10 Yaw Velocity Gain and Curvature Response . . . . . . . . . . . . . . . . . 745
14.11 Tangent Speed and Hydroplaning . . . . . . . . . . . . . . . . . . . . . . . 746
Contents xv

14.12 Neutral Steer Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 748


14.13 Modeling the Driver . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
14.14 Electronic Stability Control . . . . . . . . . . . . . . . . . . . . . . . . . . 751
14.15 Which Wheels Will Slide First? . . . . . . . . . . . . . . . . . . . . . . . . 756
14.16 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 757
14.17 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 758

15 Vehicle Dynamics—Bounce, Pitch, and Roll 763


15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 763
15.2 Sources of Excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 764
15.2.1 Ride Quality and Human Response to Vibration . . . . . . . . . . 765
15.3 Unsprung vs. Sprung Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . 766
15.4 Simple Suspension Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 766
15.5 Quarter-Car Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 770
15.5.1 Single-Degree-of-Freedom Model . . . . . . . . . . . . . . . . . . . 770
15.5.2 Change in Natural Frequency and Damping Factor Due to Payload
Weight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774
15.5.3 Wheel Hop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774
15.5.4 Two-Degrees-of-Freedom Model . . . . . . . . . . . . . . . . . . . . 775
15.6 Pitch and Bounce Motions . . . . . . . . . . . . . . . . . . . . . . . . . . . 778
15.6.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 778
15.6.2 Analysis of the Equations of Motion and Response . . . . . . . . . 780
15.7 Olley Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785
15.8 Response to Harmonic Excitation . . . . . . . . . . . . . . . . . . . . . . . 786
15.8.1 Modeling an Uneven Road . . . . . . . . . . . . . . . . . . . . . . . 786
15.8.2 Frequency Response for a Single-Degree-of-Freedom Model . . . . . 788
15.8.3 Frequency Response for Two-Degrees-of-Freedom Models . . . . . . 791
15.9 Roll Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794
15.10 Roll Center Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 798
15.11 Lateral Force Reduction Due to Weight Shift . . . . . . . . . . . . . . . . 801
15.12 Roll Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 803
15.12.1 Design for Cornering . . . . . . . . . . . . . . . . . . . . . . . . . . 806
15.13 Introduction to Suspension Systems . . . . . . . . . . . . . . . . . . . . . . 807
15.14 Suspension System Terminology and Geometry . . . . . . . . . . . . . . . 807
15.15 Axle Suspensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 811
15.16 Independent Suspensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814
15.17 Roll Center Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 817
15.18 Jacking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 820
15.19 Scrub . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821
15.20 Anti-Roll Bar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 822
15.21 Force Analysis for Anti-Squat and Anti-Dive . . . . . . . . . . . . . . . . . 824
15.21.1 Independent Suspensions and Rear Wheel Drive . . . . . . . . . . . 828
15.21.2 Front Wheel Drive . . . . . . . . . . . . . . . . . . . . . . . . . . . 828
15.21.3 Four Wheel Drive . . . . . . . . . . . . . . . . . . . . . . . . . . . . 829
15.21.4 Anti-Dive for Braking . . . . . . . . . . . . . . . . . . . . . . . . . 829
15.22 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 830
15.23 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 831

16 Appendix—Common Inertia Properties 835

Index 839
Preface

Introduction
I have been involved with research in and the teaching of dynamics and its applications for
over 30 years. While I have had the pleasure of studying and teaching from several excellent
and rigorous texts, I have always wanted to develop material to teach dynamics and its
applications where the examples are primarily from real-world situations.
Comments from students and from users of my first text, Analytical Dynamics, pointed
to the need to develop realistic examples with realistic parameters that readers could relate
to. Over the years, after developing and also locating in the literature several such examples,
I decided to take the plunge and to organize the subject material into a book. The basic
concepts from dynamics are presented within the context of applications. I have included
real-world examples including the motion of land vehicles and sports.
While there is an emphasis on applications in the book, no effort has been made to
reduce rigor or to reduce emphasis on the fundamentals. I always stress to my students
that it is important to know both the fundamentals and applications. You should be just
as comfortable working with nondimensional parameters as with realistic numbers. You
should be able to solve the describing equations, and also have a sense of the magnitudes
of the parameters involved. Always ask yourself, after solving a problem, if the answer you
obtained makes sense.

Outline
This book can be used in an advanced undergraduate course on dynamics, a graduate course
on dynamics, an undergraduate course on vibrations, or a (graduate or undergraduate level)
course on land vehicle dynamics.
Chapter 1 is an introductory chapter that discusses the concepts of degrees-of-freedom,
generalized coordinates, constraints, systems of units, linearization, dimensional analysis,
and the cause and effect principle. Numerical integration of differential equations, which is
of critical importance when dealing with applications, is also discussed.
Chapter 2 is the first of three chapters on kinematics. A thorough analysis of kinematics
of a problem is essential before studying kinetics. The chapter discusses concepts such as
position, velocity and acceleration, coordinate frames and rotations, definition of angular
velocity and angular acceleration, relative motion, and instant centers.
Chapter 3 presents several applications of kinematics. The topic of contact is discussed,
and an analysis of rolling follows. Applications from vehicle dynamics, such as bicycle model
and Ackermann steering, are discussed. Kinematic differential equations, which are very use-
ful for numerical work, are introduced. The chapter continues with the study of mechanisms
and with an expansion of the definition of instant centers for linkages. In a land vehicle, the
steering and suspension systems are modeled and built as mechanisms.
Chapter 4 discusses the basics of kinetics. The concepts of linear and angular momentum
are reviewed. The laws of motion for translational and rotational motion are presented in a

xvii
xviii Applied Dynamics

slightly different way from traditional approaches. The chapter continues with the several
different types of forces that act on bodies: gravity, contact forces, reaction forces, friction,
aerodynamic, and hydrodynamic. Spring forces are introduced and compliance in mechanical
systems is modeled. Viscous damping forces and modeling of damping are discussed.
Chapter 5 presents additional concepts from kinetics, as well as applications. The dynam-
ics of rolling is discussed and wheels of vehicles are considered. The concept of mechanical
trail, which is widely used in cars, bicycles, and casters, is introduced. Energy and mo-
mentum are discussed. The concept of equation of motion is introduced. Equilibrium and
linearization of equations of motion about equilibrium are presented. Simple concepts from
stability theory are discussed. The chapter also presents an expanded discussion of impact
and impact with friction.
While the basic relationships presented in Chapters 1–5 are valid for two- or three-
dimensional motion, most of the examples and applications are from plane motion.
After studying the modeling of dynamical systems, the next step is to solve the describing
equations and to analyze how the motion evolves. Chapters 6 and 7 do that. Chapter 6 deals
with systems with one degree of freedom, and Chapter 7 extends these concepts to multi-
degrees-of-freedom systems. Both chapters deal with linear or linearized systems and make
use of concepts from vibration theory. Some simple nonlinear analysis and stability are
included.
Chapter 6 begins with a review of methods of solution for linear ordinary differential
equations: the homogeneous plus particular solution method and the Laplace transform
method. The free motion of undamped and damped systems is considered. Natural frequen-
cies and damping factors are introduced. The general system response is discussed. The
chapter continues with harmonic excitation. The concepts of resonance and transmitted
force are introduced, both of significant importance to vehicle and other types of motion.
Chapter 7 extends the concepts in Chapter 6 to multi-degrees-of-freedom systems. It
introduces the matrix formulation of multi-degrees-of-freedom systems and demonstrates
existence of multiple natural frequencies. Concepts from linear algebra are used to solve
the matrix eigenvalue problem and to obtain the natural frequencies and modal vectors.
Applications such as mode participation, mode control, harmonic excitation, and vibration
reducing devices are discussed. A simple way of modeling damping is presented.
Chapter 8 presents techniques for obtaining the describing equations of dynamical sys-
tems that are based on energy and momentum principles. These approaches are useful
when the number of degrees of freedom is large and in the presence of nonholonomic con-
straints. A land vehicle, for example, is a nonholonomic system. The principle of virtual
work and D’Alembert’s principles are derived. These principles lead to two powerful meth-
ods: Lagrange’s equations and Kane’s equations (also known as Gibbs-Appell equations).
Also discussed are applications of analytical techniques, such as equilibrium, linearization,
damping modeling, and impulsive motion.
Chapter 9 extends the kinematics concepts in Chapters 2 and 3 to three-dimensional
motion. The Euler angles are introduced. Choices in selecting moving coordinate systems are
presented. The motions of axisymmetric bodies and rolling are discussed. Interconnections
in three dimensions are presented. The chapter concludes with the Euler and Rodrigues
parameters, which are additional parameters to describe rotational motion.
Chapter 10 discusses the center of mass and mass moments of inertia of rigid bodies.
For three-dimensional motion, the inertia matrix of a body needs to be calculated. Se-
lection of coordinate systems for calculating the inertia properties is discussed, as well as
transformation of coordinate systems and the concept of principal moments of inertia.
Chapter 11 deals with the kinetics of three-dimensional motion and applies the concepts
discussed in Chapters 4 and 5, as well as of analytical mechanics in Chapter 8, to three
dimensions. Applications of these equations for different types of bodies and by using differ-
Preface xix

ent coordinate systems are discussed. The very important cases of axisymmetric bodies and
rolling are analyzed. The concepts of energy and momentum are revisited, and the stability
of rotational motion, especially the added stability properties of axisymmetric bodies, is
discussed.
The next four chapters are dedicated to the modeling and analysis of land vehicles. These
chapters present applications as well as new developments. Some of the developments can
be studied as examples to concepts covered earlier. For example, suspension systems in
Chapter 15 serve as excellent examples for the study of single and multi-degrees-of-freedom
systems in Chapters 6 and 7. The study of acceleration and braking in Chapter 12 is an
interesting application of Newton’s Second Law in Chapter 4. The roll center of a vehicle,
discussed in Chapter 15, serves as an excellent example of mechanisms in Chapter 3.
Chapter 12 applies the basic laws of motion to longitudinal motions of vehicles. Vehicle
coordinate systems are introduced. Acceleration, braking, and their limit values are dis-
cussed. Wheel loads and their statical indeterminacy are analyzed. Power calculations are
presented. A simple model of rollover is considered and analyzed in parallel with lateral
loads that result from taking a sharp turn.
Chapter 13 considers the interaction between tires and the road surface, and the aero-
dynamic forces that act on vehicles. The various tire forces that are generated as a vehicle
moves are described, together with the resultant torques generated on the contact area be-
tween the tires and the road. The concept of tire slip and the lateral loads generated by tire
slip are crucial to the overall stability and performance of vehicles.
Chapter 14 analyzes the lateral motions of a vehicle, which consist of sideslip and yaw.
The lateral stability of the vehicle is analyzed at several levels: transient motion, steady-state
motion, and eigenvalue analysis. Relationships between steer angles and slip are developed,
leading to definitions of understeer and oversteer. Electronic stability control is described
and a simple mathematical model of the driver is presented.
Chapter 15 considers the pitch, bounce, and roll motions of a vehicle. Suspension models
and geometry are introduced. Study of suspensions requires concepts from vibration theory
in Chapters 6 and 7. This analysis is extended to roll motion. After the study of pitch, roll,
and yaw, the chapter considers the analysis and design of suspension systems, which are
modeled as mechanisms. Independent suspensions, such as MacPherson strut and short-long
arm, are discussed. The roll center is introduced.

Chapter Dynamics Dynamics Vibrations Vehicle


(Undergraduate) (Graduate) Dynamics
1 1–6, 8–9 1–11 1–7, 9 1–6, 9–10
2 1–11 1–4, 6–11 1–2 1–2, 5, 8–11
3 1–5, 8–9, 1–6 1, 3–6, 8–9
4 1–12 1–9 1–9, 11–12 1–12
5 1–2, 5–11 1–7 1, 4–8 1–5, 10
6 1–20 Use as reference
7 1–14, 16 Use as reference
8 1–10, 12 1–17 1, 10
9 1–5, 7–9 1–6, 9–10
10 1–5 1–5
11 1–6, 8–9 1–13
12 4, 5, 7 1–9
13 1–3 1–12
14 1–15
15 1, 6 1–6 1–17
xx Applied Dynamics

The table above provides suggestions to educators on the sections of the book to use for
different courses that can be taught using this book. I am interested in learning different
orders or sequences of topics that you may have used. Please contact me and let me know
your comments about the book. Also, please feel free to inform me if you spot a typo or
other error. I am creating an errata sheet. My e-mail addresses are baruh@jove.rutgers.edu,
baruh@rci.rutgers.edu, and babahaim@gmail.com.
Acknowledgments

Writing a book is a lengthy and difficult task and I wish to acknowledge the support of
many people who have taught me, helped me, and guided me. First, I am richly indebted
to all the teachers I have had in my life, from my elementary grade teacher in Istanbul,
Turkey, Mrs. Cahide Şen, to my advisor in graduate school, Dr. Leonard Meirovitch. My
employer for over 30 years, Rutgers University, Department of Mechanical and Aerospace
Engineering, has given me the time, resources, and the collegial atmosphere to conduct
scholarly work.
I am grateful to Jonathan Plant, senior editor at CRC Press, for trusting me with this
book and for his guidance during the publication process. It has been a pleasure to work
with him and the staff at CRC. Robin-Lloyd Starkes was very helpful with editorial tasks
and Marcus Fontaine provided excellent technical support. I also thank the reviewers of the
manuscript, Dr. Nilanjan Sarkar and three anonymous reviewers, who all made valuable
suggestions. My Rutgers colleagues Haym Benaroya and William J. Bottega were very
supportive and they encouraged me to publish with CRC.
When I started teaching vehicle dynamics, Dr. Tom Gillespie was very helpful, sharing
his notes and giving me insight. Doug Milliken and his late father William F. Milliken, Sr.,
have been giants of the vehicle dynamics community. I have referred to the excellent books
written by these people repeatedly, as well as to the excellent book by Reza Jazar.
I thank all the students who used the manuscript in their courses and made valuable
suggestions, as well as the following students who helped with the figures: Saugata Dutt,
Qiming Guan, Haqiqat Kalirao, Jubilee Prasad Rao, Kathleen Sindoni, Chen Yang, Wanjun
Yang, and Yicheng Zhang. All brought a different style and perspective to the figures, which
you can observe by looking at the different chapters. I am indebted to Kathleen Sindoni for
initiating the drawing project and for generating templates. I enthusiastically thank Qiming
Guan for the solutions manual and for proofreading. I am indebted to Jubilee Prasad Rao
for making corrections to the figures.
Few people are blessed to have parents like mine, Yasef and Ester Baruh, obm. They
were two people who did not eat but fed their families, who wore old, threadbare clothes
so their children could wear new clothing, who lived extremely frugally so they could send
their children to the finest schools. These two compassionate and kind human beings were
beacons of honesty and integrity in their communities and they were loved, admired, and
respected by all around them. They were natural leaders. They were true role models for
everyone. I miss them dearly.
I am very grateful to my dear uncle, Selim Baruh, M.D., obm, and his wife Elena Baruh,
M.D., who have been second parents to me all my life. They have inspired and guided me,
and their generosity made it possible for me to attend graduate school. I am blessed to have
my three wonderful children, Esther, Naomi, and Joey, as well as my son-in-law Pinchas
Fink and grandson Danny Fink. They all give me purpose in life.
I regard the seminal event of my life to be December 31, 1985, when I was invited to
a New Year’s party and met my wife to be, Rahel Baruh (née Abrams). Rahel molded me
into a better person, husband, father, and human being. I owe much of my personal and
professional development to her, a brilliant woman who knows what the important things in

xxi
xxii Applied Dynamics

life really are, a woman who works harder than anyone I know, a beautiful person unfailingly
devoted to her family and community. She also helped edit this book, even though she does
not have a science or engineering background. It is in her honor that I dedicate this book.
I also thank you, dear readers, who have chosen to read the book. Please feel free to let
me know your comments and how I can make future editions more readable and useful. My
contact information is in the Preface.
1
Introductory Concepts

1.1 Introduction
This chapter discusses introductory concepts for the study of motion. We classify objects
as particles, rigid bodies, and deformable bodies. We discuss systems of units, degrees of
freedom, constraints, linearization, dimensional analysis, and the cause-and-effect principle.
The chapter presents a broad classification of forces to which dynamical systems are sub-
jected, as well as a preliminary discussion on two problems of special interest to practicing
engineers and scientists: vibrations and vehicle motion.

1.1.1 What Is a System?


A term frequently used in dynamics is system. A system can be loosely defined as a collection
of bodies or components acting together as one. We also use the word system to describe the
behavior of a body with respect to its environment. We can choose to analyze a system as a
whole or we can split it into a series of subsystems and analyze each subsystem individually.
When an input (or excitation) is applied to a system, a certain output (or response) is
obtained, as illustrated in Figure 1.1.

FIGURE 1.1
Schematic of a system.

For example, in the airplane in Figure 1.2, the plane and its control surfaces are the
system, while the lift, drag, thrust, and weight forces comprise the input. The motions of
the airplane, such as translations in the horizontal and vertical directions, rotations pitch,
roll and yaw, and vibration of the wings and tail are the outputs. An airplane is what is
called a multi-input multi-output system.
To understand the behavior of a system, we need to develop its mathematical model.
This process involves tools from mechanics as well as mathematics. Often, we make several
assumptions, such as neglecting part of the motion, or linearization, or assuming a certain
type of energy dissipation.
It is crucial to be aware of what system we are analyzing and isolating, as well as
what assumptions we are making, and to model the interactions between the system under
consideration with its environment and with other systems.

1
2 Applied Dynamics






FIGURE 1.2
An airplane.

1.2 Particles, Rigid Bodies, and Deformable Solids


When analyzing the motion of a body, the first task is to identify the type of body. Many
times, we make assumptions regarding the type of body and we develop simplified models.1
We will consider three types of bodies: particles, rigid bodies, and deformable bodies.
All three are constant mass bodies. Mass is defined as the amount of matter contained in
a body. It can also be viewed as the resistance of a body to translational motion. Mass is
an inertial quantity, and except for relativistic mechanics problems (which involve speeds
nearing the speed of light), it does not change.
A particle is defined as a body with no physical dimensions or a body whose entire
mass is concentrated at one point. When studying particle motion the interest is in trans-
lational motion, as we cannot attribute any rotational motion to a point mass. The particle
assumption is also used when the rotational motion of a body is very small compared to its
translational motion. For example, when analyzing an airplane moving from one airport to
another thousands of miles away, the interest is in the trajectory and altitude of the air-
plane and we treat the airplane as a particle. The pilot, on the other hand, is also concerned
with the attitude (orientation) of the airplane and treats the plane not as a particle but as
a rigid body. The pilot maneuvers the airplane so that the passengers experience as little
discomfort as possible. An object may be moving along a curved path and still exhibit only
translational motion.
A rigid body is defined as a body that has physical dimensions, hence, a shape. The shape
of a rigid body does not change when forces are applied to it. That is, the distance between
any two points on the body does not change under the action of forces and moments. The
motion of a rigid body consists of translations as well as rotations.
A deformable body deforms when forces are applied to it. Such a body has elasticity,
also known as flexibility or compliance. The body may or may not return to its original
undeformed shape when the load or force is removed. Fluids, which take the shape of their
containers, are not included among deformable bodies and are classified differently. So are
variable mass systems, such as a rocket.
Considering a body as rigid assumes that it has no flexibility. The validity of such an
assumption should be continuously evaluated. The following episode from space mechanics
is a case in point. The Explorer satellites (Explorer I and III were launched in 1958) were
each in the shape of an axisymmetric rod, with protruding thin antennae, as shown in Figure
1.3. Such satellites are sent into space with an initial spin to give them attitude stability, the
same way one throws an American football. The energy dissipation due to damping of the
1 One of the most important steps in analysis is to make accurate and valid assumptions. We must

continuously monitor and assess the validity of our assumptions.


Introductory Concepts 3

antennae was not included in the mathematical model. The unmodeled damping properties
of the antennae dissipated energy, which caused the spin rate to slow down and nutational
instabilities to occur.2

FIGURE 1.3
Model of Explorer 1 satellite and its antennae at the Smithsonian Air and Space Museum.
Source: Wikimedia Commons, http://commons.wikimedia.org/wiki/File:Explorer 1 a -
Smithsonian Air and Space Museum - 2012-05-15.jpg (Last accessed August 21, 2014).

1.3 Degrees of Freedom


Development of a mathematical model requires analysis of the different actual motions that
are possible, as well as the amount of simplification that we wish to make. The term degree
of freedom (d.o.f.) is defined as the minimum number of independent variables required to
describe the motion completely. The net result of possible different actual motions minus
restrictions on the motion minus the simplifications that are made gives the number of
degrees of freedom of the system.
After determining the number of different ways a system can move, each different motion
needs to be described by a motion variable, such as position and angle. A set of independent
variables that can describe the motion of a body completely is called generalized coordinates.
In general, there are as many generalized coordinates as degrees of freedom. An exception
is nonholonomic systems, as will be demonstrated later.
Consider a bead sliding inside a rotating hoop of radius R, as shown in Figure 1.4. The
position of the bead with respect to the hoop can be described by the angle θ that the line
joining the center of the hoop and the bead makes with a line that is fixed with respect to
the hoop.
2 Chapter 11 discusses nutational instabilities that occur when there is energy dissipation in slender

bodies.
4 Applied Dynamics


FIGURE 1.4
Bead sliding inside a rotating hoop.

The number of degrees of freedom of this system depends on how we treat the rotation
of the hoop. If the rotation of the hoop is a specified quantity (for example, if there is a
servomotor regulating the angular velocity Ω), then the motion of the hoop is known and
it is not a motion variable. On the other hand, if the hoop is free to rotate (no motor
to regulate its angular velocity), then the system has two degrees of freedom. A second
generalized coordinate is needed to model the rotation of the hoop, say, φ, so that Ω = φ̇.
A particle in general motion has three degrees of freedom. A rigid body in space has six.
A deformable body has, in theory, an infinite number of degrees of freedom.

a) b)
Y Y

G
B
y yG
A
x X xG X

FIGURE 1.5
Objects on the XY plane: a) particle, b) rigid body.

A particle restricted to moving on a plane has two degrees of freedom and a rigid body in
plane motion has three degrees of freedom, as shown in Figure 1.5. For the block in Figure
1.5b we can use the coordinates XG and YG to denote the position of point G on the body
and angle θ to denote orientation of the body. Note that in defining the angle θ, a line is
drawn between two fixed points on the body A and B. The line joining A and B does not
change with respect to the body as the body moves.
One of the initial tasks when studying motion is the determination of the number of
degrees of freedom. In general, there are no given guidelines on how to determine the number
of degrees of freedom. The general relationship
d.o.f. = no. of possible motions − no. of restraints − no. of simplifications
Introductory Concepts 5

is a useful guide. Most of the time, we need to make reasonable assumptions regarding the
number of degrees of freedom.

1.3.1 Constraints
Parallel to the concept of degree of freedom, it is useful to consider the concept of a degree
of restraint (d.o.r.), which is defined as the number of possible motions minus the number
of d.o.f. Consider the door hinge in Figure 1.6a. The hinge is attached on one side to the
door frame and on the other side to the door. If the hinge is built well and firmly attached
to the door and the door frame, a one-degree-of-freedom system is the outcome, with the
hinge permitting the opening and closing of the door by rotation about the hinge axis.
The hinge provides five degrees of restraint (or five constraints) to the door, three trans-
lational and two rotational, as shown in Figure 1.6b. These constraints on the motion are
enforced by the forces in three directions and moments in two directions generated by the
hinge mechanism and screws connecting the hinge to the door and the frame.

FIGURE 1.6
a) Hinged door, b) forces on hinge.

It is important to have the supports and components of that provide restraints remain
as restraints during the lifetime of the system. If the restraints fail, the resulting system
will possess additional degrees of freedom. In general, we design mechanisms and machines
so that they have as many controllers as there are degrees of freedom. If a system gains
additional degrees of freedom due to failure of a restraint, it becomes impossible (or very
hard) to control those additional degrees of freedom. For example, when the screws con-
necting the hinge to the door or frame become loose, the hinge begins losing its degrees of
restraint and the door acquires undesirable motions, such as twisting or translating. Such
undesired motions make it harder and less accurate to open and close the door, and they
put additional stresses on the hinge.
When a car tire gets worn out and cannot hold the road, or a driver takes a turn at a
high speed, the car may start sliding. It thus loses a d.o.r., which is equivalent to adding a
degree of freedom. The vehicle is now uncontrollable.3 All the driver can do is to wait for
friction to eventually overcome the sliding. There are evasive measures the driver can take
to minimize the time it takes for the sliding to end.
3 One method that police use to destabilize a fleeing vehicle is to give it a small bump from the side

closest to the rear. Known as the PIT (Precision Immobilization Technique) maneuver, this action induces
a sliding motion to the fleeing vehicle. Sliding in a fast moving vehicle does not come to an end quickly. We
will discuss this loss of stability in Chapter 12.
6 Applied Dynamics

A restraint or a constraint is the kinematical description of a restriction on the motion


of a body or of a system of bodies. For the constraint to be enforced, there has to be an
accompanying constraint force. This duality will be observed later on in this chapter, when
discussing the cause-and-effect principle, as well as in subsequent chapters.
Consider the free-body diagram of a vehicle in plane motion in Figure 1.7. The vehicle
consists of the body, two rear wheels on an axle and the steering in the front simplified into
a single wheel, like that of a tricycle.4

 


 






 



FIGURE 1.7
Simple model of a vehicle (tricycle model).

The vehicle body has three degrees of freedom (rigid body in plane motion), and the
steer angle δ is an additional motion variable, for a total of four d.o.f. Assuming that friction
is sufficient to prevent sliding of the wheels, each wheel has a velocity in the direction of its
heading. The constraints of the rear wheels can be combined by looking at the motion of
the midpoint of the rear axle, point A. The kinematic relationships that ensue are vA = vA i
and vB = vB i0 , and the two associated constraints are

vA · j = 0 vB · j0 = 0 (1.1)

When all wheels roll without slipping or sliding, there are 4 − 2 = 2 degrees of free-
dom. Forces that act on the vehicle are the two constraint forces5 RA and RB , which are
perpendicular to the velocity of the rear and front wheels, respectively, as well as the two
propulsive forces F1 and F2 .
In a car, the propulsion and braking mechanisms control speed, while the steering wheel
controls direction. Hence, there are two degrees of freedom and two controls.
When discussing degrees of freedom and constraints, we need to distinguish between
constraints that affect position (and velocity) and constraints that affect velocity only. This
issue will be discussed in more detail in Chapters 2, 9, and 11. In general, when we discuss
degrees of freedom, we refer to degrees of freedom associated with velocity.
For the vehicle in Figure 1.7, four coordinates are necessary to describe the position,
orientation, and steer of the vehicle: the two coordinates of the center of mass, XG and YG ,
the orientation angle θ, and the steer angle δ. To describe the vehicle velocity and angular
velocity, only two velocity variables are needed, vA and θ̇. This is because the constraints
4 We will see the reasoning for this simplification in Chapter 2.
5R represents the sum of the constraint forces on the rear wheels.
A
Introductory Concepts 7

discussed above (vA · j = 0 and vB · j0 = 0) constrain only the vehicle velocity but not the
position or orientation of the vehicle. Such constraints are known as nonholonomic.

1.3.2 Generalized Coordinates


After determining the number of degrees of freedom, the next step is selection of the motion
variables that are needed to uniquely describe the motion. The set of position and orientation
variables that describe a system are known as generalized coordinates. Selection of motion
variables can be a complex task and there are no strict guidelines, other than maintaining
the uniqueness of the coordinates and simplifying the description of motion. Experience
also helps. The coordinate systems discussed in the next chapter are different choices of
generalized coordinates. The study of generalized coordinates continues throughout the
text, especially when analytical mechanics is discussed in Chapter 8.
It is important to avoid ambiguities when selecting the motion variables. For example,
the simple pendulum of fixed length L in Figure 1.8 has one degree of freedom as long as
the wire is taut. The obvious choice for the generalized coordinate is the angle θ, as any
position of the pendulum can uniquely be described by θ. The other possibilities for the
motion variable are the x or z coordinates of the pendulum. The z coordinate (z = L cos θ)
is ambiguous because, for a given value of z, there are two possible orientations for the
pendulum. Using x as the generalized coordinate (x = L sin θ) does not create ambiguities,
but it complicates the description of the position and velocity.

z L


FIGURE 1.8
Simple pendulum.

Example 1.1
Determine the number of degrees of freedom for the system shown in Figure 1.9.
The cart has one degree of freedom, and because the block of mass 2m is attached to it
with a wire of fixed length, the block does not add an additional degree of freedom as long
as the wire is taut. The block of mass m is connected to the upper block by a spring, so it
has independent motion. Hence, the degrees of freedom for this system are two, assuming
that the wire between the cart and the block of mass 2m is always taut.
We can look at this problem also by first counting the bodies that can move: three
bodies, all in rectilinear motion, for a total of three motions. Next we consider that the
8 Applied Dynamics






FIGURE 1.9
Masses with a pulley.

wire connecting the cart to the upper block provides one constraint. Therefore, the d.o.f.
= 3 − 1 = 2.

Example 1.2
How many degrees of freedom does the trailer in Figure 1.10 have? Assume that none of
the wheels slide.











FIGURE 1.10
A trailer.

The answer has to be two, as the trailer is navigated by two controls: 1) propulsive
and braking forces to change speed and 2) the steering to change direction. The tractor
(powered vehicle) has two degrees of freedom, as established in the previous section. The
trailer, when not connected to the powered front vehicle, has three degrees of freedom. The
trailer has wheels in the back, which results in one constraint (again, we assume the rear
wheels do not slide).
The front of the trailer is connected to the tractor by a hitch which, in essence, is a
ball-and-socket joint. In the plane of motion (road surface), the hitch behaves as a pin
joint. The pin joint permits the trailer to rotate about the hitch, but it does not permit
any translation in the plane of motion. The pin joint imposes two constraints. Hence, while
the trailer adds three possible motions, the no sliding condition of the rear wheels and the
Introductory Concepts 9

no translation relative to the tractor at the hitch provide three constraints, resulting in a
system that has 2 (tractor) + 3 (trailer) − 3 (wheels and pin joint) = 2 degrees of freedom.

1.4 Types of Forces and Motions


The forces acting on a body, or on a system of bodies, dictate the nature of the motion.
We will analyze commonly encountered forces in Chapter 4. Here, we present a broad
classification of forces that dynamical systems encounter, followed by the general nature of
the ensuing motion.

• Aerodynamic or hydrodynamic forces. These forces act on a body that is moving


in air or water or in another fluid. Such forces are primarily resistive forces, except for
lift forces in aerodynamics.

• Gravity. The force of gravity is the weight of a body. For a rigid body, the gravity force
acts through the center of mass.
• Restoring forces. Restoring forces always act against the motion of a body and have
the effect of bringing back the body to its original position. The presence of restoring
forces causes repetitive or vibratory motion. Springs generate restoring forces that are
functions of the displacement, and they do not dissipate energy (or dissipate very small
amounts of energy). In a pendulum, such as the one in Figure 1.8, gravity is the restoring
force. Dampers generate restoring forces that are functions of the velocity, and they have
the effect of dissipating energy.
• Persistent forces. Such forces arise from imbalances in moving bodies and in bodies
with rotating components. For example, in vehicles, a driveline shaft or tires that have
eccentricity (being off-balance), or a vehicle moving on uneven road, result in persis-
tent loading. Persistent forces are usually periodic, and they may lead to high motion
amplitudes and resonances.
• Impulsive forces. Impulsive forces are forces of large magnitude that are applied over
a very short period of time. Collisions between bodies, such as a bat hitting a baseball,
and impact forces are in this category. Impulsive forces cause sudden changes in velocity
with very little (or negligible) change in position.
• Contact forces, as well as friction. Such forces come into existence when two bodies
have contact with each other. Contact generates normal forces perpendicular to the
plane of contact and friction forces along the plane of contact. Friction forces are resistive
forces, and they dissipate energy when the bodies in contact move against each other.
Friction also makes it possible to have rolling motion.
• Internal forces, such as internal flexibility, internal damping, and hysteresis.
Internal flexibility, as in the flexibility of a beam, has the effect of a spring and cre-
ates restoring forces. Hysteresis is due to the difference in the force-displacement curve
between loading and unloading. Hysteresis and internal damping dissipate energy.
10 Applied Dynamics

1.4.1 Vibratory Motion


A very commonly encountered form of motion, and one that is of particular interest, is
vibration. A system is said to vibrate (or oscillate) when its motion is repetitive and periodic.
Vibrations occur when one or more of the following forces act:
• Internal restoring forces, due to flexibility, damping and hysteresis, as would be
encountered in an elastic beam or rod;
• External restoring forces generated by components such as springs, dampers and
elastic supports connecting a body to other bodies, such as suspension systems in vehi-
cles, also, gravity in the motion of pendulums;
• Persistent external forces, such as wind gusts, vortex shedding, eccentricities and
imbalances in rotating components, such as imbalances in shafts and wheels, as well as
road unevenness for moving vehicles.

Because vibratory motion has periodicity, we wish to develop a mathematical model


of the system and also to analyze the nature of the response. Chapter 5 discusses the
modeling of vibrating systems. Chapters 6 and 7 discuss the response of elastic bodies that
are subjected to the types loads discussed above.

1.5 Systems of Units


Two systems of units are widely used in dynamics, the Système International (SI), or metric,
and U.S. Customary. The primary difference is that the SI system is universal, and absolute,
and the U.S. system is local (or gravitational). The SI system is more widely used. In certain
industries, we can find both systems used side by side. An interesting example of this is tire
labeling.
Three fundamental quantities are needed to describe motion. The SI system uses mass
(M), length (L) and time (T), while the U.S. system uses force (F), length and time. Mass is
defined as the amount of matter contained in a body and it is an absolute quantity. We can
also define mass as the amount of resistance of a body to translational motion (see Chapter
4). Force, on the other hand, is an inertial quantity. Its value depends on the gravitational
system it is in. For example, a body on Mars weighs 3/8 of what it weighs on Earth. The
fundamental quantities used to describe motion can be treated as independent of the others
as long as the speeds involved are not close to the speed of light.
Corresponding to each fundamental quantity is a base unit (also referred to as dimen-
sion) that describes standard amounts of a fundamental quantity. Table 1.1 shows the
fundamental quantities and corresponding base units that are commonly used.
To relate the mass of an object to its weight, we use the gravitational constant g, which
is expressed in units of acceleration. Chapter 4 will discuss the general expression for the
gravitational constant. On Earth, at sea level and at a latitude of 45◦ , g is approximated as
g = 9.8066 m/s2 = 32.174 ft/sec2 . Using these values for g, on Earth an object that weighs
1 lb has a mass of 0.45359 kg.
In the SI system, force is a derived quantity. The base unit of force (F = M L/T 2 ) is
denoted by a Newton (N), where 1 N = 1 kg · m/s2 . In the U.S. system, mass is a derived
quantity and the commonly used unit of mass is a slug, with 1 slug = 1 lb · sec2 /ft.
An object of mass 1 kg has a weight of W = mg = 9.807 kg·m/s2 = 9.807 N. An object
that weighs 1 lb has a mass of m = W/g = 1/32.17 lb·sec2 /ft = 1/32.17 slugs.
Introductory Concepts 11

TABLE 1.1
Base units in the SI and U.S. Customary systems

Quantity SI Units U.S. Customary Conversion


2
Mass kilogram (kg) slug (lb · sec /ft) 1 slug = 14.5921 kg
Length meter (m) foot (ft) 1 foot = 0.3048 m
Time second (s) second (sec)
Force Newton (N = kg m/s2 ) pound (lb) 1 lb = 4.4488 N

To describe rotational displacements, we commonly use degrees (◦ ) or radians (rad).


Going around a full circle takes 360 degrees or 2π radians and is referred to as a revolution.

The angle of 1 radian = 3602π = 57.2958 . The arc length of a segment of a circle of radius
R and angle θ radians is Rθ. The relationship between arc length and radius is depicted in
Figure 1.11.

FIGURE 1.11
Arc length and radius.

Other base units also find widespread use. For example, the speed of ships and airplanes
is customarily described by knots (1 knot = 1 nautical mile/hr, where 1 nautical mile =
1.1508 mile = 1.8520 km). Also, pound-mass (lbm) is sometimes used (especially in the
thermal sciences) to describe mass. An object of mass 1 lbm has a weight of 1 lb. Hence, 1
1
lbm = 32.17 slugs.
It is important to check that the dimensions of the quantities being manipulated match.
Each term in a kinematic expression, force and moment balance equation, or other dynam-
ical relationship must have the same dimension. Checking dimensional homogeneity is a
good way of spotting errors.
We must also be careful in rounding off numbers. Many engineering problems require
an accuracy of one thousandth or better. Accordingly, in most problems in this text, four
significant digits (e.g., 0.1234, 1.234, 12.34) will be considered. Be careful to match the
accuracy of the solution to the accuracy of the different terms used in the solution. If one of
the terms in an expression is only known to a certain level of accuracy, say, 1%, the entire
expression itself cannot be evaluated to a higher level of accuracy, even if the other terms
are known more accurately. As with the weakest link in a chain, a mathematical expression
can only be as accurate as its least accurate component.
12 Applied Dynamics

1.6 Linearization
A useful analysis tool is approximation of a function by means of linearization. Nonlinear
functions are harder to analyze and nonlinear equations (algebraic or differential) are much
harder to solve than linear equations. Dealing with linear expressions makes it possible to use
a variety of well-known and powerful analysis methods, such as the method of superposition.
Linearized models find widespread use in all branches of science and engineering. Two
notable examples from dynamics involve vehicle motions and vibration amplitudes.
A function f (x), where x is the independent variable and f is the dependent variable
(depending on x, in this case), is said to be linear if it can be expressed in the form
f (x) = mx + b, where m and b are constants. Geometrically, a linear function is represented
as a straight line, where m is the slope and b is the point where the function crosses the f
axis.6
Consider the function f (x) in Figure 1.12 and approximate f (x) with a straight line in
the vicinity of a point of interest, say at x = x0 . A good straight line approximation to f (x)
at x = x0 is obtained by drawing a straight line that goes through the point (x0 , f (x0 ))
and is tangent to f (x) at x = x0 .

FIGURE 1.12
Linearizing about points x0 and x1 .


df
Denoting the slope of the tangent by m, where m = dx , and introducing a local
x=x0
variable  so that

x = x0 +  or  = x − x0 (1.2)

we can write the equation for the linear curve in terms of the local variable  as

df
f (x) ≈ f (x0 ) + m or f (x) ≈ f (x0 ) +  (1.3)
dx x=x
0

How good is the linearization approximation? Or, for what range of  will the straight
line approximation be close to the actual value of f (x)? The answer depends on the shape
of the function f (x) at x = x0 . We can observe this by drawing two circles about x0 . Inside
6 The definition of linearity, which states for two arbitrary numbers a and b the relationship f (a) + f (b) =

f (a + b), is only applicable when every term in f (x) is an explicit function of x.


Introductory Concepts 13

the inner circle the tangent line is very close to the function, but there is a deviation when
considering at the outer circle.
Let us now consider another point, say, x1 in Figure 1.12, draw the tangent line at f (x1 ),
and draw the same size circles. Here, the tangent begins to deviate from the function much
faster. This is because the function f (x) is changing more rapidly at x = x1 .
The linearization approximation can also be derived mathematically. Consider the Taylor
series expansion of the function f (x) about x = x0

1 d2 f 1 dn f

df 2
f (x) = f (x0 ) + (x − x0 ) + (x − x 0 ) + . . . + (x − x0 )n + . . . (1.4)
dx x=x 2! dx2 x=x n! dxn x=x
0 0 0

The Taylor series is an infinite series. When an infinite number of terms is used, it is an
exact expansion. The first two terms of the series give the linear approximation

df
f (x) ≈ f (x0 ) + (x − x0 ) (1.5)
dx x=x
0

which is recognized to be the same as Eq. (1.3).


The Taylor series expansion can also be used for functions of more than one variable.
For example, linearization of a function f (x, y) of two variables about x = x0 , y = y0 is

∂f ∂f
f (x, y) ≈ f (x0 , y0 ) + (x − x 0 ) + (y − y0 ) (1.6)
∂x x=x0 ∂y x=x0
y=y0 y=y0

An alternative, but mathematically equivalent, method of linearizing is to substitute the


small displacement terms directly into the nonlinear expression. To linearize f (x) about
x = x0 , introduce the variable  where x = x0 +  and substitute x0 +  into f (x). Then,
eliminate all terms nonlinear in  from the expression f (x0 + ). To this end, we can make
use of well-known relationships, such as small angle formulas and the binomial expansion.
The binomial series expansion is given by

n n (n − 1) n−2 2 n (n − 1) (n − 2) n−3 3
(a + x) = an + nan−1 x + a x + a x + ... (1.7)
2! 3!
where n is any number. For linearization in x the first two terms are retained, giving
n
(a + x) ≈ an + nan−1 x (1.8)

We can also expand (a + x)n as

n
 x n  nx 
(a + x) = an 1 + ≈ an 1 + = an + nan−1 x (1.9)
a a
When linearizing a function that is complex or lengthy, we can make use of an interesting
property of Taylor series. Consider a function f (x) that can be expressed as the product
of two functions, f (x) = f1 (x)f2 (x). The Taylor series expansion of f (x) about a certain
point, say, x0 , is equal to the product of the Taylor series expansions of f1 (x) and f2 (x)
about x = x0 . Hence, we can linearize f1 (x) and f2 (x) separately and multiply the two to
get the linearization of the entire function.
Linearity, in general, is a property not of a system but rather of the range of operation.
Most natural or man-made systems are nonlinear. Usually, we are interested in the behavior
of a system in a certain range, so we linearize the describing equations in that range. Always
check the range in which the linearization approximation is accurate.
14 Applied Dynamics

Example 1.3
Derive the small angle approximations for the sine and cosine functions.
Using the variable θ, first evaluate f (θ) = sin θ about θ = θ0 . Let  = θ − θ0 and note
that θ is measured in radians. Using the first two terms of the Taylor series expansion
sin θ ≈ sin θ0 + cos θ0  [a]
For the special case of linearization about zero, θ0 = 0, introducing the variable  =
θ − θ0 = θ, the above expression reduces to the well-known small angle approximation
sin θ ≈ sin 0 + cos 0  =  = θ [b]
The small angle approximation is valid until θ reaches a value of about 0.4 radians (≈ 25◦ ),
as can be seen by comparing the plots of θ and sin θ in Figure 1.13.
Taylor series can be used in a similar way to linearize the cosine function with the result
cos θ ≈ cos θ0 − sin θ0  [c]
and, for θ0 = 0, the above expression reduces to cos θ ≈ 1.

1
sin(θ)

−1

−2
−1.5 −1 −0.5 0 0.5 1 1.5
θ (rad)

FIGURE 1.13
Plot of sin θ vs θ.

Example 1.4
Linearize the expression f (x, y) = x2 sin (y) about x = 3, y = π/6.
Introduce the variables x and y
π
x = 3 + x y = + y [a]
6
and note that we can linearize the individual terms as
2
x2 = (3 + x ) ≈ 9 + 6x
π √
 π π 1 3
sin y = sin + y ≈ sin + cos y ≈ + y [b]
6 6 6 2 2
Multiplying the above terms and discarding the nonlinear terms gives
√ ! √
1 3 9 9 3
f (x , y ) ≈ (9 + 6x ) + y ≈ + 3x + y [c]
2 2 2 2
Introductory Concepts 15

Example 1.5
1/3
Linearize the term f (x) = (2 + 5x) about x = 0 by making use of the binomial expansion.
Rewrite the function as
  13
1 1 5
f (x) = (2 + 5x) 3 = 2 3 1+ x [a]
2

Introducing the variable y = 5x/2, the above expression can be approximated as


 
1 1 1 1
f (y) = 2 3 (1 + y) 3 ≈ 2 3 1 + y [b]
3

and, converting the variable back, we obtain the linearized expression as


   
1 1 1 5
f (x) ≈ 2 3 1+ y = 2 3 1+ x [c]
3 6

1.7 Differential Equations and the Principle of Superposition


Earlier, we learned how to linearize a nonlinear function about a point of interest. We now
consider linearizing differential equations. Linearization of differential equations permits us
to use several solution methods that are applicable only for linear systems, as well as the
principle of superposition.
Consider a system described by a differential equation in the form

Dx (t) = F (t) (1.10)

in which x (t) is the variable describing the system amplitude, or the output of the system;
F (t) is the excitation, or the input; and D is a differential operator. For example, the
operator associated with the differential equation aẍ (t) + bẋ (t) + c sin x (t) = F is D =
d2 d
a dt2 + b dt + c sin. The input-output and operator D is schematically described in Figure

1.14.

  


  

FIGURE 1.14
Input-output of a system.

Let us denote by x1 (t) the response, or output, of the system to an input F1 (t). Likewise,
let x2 (t) be the response to the input F2 (t). The system is said to be linear if the input
αF1 (t) + βF2 (t), where α and β are arbitrary constants, results in the output αx1 (t) +
βx2 (t).
A differential equation is linear if all the terms involving the variable and its derivatives
are to the first power, with no second or higher power terms, no square or cube or higher
16 Applied Dynamics

roots, no discontinuities, no transcendental terms, or no absolute value functions. A linear


n-th order ordinary differential equation can be written as
dn x (t) dn−1 x (t)
an n
+ an−1 + . . . + a0 x (t) = F (t) (1.11)
dt dtn−1
The coefficients a0 , a1 , . . . , an may be constant or they may be functions of time.
When linearizing an ordinary differential equation of order two, such as the describing
equation of a dynamical system, we linearize about an operating point. Denoting the variable
by x(t), we linearize about a point x (t) = x0 and ẋ (t) = v0 . Introducing the variable , we
express the position and velocity as
x (t) = x0 +  (t) ẋ (t) = v0 + ˙ (t) (1.12)
The expression for acceleration can be written as ẍ (t) = a0 + ¨ (t), where the value of a0
depends on the values of x0 and v0 .
Dealing with linear differential equations has two distinct advantages:
• Linearity permits us to use several methods for solving differential equations, such as
homogeneous and particular solution, Laplace transform solution, or series expansions.
• Given a linear system subjected to more than one type of input (excitation), we can
obtain the solution to each excitation separately and then combine the individual so-
lutions linearly to get the total response. This property is known as the principle of
superposition. Superposition enables us to break a lengthy problem into smaller parts
where each part can be solved more easily. Chapters 5, 6, and 7 provide examples of lin-
earization, selection of the operating point about which we linearize, and solving single
or multi-degrees-of-freedom linear differential equations by superposition.

Example 1.6
Is the differential equation 3ẍ (t) − 5 |ẋ (t)| + cos (x (t)) − t2 x (t) = 0 linear?
No, it is not. The term |ẋ (t)| is not linear, as it involves the absolute value function.
The absolute value function is not linear as it has a discontinuous derivative at ẋ (t) = 0.
The term cos(x (t)) is also nonlinear. Note that the term t2 x (t) is linear in the variable
x (t), so the differential equation 3ẍ (t) − 5ẋ (t) − t2 x (t) = 0 is a linear ordinary differential
equation.

Example 1.7
Linearize the differential equation ẍ + 3ẋ3 + 4 sin x = 0 about the operating point x0 = π,
ẋ0 = v0 = 0.
Let us introduce the local variable , where x (t) = x0 +  = π + . It follows that
ẋ = v0 + ˙ = ˙ ẍ = a0 + ¨ [a]
The value of a0 is obtained by introducing x0 and v0 into the differential equation, with
the result
a0 + 3v03 + 4 sin x0 = 0 =⇒ a0 = −3v03 − 4 sin x0 = 0 [b]
Introduction of these definitions into the differential equation results in
¨ + 3˙3 + 4 sin (π + ) = ¨ + 3˙3 + 4 sin π cos  + 4 cos π sin  [c]
Note that sin π = 0, cos π = −1, and the small angle approximation is sin  ≈ . Further-
more, ˙3 ≈ 0. Substituting these relationships to Eq. [c] yields the linearized equation
¨ − 4 = 0 [d]
Introductory Concepts 17

1.8 Dimensional Analysis and Nondimensionalization


Two valuable mathematical tools are dimensional analysis and the process of nondimen-
sionalization. Dimensional analysis is useful in experimental or numerical work, as well as
in design. In dimensional analysis, the goal is to reduce the complexity of the variables and
parameters that describe a system. The reduction is carried out by a compacting technique,
such as scaling. In the process, we develop or identify dimensionless ratios that often point to
important relationships. For example, the damping factor, derived in Chapter 6, quantifies
the amount of energy dissipation.
In dimensional analysis there are two basic approaches. One is directly to attempt to find
dimensionless parameters that influence the behavior of a system. This approach is heavily
used in fluid mechanics. The Reynolds number and the Prandtl number, for example, are
two very important dimensionless quantities. Dimensional analysis also enables us to deal
with scale models, which is very useful in experimentation.
In the second approach, we nondimensionalize the describing equations and in the pro-
cess identify important dimensionless quantities that characterize the system behavior. This
approach is especially useful for motion description.
When nondimensionalizing a describing equation, such as an equation of motion, we
select a set of reference constants that are characteristics of the particular problem. Several
choices usually exist for this selection. Be careful and consistent in order to avoid ambiguities
and redundancies. Initial values of variables and natural frequencies usually are good choices.
When nondimensionalizing an expression, it is customary to rewrite the expression in
terms of starred terms first. For a derivative term, such as ẍ, we need to nondimensionalize
it with respect to the displacement x and time t. The derivative is obtained by replacing t
with t∗ and x with x∗
d2 x d2 x∗
ẍ = =⇒ (1.13)
dt2 dt∗2
The nondimensional terms are written as
x∗
x = t = t∗ ω (1.14)
x0
where x and t are now nondimensional and x0 and ω are the reference constants, having
units of displacement and 1/time, respectively. Differentiation with respect to t∗ has the
form
d d d d2 2 d
2


= = ω = ω (1.15)
dt d (t/ω) dt dt∗2 dt2

so that the second derivative term becomes


d2 x∗ d2 x
∗2
= x0 ω 2 2 (1.16)
dt dt
Example 1.8
Gantry cranes are used in loading and unloading container ships and other heavy loads.
They lift objects by a hoist which is fitted in a trolley. The trolley can move horizontally on
a rail or on a pair of rails. A simple model of a gantry crane is shown in Figure 1.15, where
the trolley is modeled as a cart and the crane by a pendulum. The interest is in the case
when the trolley is moving, during which time the length of the pendulum L is fixed. The
18 Applied Dynamics

Trolley

x
F

L

Crane

FIGURE 1.15
Simple model of a gantry crane.

assumption is that the motion of the trolley, denoted by x, is known or specified, leading to
a one-degree-of-freedom model. The linearized equation of motion, in terms of the motion
variable θ, where θ is measured in radians, can be shown to be
g 1
θ̈ + θ = − ẍ [a]
L L
The next step is to nondimensionalize the equation of motion. Two reference constants
are needed, one for x and the other for time. Note that θ is dimensionless. Using starred
variables, the equation of motion becomes

d2 θ g 1 d2 x∗
∗2
+ θ = − [b]
dt L L dt∗2
p
Two meaningful values for the reference constants are L and ω, where ω = g/L, so
that the dimensionless quantities are x = x∗ /L, t = t∗ ω. Introduction of these values into
Eq. [a] yields
g ω2
ω 2 θ̈ + θ = − Lẍ = −ω 2 ẍ [c]
L L
Since ω 2 = g/L, dividing each term above by ω 2 yields the nondimensional equation as

θ̈ + θ = −ẍ [d]

which is in much simpler form than Eq. [a]. We can analyze the motion of the crane and
carry out any necessary motion planning (it is important for the crane operator to move
the trolley so that there is as little sway as possible in the crane) by using the dimensionless
equation [d] and not in terms of a specific pendulum length.

1.9 Numerical Integration


An important tool for obtaining the response of a system is numerical integration. Devel-
opments in computer hardware and software have made it very easy and desirable to use
Introductory Concepts 19

numerical techniques for analysis as well as simulation. This section discusses numerical in-
tegration of ordinary differential equations which are commonly used to describe the motion
of dynamical systems.
When a system is described by linear differential equations, we can make use of powerful
solution techniques to obtain the response. These techniques are discussed in Chapters 6
and 7. When the description of the system is in terms of nonlinear equations, one can either
linearize the describing equations, conduct an analytical study, or use numerical integration
to obtain the response. Numerical analysis also is very useful for obtaining the response of
linear systems, especially those that have several degrees of freedom.
Numerical integration software requires that the differential equations be written in state
form. In state form, the differential equations are of order one, there is a single derivative
on the left side of the equations, and there are no derivatives on the right side. A system
described by a higher-order ordinary differential equation has to be converted into one that
is in state form.
Consider, for example, the second-order differential equation

ẍ + 3ẋ3 + sin (4x) − 0.3x = 0 (1.17)

We convert this equation into two first-order differential equations by defining two vari-
ables, y1 (t) and y2 (t) as

y1 = x y2 = ẋ (1.18)

The derivatives of these variables cast the differential equation into state form

ẏ1 = ẋ = y2

ẏ2 = ẍ = −3ẋ3 − sin (4x) + 0.3x = −3y23 − sin (4y1 ) + 0.3y1 (1.19)

Example 1.9—MATLABr Program ode45


The MATLABr ordinary differential integration routine ode45 is a reliable program to
integrate a set of ordinary differential equations. We write two .m files, one describing the
time span (range of integration) and initial conditions, while the second .m file describes
the ordinary differential equations to be integrated.
Consider the differential equation in the previous section and select the initial conditions
as x (0) = y1 (0) = 1, ẋ (0) = y2 (0) = −0.7. There are two first-order differential equations
to integrate. The first .m file contains the following commands:
tspan = [0 10] %(defines the range of integration, selected here as from zero to 10 seconds)
yinit = [1; -0.7] %(defines the two initial conditions)
[t, y] = ode45(‘desc’, tspan, yinit) %(calls the .m file named desc.m)
Plotting commands or any other commands that you want for further analysis
The second .m file has the name desc.m and it describes the differential equations. For
the example here, the contents of this file are
function ydot = desc(t,y)
yd(1) = y(2)
yd(2) = -3*y(2)*y(2)*y(2) - sin(4*y(1)) + 0.3*y(1)
ydot = [yd(1); yd(2)]
The name desc appears in three places: the name of the second .m file, in the first line
of that file, and in the first .m file.
20 Applied Dynamics

1.10 What Is a Vehicle?


Loosely defined, a vehicle is a contraption designed and built for carrying a payload from
one location to another in an effective way. Different needs and different technologies have
led to the development of a tremendous variety of vehicles. A vehicle can be thought of
comprising of four major parts:

• A cabin for carrying the payload and passengers, to shelter them from the elements and
from hazards, as well as to provide life support;
• A power plant for generating the energy to move the vehicle;
• A propulsion mechanism to convert the energy generated by the power plant into propul-
sive forces and moments, as well as a braking system to slow down or stop the vehicle;
• A steering mechanism to steer or guide the vehicle in the desired direction.

For example, in a car, the engine is the power plant; the propulsion mechanism consists
of the transmission, drive shaft, differential, axles, suspension system, tires, and brakes; and
the steering mechanism consists of the steering wheel, steering column, kingpin, and steering
rack (or equivalent mechanism) to turn the wheels. The cabin is the passenger section and
the trunk. Not only are these systems interconnected, but there also are auxiliary systems,
such as electrical, that assist every function of the vehicle.
The distinction between the different parts of a vehicle sometimes get blurred. For
example, in an airplane, the turbines (or propellers) combine the power plant and propulsion
mechanism. The steering mechanism consists of the aerodynamic surfaces and the rudder.
Note that, unlike in a boat, the rudder on an airplane is not the primary steering mechanism;
rather, it controls yaw.
For most vehicles, it is preferable to have one propulsion mechanism and a steering
mechanism, rather than several propulsion mechanisms acting in different directions. In
a spacecraft, however, propulsion and steering are combined by using multiple thrusters.
Newer boat designs combine propulsion and steering by means of two or more propellers.
Design of the components of a vehicle often creates conflicts. For example, the cabin
designer will want to maximize the strength of the cabin, while the power plant designer
would like to minimize weight for reduced power consumption. Competing interests have to
be balanced.

1.11 Cause-and-Effect Principle


A fundamental principle of mechanics is the cause-and-effect principle, which says the fol-
lowing:

1. If there is an effect (such as a displacement, deformation, acceleration, rotation, con-


strained motion, or sliding), there must be a cause (such as a force or torque) behind
it. If there is a cause, then there is an effect that results from it.
2. In a well-defined problem, if we know the magnitude of the cause, then we do not know
the magnitude of the effect, and if we know the magnitude of the effect, we do not know
the magnitude of the cause, until the problem is solved.
Introductory Concepts 21

While this principle seems obvious and logical, many times it is forgotten or overlooked.
There are several examples of the cause-and-effect principle throughout the text. Table 1.2
shows common cause-and-effect relationships encountered in mechanics.

TABLE 1.2
Cause-and-effect relationships

Type (System) Cause (Input) Effect (Output)


Rigid body translation Force Translational acceleration
Rigid body rotation Torque (Moment) Angular acceleration
Any type of body Constraint force Constraint relation
Sliding over a body Friction force Sliding or no sliding condition
Rolling motion Friction force at contact Slip or no slip condition
Axial (extensional) spring Force Spring deflection
Torsional spring Torque Spring deflection
Deformable body Force (stress) Axial deformation (strain)
Deformable body Moment or torque (stress) Rotational deformation (strain)
Static problems Support forces & moments Zero deformation

The units of cause multiplied by the units of effect have units of work and energy (energy
per volume for stress and strain). The work done can be written as the integral
Z
Work = Cause × d (Effect)

There are certain types of causes that do not do any work, and they will be discussed in
Chapter 5.

1.12 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Benaroya, H., Mechanical Vibration: Analysis, Uncertainties, and Control, 3rd Edition,
CRC Press, 2010.
Gillespie, T.D., Fundamentals of Vehicle Dynamics, SAE Publications (R114), 1992.

1.13 Problems
Problems are marked by E—easy, M—moderate, and D—difficult.

Section 1.3—Degrees of Freedom


1.1 (E) Calculate the number of degrees of freedom for the rod in Figs. 1.16a and 1.16b.
The rods maintain contact with the surfaces they slide on.
1.2 (E) Calculate the number of degrees of freedom for the cart-rod system in Figure 1.17.
22 Applied Dynamics

!" #"
A
A L2
L1 B
 L
O 

B $%&'(

FIGURE 1.16
Figures for Problem 1.1.

  

   

 

FIGURE 1.17
Figure for Problem 1.2.

1.3 (M) The rod in Figure 1.18 is suspended by two springs. Assuming the springs move only
in the vertical direction, the rod has two degrees of freedom. Suggest two sets of generalized
coordinates that can be used to describe the motion of the rod.







FIGURE 1.18
Figure for Problem 1.3.

Section 1.5—Systems of Units


1.4 (E) A vehicle has a mass moment of inertia of I = 2500 lb·ft·sec2 . Calculate I in terms
of the SI system (kg·m2 ).
1.5 (E) A damper has a constant of c = 12 N·s/m. Calculate the damping constant in U.S.
units.
Introductory Concepts 23

1.6 (E) An axial (extensional) spring has a constant of k = 50 lb/in. Calculate the spring
constant in the SI system, using units of Newtons and millimeters.
1.7 (M) An average size apple has a mass of 100 grams. On Mars, where gMars = 0.375gEarth ,
how much would the apple weigh in Earth pounds?

Section 1.6—Linearization
1.8 (E) Linearize the expression f (θ) = sin θ about θ = π/3.
1.9 (E) Linearize the expression f (x, y) = (1 + y)3 cos (2x) about x = π/6, y = 1.2.
1/2
1.10 (M) Use the binomial expansion to linearize the function f (x) = (2x + 3.5) about
x = 1.4. 2
1.11 (M) Linearize the expression f (x) = 1 − x2 sin (2x) about x = 0.25 by splitting
f (x) into two parts and by linearizing each part separately.

Section 1.7—Differential Equations and the Principle of Superposition


2
1.12 (M) Linearize the differential equation ẍ+3ẋ cos (ẋ)+(x − 2) = 0 about x = 2, ẋ = 0.
1.13 (M) Linearize the differential equation ẍ + 3ẋ2 − 6x3 + 7 sin x = 0 about x = 0, ẋ = 0.
1.14 (M) Linearize the differential equation ẍ + 7ẋ2 + 5x3 = 0 about x = 2, ẋ = −3.

Section 1.8—Dimensional Analysis and Nondimensionalization


1.15 (M) The describing equation of a first-order system is mẋ (t) + cx (t) = 0, with initial
condition x (0) = A. The unit of x is position. Nondimensionalize this equation and initial
condition.
1.16 (M) The describing equation of a second-order system is mẍ (t) + kx (t) = 0, with
initial conditions x (0) = A, ẋ (0) = B. The unit of x is position and the unit of m is mass.
Nondimensionalize this equation and initial conditions.
1.17 (M) The describing equation of a second-order system is mẍ (t) + k2 x3 (t) = 0, with
initial conditions x (0) = A, ẋ (0) = 0. The unit of x is position and the unit of m is mass.
Nondimensionalize this equation and initial conditions.

Section 1.9—Numerical Integration


1.18 (E) The describing equation of a second-order system is mẍ + k2 x3 = F . Write this
equation in state form.
...
1.19 (M) The describing equation of a third-order system is 2 x + 5ẍ − 6ẋ2 + 2x3 = G. Write
this equation in state form.
1.20 (M) Consider Problem 1.12 and cast the differential equation into state form.
1.21 (M) Consider Example 7.4 and cast the equations of motion that are obtained into
state form.
1.22 (D) Consider Example 7.7 and cast the equations of motion into state form given the
values M = 4, m = 3, L = 1, k = 8, g = 10, c = 0.
2
Kinematics Fundamentals

2.1 Introduction
This chapter discusses the kinematics of motion, that is, looking at the nature of the motion
without examining the forces that cause the motion. We will focus on the two- and three-
dimensional kinematics of particles, as well as planar kinematics of rigid bodies. Three-
dimensional kinematics of rigid bodies will be discussed in Chapter 9.
We begin with coordinate systems and the kinematics of particles. The motion of par-
ticles is purely translational. Then, rotating reference frames, rotation parameters, angular
velocity, and angular acceleration are discussed. Relative velocity and acceleration equations
are developed. Instant centers are introduced, as they are crucial to the analysis of vehicles
and mechanisms.
Kinematic analysis serves two purposes: First and foremost, it is a precursor to kinetic
analysis, a topic that will be discussed in the next chapter. We cannot analyze the kinetics
of a system without first studying its kinematics. In addition, kinematic analysis by itself
is a valuable tool and is widely used in the design of mechanisms and vehicle suspensions,
as well as in motion planning. Chapter 3 discusses applications of kinematics.

2.2 Position, Velocity, and Acceleration


When studying the kinematics of a particle, that is, the translational motion of a point, we
need to describe its position, velocity, and acceleration. The description must be made with
respect to a reference point or origin using a coordinate system. A coordinate system or
coordinate frame is characterized by a set of coordinate axes, the positive directions of these
axes, and unit vectors along these axes. Several types of coordinate systems exist. We select
the coordinate system (or frame) that will make the analysis easier and more meaningful.
Selecting a coordinate frame is, in essence, selecting the set of motion variables.
Figure 2.1 depicts a particle that is moving. The position is measured from a reference
point that is fixed, and the position vector is denoted by r (t). Consider next the position
of the same particle at a time increment ∆t later, at time t + ∆t. The position vector at
t + ∆t is r (t + ∆t). The velocity v (t) and acceleration a (t) of the particle are defined as

dr ∆r r (t + ∆t) − r (t)
v (t) = = lim = lim
dt ∆t→0 ∆t ∆t→0 ∆t

dv v (t + ∆t) − v (t)
a (t) = = lim (2.1)
dt ∆t→0 ∆t
The rate of change of acceleration is of interest in several applications. Examples include

25
26 Applied Dynamics

FIGURE 2.1
A particle and its path.

vehicle dynamics and human motion analysis. The commonly used terms for the rate of
change of acceleration are shock are jerk. The occupants of a vehicle get shaken and thus
experience discomfort if the acceleration profile undergoes a sudden change, such as when
accelerating or braking rapidly and when taking sharp turns. A vehicle and its components
wear out sooner when they are repeatedly subjected to sudden accelerations.

2.3 Reference Frames: Single Rotation in a Plane


Coordinate systems are used in kinematics to observe motion. We decide on which coordi-
nate system to use by considering the nature of the motion. This section develops relation-
ships between different coordinate systems and transformations from one coordinate system
to another. Only right-handed coordinate systems are considered here. When we point our
right hand towards the positive direction of one of the axes (say, x, with unit vector i) and
rotate our fingers towards the positive direction of the second axis (say, y, with unit vector
j), the thumb points in the positive direction of the third axis z, with unit vector k = i × j.

Y
y
sin
J


cos x
j
i sin

X
cos I

FIGURE 2.2
The XY and xy coordinate systems.

Consider plane motion and a planar coordinate system XY , as shown in Figure 2.2. The
Kinematics Fundamentals 27

unit vectors along the X and Y directions are I and J, respectively. Also shown in the same
figure is a coordinate system xy (with unit vectors i and j) that is obtained by rotating
the XY axes by an angle θ in the counterclockwise direction. The rotation is about the
Z axis (not shown here) perpendicular to the plane. The Z and z axes are the same. The
relationship between the unit vectors of the two coordinate systems is
i = cos θI + sin θJ j = − sin θI + cos θJ

I = cos θi − sin θj J = sin θi + cos θj (2.2)


The unit vector perpendicular to the plane of motion is common to both coordinate systems,
so that k = K. The above relationships can be expressed in matrix form as
         
i cos θ sin θ I I cos θ − sin θ i
= = (2.3)
j − sin θ cos θ J J sin θ cos θ j
Define the rotation matrix [R] as
 
cos θ sin θ
[R] = (2.4)
− sin θ cos θ
The matrix [R] is unitary; its determinant is equal to 1 and its inverse is equal to its trans-
pose, [R]−1 = [R]T . The two relationships in Equation (2.3) are inverse transformations.
Next, consider a point P on the plane (Figure 2.3) and express the coordinates of point
P as (XP , YP ) in the XY coordinates and (xP , yP ) in the xy frame, as shown in Figure 2.3.
A set of independent coordinates that describe the orientation of a system completely are
called generalized coordinates. The vector rP , which denotes the position of point P , can
then be written in terms of the two coordinate systems as
rP = XP I + YP J = xP i + yP j (2.5)

Y
y
YP P

yP

rP

J x
j
i xP

X
I XP

FIGURE 2.3
Representation of vector rP in XY and xy coordinates.

It is of interest to explore the relationship between the components of the two descrip-
tions of rP . To this end, introduction of Equation (2.3) to the above equation results in
rP = XP I + YP J = XP (cos θi − sin θj) + YP (sin θi + cos θj)
28 Applied Dynamics

= (XP cos θ + YP sin θ) i + (YP cos θ − XP sin θ) j = xP i + yP j (2.6)


from which we conclude that
xP = XP cos θ + YP sin θ yP = YP cos θ − XP sin θ (2.7)
At this stage, we introduce the column vector representations of the unit vectors and of
the location of point P in the two reference frames:
   
xP XP
{xyz rP } =
XY Z
rP = (2.8)
yP YP
and express the transformation between the two column vectors as
   
xP XP
{xyz rP } = [R] XY Z rP

= [R] or (2.9)
yP YP
The inverse relationship between the two vectors is
rP = [R]−1 {xyz rP }
XY Z XY Z
or rP = [R]T {xyz rP } (2.10)
The above relationships can be generalized to three dimensions by redefining the column
vector representations and the rotation matrix as
     
xP XP cos θ sin θ 0
{xyz rP } =  yP  XY Z

rP =  YP  [R] =  − sin θ cos θ 0  (2.11)
zP ZP 0 0 1
We see that the same relationship that governs the position vectors is also valid for
the unit vectors. Indeed, expressing the unit vectors in column vector format [i j k]T and
[I J K]T , we can write
       
i I I i
 j  = [R]  J   J  = [R]T  j  (2.12)
k K K k

Example 2.1
The XY Z coordinate system is rotated about the Z axis by 30◦ clockwise to obtain the
xyz coordinates. Consider the vector r = 3i − 4j and express it in terms of the XY Z frame.
Using column vector notation, r is
 
xyz 3
{ r} = [a]
−4
The transformation angle is θ = −30◦ , so that the matrix between the two coordinate
systems is    
cos θ sin θ 0.8660 −0.5000
[R] = = [b]
− sin θ cos θ 0.5000 0.8660
The second part of Equation (2.10) can be used to express r in the XY Z frame, with
the result
    
XY Z 0.8660 −0.5000 3 0.5981
r = [R]T {xyz r} = = [c]
0.5000 0.8660 −4 −4.9641

The results can be checked by noting that the magnitude of {xyz r} and XY Z r must
be the same. Therefore,
XY Z T XY Z
{xyz r}T {xyz r} = 32 + 42 = 25 r r = 0.59812 + 4.96412 = 25.00001
[d]
where the difference is due to roundoff error.
Kinematics Fundamentals 29

2.4 Column Vector Representation


This section discusses two notations to represent vectors. Consider a coordinate system with
unit vectors e1 , e2 , and e3 , which form a mutually orthogonal set. Also consider two vectors
r and q defined as

r = r1 e1 + r2 e2 + r3 e3 q = q1 e1 + q2 e2 + q3 e3 (2.13)

Vectors described this way are referred to as geometric vectors or spatial vectors. The dot
and cross products of these vectors yield

r · q = r1 q1 + r2 q2 + r3 q3

r × q = (r2 q3 − r3 q2 ) e1 + (r3 q1 − r1 q3 ) e2 + (r1 q2 − r2 q1 ) e3 (2.14)

The previous section demonstrated that the vectors r and q can be expressed in column
vector format as
   
r1 q1
{r} =  r2  {q} =  q2  (2.15)
r3 q3

The column vectors are also referred to as algebraic vectors. Using this description, we
can express the dot product of two geometric vectors in column vector format as

r · q =⇒ {r}T {q} (2.16)

The skew-symmetric matrix [r̃] associated with the column vector {r} is a compact way of
expressing a cross product. Define it as
 
0 −r3 r2
[r̃] =  r3 0 −r1  (2.17)
−r2 r1 0

so that
 
r2 q3 − r3 q2
r × q =⇒ [r̃] {q} =  r3 q1 − r1 q3  (2.18)
r1 q2 − r2 q1

Note that because r × q = −q × r, the relationship [r̃] {q} = − [q̃] {r} also holds.
In kinematics we frequently encounter the vector product r × (r × q). The expression is
commonly shortened to r × r × q, with the understanding that the cross product between
r and q is performed first. Using the notation introduced above,

r × (r × q) =⇒ [r̃] [r̃] {q} (2.19)

The matrix multiplications in [r̃] [r̃] {q} can be performed in any order.
Another use of the column vector notation arises when taking derivatives of a function
with respect to a set of variables, or when taking the derivative of a scalar with respect
to a vector. Consider a vector {q} = [q1 q2 . . . qn ]T of dimension n, where the elements
q1 , q2 , . . . , qn are variables that are independent of each other, and a scalar S which is a
function of these variables, S = S (q1 , q2 , . . . , qn ). The derivative of S with respect to the
30 Applied Dynamics

vector {q} is defined as the n-dimensional row vector dS/d{q}, whose elements have the
form
 
∂S ∂S ∂S ∂S
= ... (2.20)
∂{q} ∂q1 ∂q2 ∂qn
The derivative of one column vector with respect to another can be obtained in a similar
fashion. Consider the column vector {v} of order m, where {v} = [v1 v2 . . . vm ]T , where
the elements of {v} are functions of q1 , q2 , . . . , qn . The derivative of {v} with respect to
{q} is a matrix of order m × n having the form
 ∂v1 ∂v1 ∂v1 
∂q1 ∂q2 . . . ∂q n
 ∂v2 ∂v2 ∂v2 
d{v}  ∂q ∂q2 . . . ∂qn 

=  1
(2.21)
d{q}  ...
 

∂vm ∂vm
∂q1 ∂q2 . . . ∂v
∂qn
m

For the special case when the scalar S is in quadratic form and expressed as S =
{q}T [D] {q}, where the elements of the matrix [D] are not functions of the variables
q1 , q2 , . . . , qn , the derivative of S with respect to {q} has the form
dS d{q}T [D] {q} T
= = {q}T [D] + {q}T [D] (2.22)
d{q} d{q}
and when the matrix [D] is symmetric we obtain
dS d{q}T [D] {q}
= = 2{q}T [D] (2.23)
d{q} d{q}

Example 2.2
Consider the scalar S = 3x2 + 4y 2 − 5xy and express it in terms of a symmetric matrix [D].
Writing the variables in vector form as {q} = [x y]T and taking the derivative of S with
respect to {q} gives
dS
{v}T = = [6x − 5y 8y − 5x] = 2{q}T [D] [a]
d{q}
from which it follows that
T
{v} = 2 [D] {q} = 2 [D] {q} [b]
and  
d{v} 6 −5
= 2 [D] = [c]
d{q} −5 8
Thus, the matrix [D] is  
3 −2.5
[D] = [d]
−2.5 4

2.5 Commonly Used Coordinate Systems


This section discusses four coordinate systems that are commonly used to describe motion.
One of these, rectilinear coordinates, involves unit vectors that are fixed in space, and the
other three are moving coordinate systems.
Kinematics Fundamentals 31

2.5.1 Rectilinear Coordinates

Z
P

$P
#
ZP
! Y
"
! XP
X

YP

FIGURE 2.4
Rectilinear coordinates.

The axes of a rectilinear coordinate system are fixed in direction. The unit vectors
along the coordinate axes are also fixed, and hence, their derivatives are zero. Consider
a coordinate system XY Z with unit vectors I, J, and K along the X, Y , and Z axes,
respectively. Another commonly used coordinate set is xyz with unit vectors i, j, and k.
Also, consider a reference point O and a point P , as shown in Figure 2.4. The position
vector rP , which describes the position of point P , has the form
rP = XP I + YP J + ZP K (2.24)
in which XP , YP , and ZP are the coordinates of point P , that is, their distance from the
origin O along the X, Y , and Z axes. To obtain the velocity of point P , denoted by vP ,
the above expression is differentiated with respect to time. Noting that the time derivative
of the unit vectors is zero,
drP
vP = = ẊP I + ẎP J + ŻP K + XP İ + YP J̇ + ZP K̇ = ẊP I + ẎP J + ŻP K (2.25)
dt
with the overdots denoting differentiation with respect to time. Similarly, the acceleration
of point P , which is denoted by aP , has the form
dvP
aP = = ẌP I + ŸP J + Z̈P K (2.26)
dt
The advantage of using a rectilinear coordinate system is its simplicity and the ease
with which components of the motion in other directions are viewed. The simplicity of
rectilinear coordinates, however, is also their disadvantage, as rectilinear coordinates provide
no information about the nature of the path that is followed.
Rectilinear coordinates are useful when components of the motion can be separated from
each other. A common application is projectile motion.

Example 2.3
A basketball player wants to shoot the basketball into the hoop. The player is at a distance
L from the basket and the basket is at a height h from the player’s chest, from where
32 Applied Dynamics

the player launches the ball. The player wants the ball to travel as a projectile and reach
a height of 5h/4 before it begins its descent towards the basket, as shown in Figure 2.5.
Calculate the initial velocity v0 and angle θ with which the player needs to launch the ball.

z A
B

!
h
vo "
h


x
L

FIGURE 2.5
Basketball on a trajectory.

Neglecting the aerodynamics, the accelerations of the projectile in the x and z directions
are
ax = 0 az = −g [a]
and the initial velocities in the x and z directions are
vx0 = v0 cos θ vz0 = v0 sin θ [b]
It follows that the expressions for velocity and acceleration can be written separately in
the x and z directions as
vx = v0 cos θ x = v0 cos θt
1 2
vz = v0 sin θ − gt z = v0 sin θt −
gt [c]
2
Point A is the point where the peak amplitude is reached. This happens at time tA . At
this point, the vertical velocity is zero, or vz = v0 sin θ − gtA = 0. Solving for the time tA
we obtain
v0 sin θ
tA = [d]
g
and the height reached at this point is
2
1 v02 sin2 θ

v0 sin θ 1 v0 sin θ 5
z (tA ) = v0 sin θ − g = = h [e]
g 2 g 2 g 4
Next, consider the time it takes for the ball to reach the basket, that is, point B. The
horizontal distance traversed, L, can be expressed as x = L = v0 cos θtB . Solving for the
time tB to reach B we obtain
L
tB = [f ]
v0 cos θ
Introduction of the above expression into the height at time t = tB gives the height at
time tB as  2
L 1 L
z (tB ) = h = v0 sin θ − g [g]
v0 cos θ 2 v0 cos θ
Kinematics Fundamentals 33

There are two unknowns, v0 and θ. The two equations that need to be solved are Equa-
tions [e] and [g]. The solution can be simplified by introducing the variables u = v0 cos θ
and w = v0 sin θ. Equation [e] can be rewritten as

1 w2 5 5
= h =⇒ w2 = gh [h]
2 g 4 2

Introducing this result into Equation [g] and rearranging leads to a quadratic equation in
terms of u r
5g 1 gL2
u2 − Lu + = 0 [i]
2h 2 h
which can be solved as
r r r ! r r r !
1 gL2 5 5 1 gL2 5 1
u = ± −2 = ± [j]
2 h 2 2 2 h 2 2

Note that there are two solutions. After calculating u and w, the next step is to solve
for the launch angle θ using the relation
w v0 sin θ
= = tan θ [k]
u v0 cos θ

from which the angle θ is obtained as θ = tan−1 (w/u). We can then introduce the value of
θ to any one of the expressions for u or w to find the launch speed.
This example can be used as a parametric study to determine the best options for
maximizing the possibilities of scoring a basket.

2.5.2 Normal-Tangential Coordinates


Normal-tangential coordinates take into consideration the properties of the path taken by
the moving body, which is extremely useful. On the other hand, normal-tangential coordi-
nates are not very useful in describing position.

 

 
 
 





FIGURE 2.6
Particle on a curved path.

Consider a particle moving along a curved path. The normal-tangential coordinate sys-
tem is a moving coordinate system attached to the particle. Two principal directions describe
the motion, normal and tangential. To obtain these directions, consider the position of the
particle after it has traveled distances s and s + ∆s along the path, as shown in Figure 2.6.
34 Applied Dynamics

The associated position vectors, measured from a fixed location, are denoted by r (s) and
r (s + ∆s), respectively. Define by ∆r the difference between r (s) and r (s + ∆s); thus,

∆r = r (s + ∆s) − r (s) (2.27)

As ∆s becomes small, ∆r and ∆s have the same length and become parallel to each
other. Further, ∆r becomes aligned with the tangent to the curve. The tangential direction
is defined as the direction tangent to the curve with the positive direction in the same
direction as the velocity. The unit vector in the tangential direction is defined as
∆r dr
et = lim = (2.28)
∆s→0 ∆s ds
The tangential direction is shown in Figure 2.7. The unit vector et changes direction as the
particle moves.

 

 










FIGURE 2.7
Normal and tangential directions.

The velocity is obtained by differentiating the displacement vector with respect to time.
Using the chain rule for differentiation gives
dr dr ds
v = = (2.29)
dt ds dt
Using the definition of et from Equation (2.28) and noting that the speed v is the rate
of change of the distance traveled along the path, v = ds/dt, the expression for velocity
becomes

v = vet (2.30)

The second principal direction is defined as normal to the curve and directed toward
the center of curvature of the path, and it is shown in Figure 2.7. This direction is defined
as the normal direction (n), and the associated unit vector is denoted by en . The center
of curvature associated with a certain point on a path lies along a line perpendicular to
the path. An infinitesimal arc of the curve in the vicinity of that point can be viewed as
Kinematics Fundamentals 35

a circular path, with the center of curvature at the center of the circle. The radius of the
circle is called the radius of curvature and is denoted by ρ. The two unit vectors introduced
above are orthogonal, that is, et · en = 0.
Differentiation of Equation (2.30) with respect to time gives the acceleration of the
particle a as

a = v̇ = v̇et + v ėt (2.31)

et(s)

et(s+ds) det en t

FIGURE 2.8
Infinitesimal change in tangential direction.

The derivative of et is needed to calculate the acceleration. To this end, consider Figure
2.7 and displace the particle by an infinitesimal distance ds along the path. The unit vectors
associated with the new location are et (s + ds) and en (s + ds). The center of curvature
remains the same as the particle is moved infinitesimally, so the arc length can be expressed
as ds = ρdφ in which dφ is the infinitesimal angle traversed as the particle moves by
a distance ds. Define the vector connecting et (s + ds) and et (s) by det , so that det =
et (s + ds) − et (s). From Figure 2.8, the angle between et (s + ds) and et (s) is small, so
that
ds
|det | ≈ sin dφ |en (s)| ≈ dφ = (2.32)
ρ
or

det 1
ds = ρ (2.33)

The radius of curvature is a measure of how much a curve bends. For motion along a
straight line, the curve does not bend and the radius of curvature has the value of infinity.
For plane motion, using the coordinates x and y such that the curve is described by y = y(x),
the expression for the radius of curvature can be shown to be
2
1 d y/dx2
= h i3/2 (2.34)
ρ 2
1 + (dy/dx)

The absolute value sign in the above equation is necessary because the radius of curvature is
defined as a positive quantity. Considering the sign convention adopted above, the derivative
of the unit vector in the tangential direction becomes
det en
= (2.35)
ds ρ
36 Applied Dynamics

Using the chain rule, the time derivative of et becomes


det ds v
ėt = = en (2.36)
ds dt ρ

Introduction of this relationship to Equation (2.31) yields the acceleration as

v2
a = v̇ = v̇et + en (2.37)
ρ
The first term on the right in this equation is the component of the acceleration due
to a change in speed, referred to as tangential acceleration (at ). The second term is the
contribution due to a change in direction, referred to as the normal acceleration (an ). The
acceleration expression can be written as

a = at et + an en (2.38)

with at = v̇ and an = v 2 /ρ.


The normal and tangential directions define the instantaneous plane of motion, also
known as the osculating plane. The velocity and acceleration vectors lie on this plane. The
orientation of osculating plane changes direction (twists), as the particle moves.

Example 2.4—Road Curvature Design


Roads that change direction have to be designed with a curvature. The amount of curvature
depends on the maximum normal acceleration that a vehicle can have and not slide. When
designing a curving road, two important considerations are the amount of curvature and
the variation of the curvature as the vehicle enters and leaves a curve.1

b
an 0
an v 2/ b

b
b
2
an v / b
Center of
curvature
an 0

FIGURE 2.9
Connecting two roads by a quarter circle.

Suppose you are designing a connection between two roads. The roads are perpendicular,
as shown in Figure 2.9. One way is to fit a quarter circle (say, of radius b) to connect the
roads. The disadvantage of this design is that, even for a vehicle moving with constant
speed, the lateral acceleration will be zero before entering the curve and v 2 /b immediately
1 Banking of the curve is another important factor. We will address this issue in the next chapter, when

discussing kinetics.
Kinematics Fundamentals 37

after. This jump in acceleration may not be sustained by the friction between the tires
and the road surface. Further, it contributes to shock (or jerk, defined earlier as the time
derivative of acceleration) and causes discomfort.
A wiser curved road design is one where the radius of curvature changes gradually. Two
examples of such a curve are shown in Figure 2.10. The slow initial change in curvature
is usually compensated for by a higher curvature in the middle of the curve. The second
curve in Figure 2.10, where the vehicle first turns away from the curve (seems contrary to
intuition), is another approach to increase the smallest radius of curvature. This counter
steer action is also what a bicycle rider (or speed skater) does when taking a turn.

Countersteer
path
an = small

an = 0
an= large

an = small

an = 0

FIGURE 2.10
Improved road curvature design.

Example 2.5
The motion of a point is described in Cartesian coordinates as x (t) = 2t2 + 4t, y (t) =
0.1t3 +cos t, z (t) = 3t. Find the radius of curvature and normal and tangential accelerations
at t = 0.
To find the radius of curvature, we need to first calculate the normal and tangential
directions, as well as the speed and the acceleration components in terms of the normal and
tangential coordinates. The position, velocity, and acceleration vectors are

r (t) = 2t2 + 4t i + 0.1t3 + cos t j + 3t k


 
[a]

v (t) = dr (t) /dt = (4t + 4) i + 0.3t2 − sin t j + 3k



[b]
a (t) = dv (t) /dt = 4i + (0.6t − cos t) j [c]
The speed is

q
2 2
v = v·v = (4t + 4) + (0.3t2 − sin t) + 9 [d]
and the unit vector in the tangential direction can be written as et = v/v. At t = 0 the
velocity and speed are
p
v (0) = 4i + 3k v (0) = 42 + 32 = 5 [e]
38 Applied Dynamics

so that the unit vector in the tangential direction is


v 4 3
et = = i+ k [f ]
v 5 5
The value of the acceleration at t = 0 is

a (0) = 4i − j [g]

The tangential acceleration can be obtained from

at (0) = a (0) · et = (4i − j) · (4i + 3k) /5 = 3.2 [h]

and the normal acceleration becomes

an = an en = a − at et = (4i − j) − 3.2 (4i + 3k) /5 = 1.44i − j − 1.92k [i]

The magnitude of the normal acceleration is


√ p
an = an · an = 1.442 + 1 + 1.922 = 2.6 [j]

so the unit vector in the normal direction is


an 1.44i − j − 1.92k
en = = [k]
an 2.6
Taking the dot product with the unit vector in the tangential direction, we can confirm
that the two unit vectors are orthogonal to each other.
The radius of curvature can be calculated using

v2 52
ρ = = = 9.615 [l]
an 2.6

2.5.3 Cylindrical Coordinates


Cylindrical coordinates and their two-dimensional counterpart, polar coordinates, are pre-
ferred when motion is along a curved path, the distance of a point from an origin is of
interest, and one component of the motion can be separated from the other two. A common
use of polar coordinates is in orbital mechanics.
Consider a point P and an inertial coordinate system XY Z with point O acting as the
reference point from which P is observed, as shown in Figure 2.11a. The position of point
P can be described by first taking the projection of P onto the XY plane, denoted by
P 0 . The distance ZP from points P to P 0 is along the vertical direction, and it is one of
the parameters describing the motion. Next, draw a line from point O towards P 0 and call
the direction of this line the radial direction. The radial direction is also denoted as the r
direction and the associated unit vector along is denoted by er .
The distance R from points O to P 0 is the second parameter describing the motion. The
third parameter is the angle between the radial direction and X axis and is denoted by θ
and measured in radians. The unit vector along the radial direction, er , is

er = cos θI + sin θJ (2.39)

The transverse direction lies on the XY plane and it is perpendicular to the radial
direction. Its positive direction is along the direction of a positive rotation of θ. The unit
Kinematics Fundamentals 39

  





  
 


 
 
    
 
  


FIGURE 2.11
Cylindrical coordinates. a) Path of an object, b) the XY and rθ axes.

vector along the transverse direction, denoted by eθ , is along this direction and it obeys the
rule er × eθ = K. From Figure 2.11b, eθ is

eθ = cos θJ − sin θI (2.40)

The position of point P is expressed in cylindrical coordinates as

rP = XP I + YP J + ZP K = Rer + ZP K (2.41)

The unit vectors er and eθ change direction as point P moves. To obtain the velocity,
we need to differentiate the above equation

vP = ṙP = Ṙer + Rėr + ŻP K (2.42)

which requires the derivative of the unit vector in the radial direction. To calculate this
derivative, consider the projection of the motion onto the XY plane and that the particle
has moved to point Q, whose projection is Q0 . Consequently, the coordinate system has
moved by ∆θ, as shown in Figure 2.12.
The unit vectors of the new coordinate system are denoted by er (θ + ∆θ) and
eθ (θ + ∆θ) and related to er (θ) and eθ (θ) by

er (θ + ∆θ) = er (θ) cos ∆θ + eθ (θ) sin ∆θ

eθ (θ + ∆θ) = −er (θ) sin ∆θ + eθ (θ) cos ∆θ (2.43)

Using a small angle assumption of sin ∆θ ≈ ∆θ, cos ∆θ ≈ 1, and taking the limit as ∆θ
approaches zero, the derivatives of the unit vectors become

er (θ + ∆θ) − er (θ) der


lim = = eθ
∆θ→0 ∆θ dθ

eθ (θ + ∆θ) − eθ (θ) deθ


lim = = −er (2.44)
∆θ→0 ∆θ dθ
40 Applied Dynamics

 
 








FIGURE 2.12
Polar coordinate system moved by ∆θ.

The time derivatives of the unit vectors become


der dθ deθ dθ
ėr = = θ̇eθ ėθ = = −θ̇er (2.45)
dθ dt dθ dt
which, when substituted in the expression for velocity in Equation (2.42), results in

v = Ṙer + Rθ̇eθ + ŻP K (2.46)

The first term on the right side corresponds to a change in the radial distance and the
second term to a change in angle.
In a similar fashion we can find the expression for acceleration. Differentiation of Equa-
tion (2.46) yields

aP = v̇P = R̈er + Ṙėr + Ṙθ̇eθ + Rθ̈eθ + Rθ̇ėθ + Z̈P K (2.47)

Substituting in the values for the derivatives of the unit vectors and combining terms gives
   
aP = R̈ − Rθ̇2 er + Rθ̈ + 2Ṙθ̇ eθ + Z̈P K (2.48)

We can attribute a physical meaning to the acceleration terms. The first term, R̈, de-
scribes the rate of change of the component of the velocity in the radial direction. The
second term, Rθ̇2 , is the centripetal acceleration. This term is always in the negative radial
direction, as R is always positive. The term Rθ̈ describes the acceleration due to a change
in the angle θ. The next term, 2Ṙθ̇, is known as the Coriolis acceleration, named after the
French military engineer Gustave G. Coriolis (1792–1843). The Coriolis acceleration is due
to two sources. Both deal with a changing distance in a rotating system.
Cylindrical coordinates are suitable to use when one component of the motion, which is
selected as the Z (or z) direction, is separable from the other two.

Example 2.6
For the mechanism in Figure 2.13, the crank is at an angle γ = 30◦ and is rotating at the
rate of γ̇ = 0.2 rad/s, which is increasing by γ̈ = 0.1 rad/s2 . The crank causes the slotted
Kinematics Fundamentals 41



 

 
 

FIGURE 2.13
Crank and slotted link.

link to rotate. Using cylindrical coordinates, calculate ṙ and r̈ associated with point P on
the slotted link for the special case when b = a.
As in any kinematics problem, the analysis begins with examining the position, continues
on to velocity analysis, and then to accelerations. Because OP B is an isosceles triangle,
θ = γ/2, so that
γ̇ γ̈
θ̇ = θ̈ = [a]
2 2
The polar coordinates for the slotted link r, θ, the normal-tangential coordinates for the
crank t, n and the inertial coordinates X, Y are shown in Figure 2.14. The length r can
be shown to be
r = 2a cos θ [b]

 




 





FIGURE 2.14
Coordinate systems for crank and slotted link (for when a = b).

The velocity at the tip of the crank is


vP = vet = aγ̇et = 2aθ̇et [c]
where v = aγ̇. The velocity of point P in terms of polar coordinates is
vP = ṙer + rθ̇eθ [d]
Equating Equations [c] and [d] leads to the expression for ṙ. From Figure 2.14, the unit
vector in the tangential direction becomes
et = − sin θer + cos θeθ [e]
42 Applied Dynamics

so that
vP = 2aθ̇et = −2aθ̇ sin θer + 2aθ̇ cos θeθ [f ]
and considering Equation [c] gives

2aθ̇ cos θ γ̇
ṙ = −2aθ̇ sin θ θ̇ = = [g]
r 2
Note that ṙ can also be obtained by direct differentiation of Equation [b].
To find the second derivatives of r and of θ, we can either differentiate the above equation,
which yields
γ̈
r̈ = −2aθ̈ sin θ + 2aθ̇2 cos θ θ̈ = [h]
2
or obtain the acceleration terms by equating the normal-tangential and polar components
of the acceleration. In normal-tangential coordinates, the acceleration is

v2
aP = v̇et + en [i]
ρ

where v̇ = aθ̈ and ρ = a, so that

aP = aθ̈et + aθ̇2 en [j]

The acceleration components in polar coordinates are


   
aP = r̈ − rθ̇2 er + rθ̈ + 2ṙθ̇ eθ [k]

and the value for θ can be obtained by relating the components of the unit vectors in the
two coordinate systems.
It should be noted that when b 6= a, the solution becomes much more complicated from
an algebraic point of view, as r, θ, and γ are related by

r cos θ = a + b cos γ r sin θ = b sin γ [l]

This example shows that we can obtain solutions to kinematics problems either by
selecting coordinate appropriate systems or by finding algebraic relationships that describe
the geometry and differentiating these equations.

2.5.4 Spherical Coordinates


Spherical coordinates express position in terms of one displacement and two angular co-
ordinates. An important use of spherical coordinates is describing the position of a point
on Earth in terms of the point’s latitude and longitude. The configuration of spherical
coordinates is shown in Figure 2.15a.
There are several different conventions used to define the principal directions associated
with spherical coordinates. We use here the distance R of the point from a reference point
O, and two angles θ and φ, referred to as the azimuthal and zenith angles, respectively. The
parameter R here (total distance from reference point) is different from the R (distance
from reference point to projection onto the XY plane) used in cylindrical coordinates. The
azimuthal angle is the same as the polar angle in cylindrical coordinates.
The principal directions are referred to as the radial, azimuthal, and zenith. The radial
direction connects reference point O and point P , with the positive direction as outward.
Kinematics Fundamentals 43

  





   


 
  

FIGURE 2.15
a) Spherical coordinate system, b) side view.

The corresponding unit vector is denoted by eR , so that the position vector for P has the
form

rP = ReR (2.49)

To define the azimuthal and zenith directions, it is necessary to first select and orient an
inertial XY Z coordinate system. In Earth geometry, the equatorial plane is the XY plane
with the Z axis towards the north. Projection of point P onto the XY plane is denoted by
P 0 . Next, rotate the XY Z coordinates about the Z axis by the azimuthal angle θ to get an
xyz coordinate system, noting that the x axis goes through point P 0 . In Earth geometry,
the azimuth angle is the longitude. The zenith angle φ is defined as the angle that the Z
axis makes with the radial direction, as shown in Figure 2.15b. In Earth coordinates, the
zenith angle φ is known as the colatitude or 90◦ minus the latitude.
The unit vector in the radial direction can be expressed in terms of the xyz coordinates
as

eR = sin φi + cos φk = sin φi + cos φK (2.50)

Noting that the unit vector in the x direction is

i = cos θI + sin θJ (2.51)

the unit vector in the radial direction in terms of the XY Z coordinates becomes

eR = sin φ cos θI + sin φ sin θJ + cos φK (2.52)

As shown in Figure 2.16, the unit vector in the azimuthal direction is selected as similar
to its counterpart in cylindrical coordinates, the polar direction, so that

eθ = j = − sin θI + cos θJ (2.53)

We can show that eR · eθ = 0, so the two unit vectors are orthogonal. The unit vector
associated with the zenith angle satisfies the relationship

eφ = eθ × eR = cos φ cos θI + cos φ sin θJ − sin φK = cos φi − sin φk (2.54)


44 Applied Dynamics

 



 


FIGURE 2.16
Top view of spherical coordinate system.

The unit vectors eR , eθ and eφ form a mutually orthogonal set with eR and eφ lying on
the xz (or xZ) plane. We need to obtain the derivatives of the unit vectors associated with
the spherical coordinates in order to calculate velocities and accelerations. The procedure
is tedious and only the results are stated here:
ėR = sin φθ̇eθ + φ̇eφ ėθ = −θ̇ (sin φeR + cos φeφ ) ėφ = −φ̇eR + cos φθ̇eθ (2.55)

d
vP = ṙP = (ReR ) = ṘeR + Rθ̇ sin φeθ + Rφ̇eφ (2.56)
dt
   
aP = R̈ − Rφ̇2 − Rθ̇2 sin2 φ eR + Rθ̈ sin φ + 2Ṙθ̇ sin φ + 2Rφ̇θ̇ cos φ eθ

 
+ Rφ̈ + 2Ṙφ̇ − Rθ̇2 sin φ cos φ eφ (2.57)

Example 2.7
The airplane in Figure 2.17 is traveling in the Y Z plane and, at the instant shown, it
is executing a maneuver so that it is at the bottom of a vertical loop that has a radius of
curvature of 1500 m. The speed of the airplane is constant at 550 km/hr. A radar is tracking
the airplane. Using spherical coordinates and a radar at point O, find the values of R̈ and
φ̈.
Since the azimuthal angle θ is not of interest (but its derivative is), without loss of
generality it can be set equal to zero in Figure 2.17, so that the xyz and XY Z coordinate
systems coincide. Shifting the reference point to O, as shown in Figure 2.18, OP P 0 forms a
5-12-13 triangle, so
φ = tan−1 (12/5) = 67.38◦ . The unit vectors are
12 5 5 12
eR = I+ K eθ = J eφ = eθ × eR = I− K [a]
13 13 13 13

and the radial distance is R = 12002 + 5002 = 1300 m.
The next step is velocity analysis. The velocity of the airplane is v = 550J km/h, and
in m/s it is
km km 1000 m 1 hr
550 = 550 = 152.77 m/s [b]
hr hr 1 km 3600 s
Kinematics Fundamentals 45





 


 
  







FIGURE 2.17
Airplane tracked by radar.

eR
P

5 e
X
O 12 P'

FIGURE 2.18
Side view, with shifted axes.

The velocity is in the Y direction. The unit vectors in the radial and zenith directions
do not have components in the Y direction. Comparing with Equation (2.56) leads to the
conclusion that
12
v · eR = Ṙ = 0 v · eθ = Rθ̇ sin φ = 1300θ̇ = 152.77 m/s v · eφ = Rφ̇ = 0 [c]
13
Solving for θ̇ gives θ̇ = 152.77/1200 = 0.1273 rad/s.
The acceleration analysis is next. The aircraft is moving with constant speed on a curved
path. Considering normal-tangential coordinates, the only component of the acceleration is
in the normal direction (Z axis) and
v2 152.772 2
a = an = K = 15.56K m/s [d]
ρ 1500
The components of the acceleration in the radial, azimuthal, and zenith directions are
 
12 5 2
a · eR = 15.56K · I+ K = 5.985 m/s a · eθ = 0
13 13
 
5 12 2
a · eφ = 15.57K · I− K = −14.36 m/s [e]
13 13
46 Applied Dynamics

From Equation (2.57) and considering from Equation [c] that Ṙ = 0, φ̇ = 0, the compo-
nents of the acceleration are
a · eR = R̈ − Rθ̇2 cos2 φ a · eθ = Rθ̈ cos φ a · eφ = Rφ̈ − Rθ̇2 sin φ cos φ [f ]
We solve for the second derivatives of R, θ, and φ by equating Equations [e] and [f]. The
azimuthal component yields θ̈ = 0, and the radial and zenith directions give
 2
2 2 2 12 2
R̈ = 5.985 + Rθ̇ cos φ = 5.985 + 1300 × 0.1273 × = 23.94 m/s
13
 
1 5 12 2
φ̈ = −14.36 + 1300 × 0.12732 × × = −0.0053 rad/s [g]
1300 13 13

2.6 Moving Reference Frames


Section 2.3 discussed transformation of a coordinate system by a rotation about a coordinate
axis. Then, we considered coordinate systems commonly used in dynamics. This section
extends the concept of coordinate transformations to the most general case.
Rotating coordinate systems are essential to the study of dynamics. There are several
cases where, either by choice or by necessity, we need to use a coordinate system that
rotates to describe motion. For example, when describing certain motions in the vicinity of
the Earth, such as hurricane formation and satellite launching, it is necessary to consider
the rotation of the Earth in the mathematical model. The treatment of rigid body motion
also is facilitated by the use of rotating coordinate frames.
We begin by exploring relationships between different reference frames and associated
unit vectors. Two notations will commonly be used to distinguish between different reference
frames:
1. Denote one of the frames by XY Z and the other by xyz with unit vectors I, J, K and
i, j, k, respectively. In general, XY Z denotes a fixed frame and xyz one that moves.
Intermediate frames are usually referred to as X 0 Y 0 Z 0 , X 00 Y 00 Z 00 , and so on.
2. In the second description, which is especially useful when several frames are involved, a
letter is assigned to each frame: for example, frames A and B. The coordinate axes of
the frames are called a1 a2 a3 and b1 b2 b3 and the unit vectors along the coordinate axes
are defined as a1 , a2 , and a3 and b1 , b2 , and b3 . Intermediate frames are defined in
a similar fashion.
Consider a vector q as viewed in the B frame. This vector can be expressed as
q = q1 b1 + q2 b2 + q3 b3 (2.58)
in which q1 , q2 , and q3 are the components of q. Because the unit vectors are orthogonal,
we can express each component as qj = q · bj , (j = 1, 2, 3), so that
q = (q · b1 ) b1 + (q · b2 ) b2 + (q · b3 ) b3 (2.59)
The vector q can be resolved in the A frame, as well. However, referring to the compo-
nents as qi (i = 1, 2, 3) will not make it possible to distinguish them from their counterparts
associated with the B frame. Rather, these components can be expressed as
A
q = q 1 a1 + A q2 a2 + A q3 a3 = B
q1 b1 + B q2 b2 + B q3 b3 (2.60)
Kinematics Fundamentals 47

so that in column vector format q becomes


A T T
q = A q1 A A
B
q = B q1 B B
 
q2 q3 q2 q3 (2.61)

This notation is similar to the notation used in Sec. 2.3 with XY Z and xyz as the coordinate
axes.

b3

q = qe e

3 q
e 2
b2
1

b1

FIGURE 2.19
Direction angles of a vector.

It is of interest to investigate relationships between the unit vectors of two coordinate


systems. Consider the vector q and the unit vector e along it, so that q = qe, where q is the
magnitude of q. The angles that q makes with the axes of a coordinate system are denoted
by θ1 , θ2 , and θ3 and are called direction angles, as shown in Figure 2.19. The direction
cosines of q (and of e) are denoted by c1 , c2 , and c3 , and they are the cosines of the angles
that e makes with the coordinate axes

e = cos θ1 b1 + cos θ2 b2 + cos θ3 b3 (2.62)

The direction cosines are written as cj = e · bj = cos θj , (j = 1, 2, 3), so that q = qc1 b1 +


qc2 b2 + qc3 b3 .
Now, consider the unit vectors of the A and B frames. Considering the above equation,
we can write

bj = (bj · a1 ) a1 + (bj · a2 ) a2 + (bj · a3 ) a3 j = 1, 2, 3

aj = (aj · b1 ) b1 + (aj · b2 ) b2 + (aj · b3 ) b3 j = 1, 2, 3 (2.63)

The direction cosine between two coordinate axes aj and bk is the cosine of the angle
between the two axes and is denoted by cjk = aj · bk = cos θjk (j, k = 1, 2, 3) . Define the
column vectors containing the unit vectors of the two frames as {a} and {b}, where
T T
{a} = [a1 a2 a3 ] {b} = [b1 b2 b3 ] (2.64)

and the direction cosine matrix [c]


 
c11 c12 c13
 c21 c22 c23  (2.65)
c31 c32 c33
48 Applied Dynamics

from which we can show that

{b} = [c]T {a} (2.66)

Equation (2.66) can be inverted to yield

{a} = [c]−T {b} (2.67)

Comparing Equation (2.67) with Equation (2.66) leads to the conclusion that the direc-
tion cosine matrix is unitary (also called orthonormal), that is, its inverse is equal to its
transpose, or

[c]−1 = [c]T [c] [c]T = [1] (2.68)

where [1] is the identity matrix. Note that the determinant of a general unitary matrix is
±1, but for the direction cosine matrix det [c] = 1.
Equation (2.66) also applies when relating the components of a vector resolved along
the axes of two coordinate systems. Indeed, consider the vector q and its column vector
representation in the A and B frames. We can show that
B
q = [c]T A q
A
q = [c] B q
 
(2.69)

The definition of direction cosine above is not universally accepted. Some texts instead
define the direction cosine by cjk = bj · ak = cos θkj (j, k = 1, 2, 3).
A total of nine direction cosines are defined, one for each angle between the j-th and k-th
coordinate axes (i, j, k = 1, 2, 3). These nine direction cosines are not independent of each
other. Equation (2.68) represents six independent equations that relate the direction cosines
(six because of the symmetry of [c] [c]T ), reducing the number of independent direction
cosines to three.2 It follows that, at most, three independent parameters are necessary to
represent the transformation from any given configuration of coordinate axes to another
one. Issues that are of interest are the following:

• How do we select these three rotation parameters?


• Given the three parameters, how do we calculate the direction cosine matrix?
• Given the direction cosine matrix, what are the three rotation parameters associated
with that matrix?

2.7 Selection of Rotation Parameters


This section considers two approaches for selecting the rotation parameters. The first in-
volves three rotations about independent axes. The second approach uses a single rotation
about an axis whose orientation is specified.
2 We can demonstrate this by writing [c] as three column vectors [{c }{c }{c }]. These vectors are
1 2 3
orthogonal vectors and they represent the direction angles of the axes of the transformed coordinates. It
follows that Equation (2.68) represents the six possible dot products among these vectors.
Kinematics Fundamentals 49

2.7.1 Transformation by Three Rotation Angles


In the first approach, the three parameters are selected as three independent rotations. This
necessitates selection of the axes about which these rotations are made. To this end, there
are infinite choices. These axes are selected in a way to simplify calculations.
The choices of axes about which rotations are made are narrowed down by carrying out
the rotations about the axes of coordinate frames. Denote the initial position of the axes by
frame A, with axes a1 a2 a3 , and rotate frame A counterclockwise by angle θ about one of the
axes to get a rotated frame B, whose axes are b1 b2 b3 . This rotation convention ensures that
det [c] = 1. Rotation about the a1 axis is called a 1 rotation, about the a2 axis a 2 rotation
and rotation about the a3 axis a 3 rotation. These rotations are illustrated in Figure 2.20.

a) b) c)
b3 a3 b3 a3 a3, b3

b2
b2
a2 a2 , b2 a2

a1, b1 a1 a1
b1 b1

FIGURE 2.20
Rotation types: a) a 1 rotation; b) a 2 rotation; c) a 3 rotation.

For a 1 rotation the unit vectors of the A and B frame are related by

b1 = a1 b2 = cos θa2 + sin θa3 b3 = − sin θa2 + cos θa3 (2.70)

or, in matrix form,


−1
{b} = [c] {a} = [R] {a} {B q} = [R] {A q} (2.71)

where [R] = [c]−1 is called the rotation matrix and has the form
 
1 0 0
For a 1 rotation [R] =  0 cos θ sin θ  (2.72)
0 − sin θ cos θ

When the rotation is performed about the a2 axis, that is, a 2 rotation, the rotated
vectors become

b1 = cos θa1 − sin θa3 b2 = a2 b3 = sin θa1 + cos θa3 (2.73)

and the rotation matrix has the form


 
cos θ 0 − sin θ
For a 2 rotation [R] =  0 1 0  (2.74)
sin θ 0 cos θ
50 Applied Dynamics

For a 3 rotation, the rotated vectors are

b1 = cos θa1 + sin θa2 b2 = − sin θa1 + cos θa2 b3 = a3 (2.75)

and the rotation vector is


 
cos θ sin θ 0
For a 3 rotation [R] =  − sin θ cos θ 0 (2.76)
0 0 1

For all three rotations above, the determinant of the rotation matrix is 1, or det[R] = 1.
This verifies our earlier statement that the determinant of the direction cosine matrix is
unity.
Consider now the three rotation angles needed for a general rotation sequence and how
we can accomplish the three rotations. To this end, two approaches can be identified: body-
fixed rotation sequence and space-fixed rotation sequence.

Body-Fixed Rotation Sequence


This rotation sequence can be visualized by considering a box and performing the rotations
about a set of axes attached to the box. Begin with an initial coordinate set a1 a2 a3 , align
the box with it, and rotate the box about one of the axes by an angle θ1 . Let us call the
resulting orientation of the box the D frame with the axes d1 d2 d3 , which are now aligned
with the box. There are three choices in the selection of the rotation axis.
The procedure is shown in Figure 2.21a for a 3 rotation, that is, a rotation about the a3
axis. The associated rotation matrix is denoted by [R1 ]. Note that the subscript 1 denotes
that this is the first rotation and not the axis about which the rotation takes place. As a
result of the rotation, points P QR on the box move to their new positions P 0 Q0 R0 in Figure
2.21b. These positions do not change relative to the box.

 
 

  
   


 
   


 
 

FIGURE 2.21
Body-fixed rotation about a3 : a) initial position, b) after rotation by θ.

Next, rotate the box about one of the d1 d2 d3 axes by an angle θ2 to obtain the new
orientation of the box, denoted as the H frame, with axes h1 h2 h3 . The rotation matrix
associated with this rotation is denoted by [R2 ]. Two axes can be chosen for this rotation.
If the first rotation is carried out about, say, a3 , then d3 = a3 and the second rotation can
only be carried out about d1 or d2 . Otherwise, the two rotations cannot be distinguished
from each other.
Kinematics Fundamentals 51

The third rotation is carried out by rotating the box about one of the axes h1 h2 h3 by
θ3 to obtain the final frame B with axes b1 b2 b3 . For the same reason described above, there
are two axes to rotate about. The associated rotation matrix is [R3 ]. It follows that there
are 3 × 2 × 2 = 12 different ways to carry out a body-fixed rotation sequence. These possible
ways of selection are known as Euler angle sequences, and they are denoted by the number
of the axes about which the rotations are made.
For example, if the first rotation is about the a3 axis, the second rotation is about the
d1 axis, and the third rotation is about the h2 axis, the rotation sequence is called 3-1-2.
Chapter 9 will quantify rotation sequences and discuss their applications in more detail.
Consider next the combined rotation. For the three rotations above

{d} = [R1 ] {a} {h} = [R2 ] {d} {b} = [R3 ] {h} (2.77)

Combining the three rotations gives

{b} = [R3 ] [R2 ] [R1 ] {a} = [R] {a} (2.78)

and the combined rotation matrix is [R] = [R3 ] [R2 ] [R1 ].

Space-Fixed Rotation Sequence


Here, we consider an initial coordinate system and perform all three rotations about the
axes of the initial coordinates. Let a1 a2 a3 be the initial axes. The first rotation is about
one of a1 , a2 , or a3 axes to yield the d1 d2 d3 axes, with {d} = [R1 ]{a}. Then, the coordinate
axes d1 d2 d3 are rotated about one of the a1 , a2 , or a3 axes, but not about the same axis
about which the first rotation was made. Note that the axis for the second rotation is not
one of the axes of the rotated frame. The third rotation is carried out about another one of
a1 , a2 , or a3 .
We can show that the combined rotation matrix has the form

{b} = [R] {a} = [R1 ] [R2 ] [R3 ] {a} (2.79)

where the combined rotation matrix is in reverse order of the body-fixed transformations.
Space-fixed rotations are not used as frequently as body-fixed rotations in dynamics.

Inversion
We have obtained values for the rotation matrix [R] (and hence direction cosine matrix
[c]) in terms of three rotation angles. The inverse problem, that is, given [R] or [c] to find
the three rotation angles, is not as straightforward. To solve the inverse problem one needs
to be given, in addition to [R] (or [c]), the sequence under which the individual rotation
transformations are made. Given this information, the rotation matrix [R] is constructed
and the unknowns are solved for. In general, we look at the general form of [R] and begin
with the entries of [R] that are the simplest to solve for.

2.7.2 Resolving a Rotated Vector


So far, we have developed two coordinate systems and looked at the same vector in the
two coordinate systems. That is, if the coordinates of a particular vector are {A q} =
[A q1 A q2 A q3 ]T in the A frame, and the coordinates of that same vector are {B q} =
[B q1 B q2 B q3 ]T in the B frame, then
B
q = [R] A q

(2.80)
52 Applied Dynamics

a) b)
a3 b3 a3
P'
E P b2
E' qf C'
qi
O a2 O a2
C

a1 a1

FIGURE 2.22
a) Initial position of vector, b) rotated vector.

Another problem of interest is to take a vector, to rotate the vector, and to find the
coordinates of the rotated vector in the original coordinate system. Consider the A frame
and attach a box (or a rectangle) to it, as shown in Figure 2.22a. Let us select point P
on the box and associated vector qi or {A qi }. Next, we rotate the vector qi (or the box to
which the vector is attached).
Figure 2.22b shows a rotation by θ about the a1 axis. The rotated vector is denoted by
qf or {A qf }, and this vector connects points O and P 0 , where P 0 is the point to which P
moves after the rotation. Note that rotating the vector is equivalent to rotating the box (or
rectangle) in which the vector is defined.
Consider now a second frame B, which is obtained by rotating the A frame the same
way the vector qi was rotated. The initial and rotated vectors can also be expressed in
terms of the B frame as {B qi } and {B qf }. We relate the initial and final vectors in the two
coordinate systems by
B
qi = [R] A qi
B
qf = [R] A qf
 
(2.81)
Because the vector is attached to the reference frame, the initial vector in the A frame
has the same coordinates as the rotated vector in the B frame. Hence,
B
qf = A qi

(2.82)
Combining the above two equations, the initial and final positions of the vector in the initial
A frame are expressed as
A
qf = [R]T A qi

(2.83)
Let us compare Equations (2.80) and (2.83). In Equation (2.80) the vector is fixed, but
is viewed from two different coordinate frames. In Equation (2.83) the vector is rotated,
but its initial and final locations are viewed from the same reference frame. In one case the
transformation matrix is [R] and in the other it is [R]T , denoting inverse transformations. It
follows that viewing a fixed vector from a rotated frame is exactly the opposite of rotating
that vector (by the same amount) and viewing the rotated vector from the fixed frame.

2.7.3 Single Rotation about a Specified Axis


This rotation approach is based on Euler’s theorem, which states that the most general
transformation of a rigid body with one fixed point can be described as a single rotation
Kinematics Fundamentals 53

about a certain axis going through the fixed point. The axis about which the rotation is
made is called the principal line and is denoted by n, as shown in Figure 2.23a. The unit
vector along this axis is expressed as n. The rotation angle is called the principal angle and
is denoted by Φ.

a) b) c)
n
n
O a2 (or a'2 )
P' O O
P P (or P')
qi qf a'2 a'1
a2
P'
P a1

FIGURE 2.23
a) General rotation of a vector about an axis n by Φ, b) top view, c) side view.

The direction cosines of the axis of rotation {n} = [n1 n2 n3 ]T are not independent and
they obey the relationship n21 + n22 + n23 = 1. Given two direction cosines, the third can be
ascertained to within a ± value. Hence, only two of the parameters are independent. The
rotation angle Φ becomes the third parameter. When the sign of the third direction cosine
is changed, this means the principal line is pointing in the opposite direction and the same
transformation can be achieved by reversing the sign of Φ.
Given a vector qi and rotating it by Φ about axis n, the rotated vector, denoted by qf ,
can be expressed as

qf = cos Φqi + (1 − cos Φ)(qi · n)n + sin Φn × qi (2.84)

or, in column vector notation

{qf } = cos Φ {qi } + (1 − cos Φ) {n}{n}T {qi } + sin Φ [ñ] {qi } (2.85)

or

cos Φ [1] + (1 − cos Φ){n}{n}T + sin Φ [ñ] {qi }


 
{qf } = (2.86)

where [ñ] is the skew-symmetric matrix associated with the column vector {n}.
The validity of the above relationship can be demonstrated by defining a coordinate
system with directions n, a1 , a2 and rotating the coordinate system about n by Φ. Referring
to the rotated coordinate system as n, a01 , a02 , as shown in Figure 2.23b, the two sets of
vectors are related by

a01 = cos Φa1 + sin Φa2 a02 = − sin Φa1 + cos Φa2 (2.87)

The angle θ is not a rotation parameter; rather, it is the angle between the principal
line and the vector q. Figure 2.23c shows the side view.
Now, consider the initial and rotated vectors qi and qf and denote the magnitude of
54 Applied Dynamics

these vectors by q. Because qf is obtained by rotating qi about n, the two vectors have the
same components along the initial and rotated axes. These vectors are written as
qi = q cos θn − q sin θa2 qf = q cos θn − q sin θa02 (2.88)
Substitution of the value for a02 from the above equation leads to
qf = q cos θn + q sin θ sin Φa1 − q sin θ cos Φa2 (2.89)
The above equation can be rewritten as
qf = q cos Φ (cos θn − sin θa2 ) + q (1 − cos Φ) cos θn + q sin θ sin Φa1 (2.90)
The first term following the equal sign in the above equation is recognized as cos Φqi .
Noting that cos θ = qi · n, the second term can be written as (1 − cos Φ)(qi · n)n. Finally,
we note that
sin Φn × qi = sin Φn × q (cos θn + sin θa2 ) = q sin Φ sin θa1 (2.91)
The conclusion is that Equation (2.84) and Equation (2.90) are equivalent.
Equations (2.85)–(2.86) give the relationship between the initial and final positions of a
vector in column vector format. Hence, considering an initial frame and rotating it by angle
Φ about n to get a rotated frame, the direction cosine matrix between the two frames has
the form
[c] = [R]T = cos Φ [1] + (1 − cos Φ) {n}{n}T + sin Φ [ñ] (2.92)
which is an additional way of calculating the rotation matrix. Given a rotation matrix [R],
calculation of the associated principal line and rotation angle will be discussed in Chapter
9.

2.7.4 Finite Rotations Do Not Commute


The preceding analysis leads to the conclusion that the order in which rotations of coor-
dinates are performed makes a difference in the orientation of the transformed coordinate
system. This holds true whether the rotations are performed as a body-fixed rotation se-
quence or as space-fixed rotations. We can verify this visually by taking a book and rotating
it about two axes in different sequences.
The procedure is illustrated in Figure 2.24 for a body-fixed rotation sequence. Begin
with the XY Z frame, where the book lies on the XZ plane. Rotate the book by 90◦ about
the X axis to get the X 0 Y 0 Z 0 coordinates and then rotate about the Z 0 axis by 90◦ to get
the xyz coordinates. Repeating the procedure with first rotating about the Z and then X 0
coordinates leads to a different orientation of the book.
The conclusion is that sequences of finite rotations in three dimensions cannot be ex-
pressed as vectors, as the commutativity rule does not hold. It follows that there does not
exist an angular position vector to differentiate in order to obtain angular velocity, except
for the special case of plane motion. The subsequent sections will provide definitions for the
angular velocity of a reference frame.

Example 2.8
Given the rotation matrix
 
−0.1768 0.8839 −0.4330
[R] =  −0.9186 −0.3062 −0.2500  [a]
−0.3536 0.3536 0.8660
Kinematics Fundamentals 55

  

  
  
 



   


 
 

  
  

  

  

  




FIGURE 2.24
Finite rotations do NOT commute.

find the rotation angles if the coordinate axes are obtained by a 3-1-3 transformation.
For a 3-1-3 transformation with angles θ1 , θ2 , and θ3 , the combined rotation matrix is
expressed as
[R] = [R3 ][R2 ][R1 ] [b]
where    
cos θ1 sin θ1
0 1 0 0
[R1 ] =  − sin θ1 cos θ1
0 [R2 ] =  0 cos θ2 sin θ2 
0 0 1 0 − sin θ2 cos θ2
 
cos θ3 sin θ3 0
[R3 ] =  − sin θ3 cos θ3 0  [c]
0 0 1
Carrying out the algebra, the combined rotation matrix has the form
 
c3 c1 − s3 c2 s1 c3 s1 + s3 c2 c1 s3 s2
[R] =  −s3 c1 − c3 c2 s1 −s3 s1 + c3 c2 c1 c3 s2  [d]
s2 s1 −s2 c1 c2

where the compact notation is used, where s denotes the sine and c to denotes the cosine
functions. For example, c1 = cos θ1 .
When proceeding to identify the rotation angles, it is convenient to begin with the
simplest expression, which in this case is R33

R33 = cos θ2 = 0.8660 =⇒ θ2 = ±30◦ [e]


56 Applied Dynamics

where the rotation angles in the range of −180◦ to 180◦ . Comparing the other simple
elements in the rotation matrix yields
R13 −0.433 √
= tan θ3 = = 3 =⇒ θ3 = 60◦ or − 120◦ [f ]
R23 −0.250
R31
= − tan θ1 = −1 =⇒ θ1 = 45◦ or − 135◦ [g]
R32
Begin with the assumption that θ2 = 30◦ . Examining the values of R31 and R13 leads
to
R31 = sin θ2 sin θ1 = −0.3536 R13 = sin θ3 sin θ2 = −0.4330 [h]
It follows that the sines of both these angles are negative, and we conclude from Equations
[f] and [g] that
θ1 = −45◦ θ2 = 30◦ θ3 = −120◦ [i]
The next step is to check the accuracy of the assumption that we made. Calculating any
one of the remaining elements of [R], say R11 , gives
√ √ √ √
1 2 3 3 2
R11 = cos θ3 cos θ1 − sin θ3 cos θ2 sin θ1 = − + = 0.1768 [j]
2 2 2 2 2
But R11 = −0.1768, so the assumption made earlier about θ2 is incorrect. It follows
that θ2 = −30◦ . Following the same procedure as in Equation [h] leads to
R31 = sin θ2 sin θ1 = −0.3536 R13 = sin θ3 sin θ2 = −0.4330 [k]
and the conclusion is
θ1 = 45◦ θ2 = −30◦ θ3 = 120◦ [l]
Substituting these values into R11 or in any other element of [R], the correctness of
results in Equation [l] is confirmed.

Example 2.9
An xyz coordinate system is obtained by rotating the XY Z coordinates about a line n that
makes an angle of 30◦ with the X axis, 60◦ with the Y axis and is perpendicular to the
Z axis, as shown in Figure 2.25. Calculate the direction matrix [c] between XY Z and xyz
when the rotation angle is Φ = −66◦ . Also, find the coordinates after rotation of point P ,
whose position before rotation is given by rP = 2I.
The direction cosines of the principal line are

3 1
nX = nY = nZ = 0 [a]
2 2
and the associated [ñ] matrix is
   
0 −nZ nY 0 0 1
1 √ 
[ñ] =  nZ 0 −nX  =  0 0 − 3 [b]
 
2 √
−nY nX 0 −1 3 0
Noting that cos Φ = cos (−66◦ ) = 0.4067, sin Φ = sin (−66◦ ) = −0.9135, and using Equa-
tion (2.92) results in
   √   √ T
1 0 0 3 3
1
[c] = 0.4067 0
 1 0  + 0.5933 ×  1   1 
4
0 0 1 0 0
Kinematics Fundamentals 57

n
60°
n
30°
X
90° P
rP
Z

FIGURE 2.25
Rotation axis n.

 
0 0 1
 
√  0.8517 0.2569 −0.4568
1
−0.9135 ×  0 0 − 3  =  0.2569 0.5551 0.7912  [c]
2 √
−1 3 0 0.4568 −0.7912 0.4067

To find the location of point P after the rotation, we can use the relationship {qf } =
[c]{qi }, where {qi } = [2 0 0]T , with the result
 
0.8517 0.2569 −0.4568
  
2 1.7034
{qf } = [c]{qi } =  0.2569 0.5551 0.7912   0  =  0.5138  [d]
 
0.4568 −0.7912 0.4067 0 0.9136

Note that the rotated vector {qf } is expressed in terms of the XY Z frame.
The same result can also be obtained by successive coordinate transformations. Noting
that for body-fixed transformations the rotations are conducted about one of the coordinate
axes, the XY Z frame is rotated so that one of the axes of the rotated frame is the n axis.
This can be accomplished by rotating about the Z axis by 30◦ and calling the rotated axes
X 0 Y 0 Z 0 . The associated rotation (a 3 rotation) matrix is
√ 
3/2 1/2 0

[R1 ] =  −1/2 3/2 0  [e]
 

0 0 1

and X 0 axis is the same as the n axis in Figure 2.25. Then, rotate the X 0 Y 0 Z 0 frame and the
vector rP = 2I about the X 0 axis by −66◦ to get the xyz frame. The associated rotation (a
1 rotation) matrix is  
1 0 0
[R2 ] =  0 0.4067 −0.9135  [f ]
0 0.9135 0.4067
The position of the vector rP after the rotation, in terms of the X 0 Y 0 Z 0 frame, is
0
Y 0Z0 T 0
Y 0Z0
{X qf } = [R2 ] {X qi } [g]

Noting that
X0
   
X
 Y 0  = [R1 ]  Y  [h]
Z0 Z
58 Applied Dynamics
0
Y 0Z0
XY Z
so that {X qf } = [R1 ] qf , Equation [g] can then be expressed in the XY Z frame
as XY Z T
= [R1 ]T [R2 ] [R1 ]
XY Z
qi = [R]T XY Z qi

qf [i]
where the rotation matrix between the initial and final frames is

[R] = [R1 ]T [R2 ] [R1 ] [j]

The above result can be explained by noting that the first rotation [R1 ] merely shifts to
a different coordinate frame but does not rotate the vector rP . The second rotation rotates
the vector rP . So, the two rotations are completely different in nature. Carrying out the
matrix multiplications, Equation [j] gives the same results as Equations [c] and [d].

2.8 Rate of Change of a Vector, Angular Velocity


The previous section demonstrated that consecutive rotations of coordinate frames by finite
angles do not lend themselves to representation as vectors. Hence, we do not have a vector
to differentiate in order to represent rotation rates. This section explores ways of looking at
rotation rates and defining the angular velocity vector.
When looking at the rate of change of a quantity, we must distinguish between derivatives
taken in different reference frames. For example, consider a moving vehicle and attach a
reference frame to the vehicle. When something inside the vehicle moves, say a passenger
throws a ball up and down, someone inside the vehicle sees the ball as moving up and down
only. An observer outside the vehicle sees the ball moving with the vehicle as well as moving
with respect to the vehicle.
We can define rates of rotations and angular velocities in a variety of ways. One is to
use infinitesimal values for the angles in a body-fixed (or space-fixed) rotation sequence.
This approach is intuitive, but not mathematically sound. Two more rigorous approaches
are presented here; one that makes use of column vector notation and one that is based on
Euler’s theorem. This section begins with the definition of angular velocity for plane motion
and then considers the three-dimensional case.

2.8.1 Angular Velocity for Plane Motion


Consider the two reference frames XY Z and xyz in Section 2.3, where the XY Z coordinates
are rotated by θ about the Z axis to arrive at the xyz coordinate system. Let us consider
XY Z to be a fixed reference frame and evaluate the time derivative of the unit vectors
associated with the moving frame xyz. Differentiating Equation (2.2) leads to

d d
i = θ̇ (− sin θI + cos θJ) = θ̇j j = θ̇ (− cos θI − sin θJ) = −θ̇i (2.93)
dt dt
d
and dt k = 0, as the z direction is the same as the Z direction, which is fixed. The above
relationships can be expressed in terms of the cross product involving the angular velocity
vector ω = ωk = θ̇k and by writing
d d d
i = ω × i = θ̇k × i = θ̇j j = ω × j = −θ̇i k = 0 (2.94)
dt dt dt
Indeed, when dealing with a single rotation about a fixed axis, the rotation can be
Kinematics Fundamentals 59

 

 

 





FIGURE 2.26
Cross products of unit vectors in xy plane.

represented as a vector, θk. Taking the derivative of this vector gives the angular velocity
ω = θ̇k. The cross products above are demonstrated in Figure 2.26.
The rate of change of a vector in the fixed and moving frames is of interest. Consider
first a vector that is fixed in the xyz frame. This vector can be any type of vector, such
as position, velocity or acceleration, linear or angular momentum, or angular velocity. To
an observer attached to the xyz frame, this vector does not move. As the xyz coordinate
system is moving, an observer in the XY Z system sees the vector move.
The vector of interest is q = qx i + qy j = qX I + qY J. Because this vector is fixed in the
xyz frame, its time derivative becomes
d
q = q̇ = qx i̇ + qy j̇ = qx ω × i + qy ω × j = ω × q (2.95)
dt
The rate of change of a vector that is fixed in a moving frame is obtained by the cross product
between the angular velocity of the moving frame and the vector itself. This relationship is
valid for both two- and three-dimensional motion.
Next, we consider the case when the vector q is not fixed in the moving frame xyz but
instead is changing with respect to the moving frame. The derivatives q̇x , q̇y are no longer
zero and
d
q = q̇x i + q̇y j + ω × q (2.96)
dt
Denoting by q̇rel = q̇x i + q̇y j the local derivative or relative rate of change of q, or the rate
of change of the vector q in the moving frame, we can write the above equation as
d
q = q̇ = q̇rel + ω × q (2.97)
dt
The total derivative of a vector comprises of two components: i) the change of the vector
as viewed from the moving frame, plus ii) the change due to the rotation of the moving
frame. The above equation is known as the transport theorem, and it is valid for any
vector observed from a moving (rotating) frame. The name transport reflects the fact that
the derivative is being “transported” from one reference frame to another. The transport
theorem is depicted in Figure 2.27.

2.8.2 Angular Velocity for Three-Dimensional Motion


As discussed earlier, for three-dimensional rotation, the cumulative effect of rotation se-
quences cannot be described as vectors, and thus, no rotation vector exists for us to dif-
60 Applied Dynamics

q =×q+q rel

×q

q

x
q rel

FIGURE 2.27
Derivative of a vector using transport theorem.

ferentiate. Instead, we define angular velocity for three-dimensional motion by means of a


column vector formulation.
Begin with the XY Z and xyz frame representations. The frame xyz rotates with respect
to the XY Z frame. Consider a vector q that is represented in terms of the two reference
frames as

q = qx i + qy j + qz k = qX I + qY J + qZ K (2.98)

Using the notation in Section 2.3, the vector q in column vector format has the form
   
XY Z qX qx
q =  qY  {xyz q} =  qy  (2.99)
qZ qz

These two column vector representations are needed to indicate that the vector is ex-
pressed in two different reference frames. The two vectors are related by

{xyz q} = [R] XY Z q
XY Z
q = [R]T {xyz q}

or (2.100)

where [R] is the 3×3 rotation matrix. The form for [R] for plane motion is given in Equation
(2.4).
Differentiation of the expression on the right in Equation (2.100) results in
XY Z
q̇ = [Ṙ]T {xyz q} + [R]T {xyz q̇} (2.101)

The term {xyz q̇} is the derivative of the vectoras viewed from the moving frame. The term
[Ṙ] is the derivative of the rotation vector, and XY Z q̇ is the derivative of the vector viewed
from the nonmoving (inertial) frame, that is, the total derivative. Introducing the notation
d
for the derivative of qas v = q̇ = dt q, the rate of change vector v in terms of the
 XY Z and
XY Z
xyz frames becomes v and {xyz v}, respectively. This way, {xyz v} = [R] XY Z v .
Note that while the relationship
XY Z
v = XY Z q̇

(2.102)

is valid, the similar-looking counterpart for xyz frame is not:

{xyz v} 6= {xyz q̇} (2.103)


Kinematics Fundamentals 61

This is because {xyz v} describes the rate of change vector in terms of the coordinates of the
moving frame, while {xyz q̇} denotes the local derivative of {xyz q} as viewed from the moving
(rotating) frame. Introducing Equation (2.101) to Equation (2.102) and left multiplying by
[R] gives

[R] XY Z v = {xyz v} = [R][Ṙ]T {xyz q} + {xyz q̇}



(2.104)

Let us examine the matrix product [R][Ṙ]T more closely. For the special case of plane
motion (i.e., two-dimensional)
     
cos θ sin θ − sin θ − cos θ 0 −1
[R][Ṙ]T = θ̇ = θ̇ (2.105)
− sin θ cos θ cos θ − sin θ 1 0

which is recognized as the matrix representation [ω̃] of the angular velocity ω = θ̇k in the
xyz frame. It follows that the above equation is yet another way to define the angular
velocity vector, and we can write

{xyz v} = [ω̃] {xyz q} + {xyz q̇} (2.106)

which is the representation of the transport theorem in column vector notation. Note that
all of the terms in the above equation are expressed in terms of parameters associated with
the xyz frame.
The same procedure can be carried out for the rotation matrix in three dimensions.
Regardless of the rotation sequence, we can show that [R][Ṙ]T is a skew-symmetric matrix.
Consider the identity [R] [R]T = [1] and differentiate it, which yields
T
[R] [Ṙ]T + [Ṙ] [R] = [0] (2.107)

The two terms in the preceding equation are transposes of each other. Indeed, denoting
T
by [W ] = [R] [Ṙ]T , it follows that [W ]T = [Ṙ] [R] . A matrix which, when added to its
transpose, yields a null matrix must be skew-symmetric. The matrix [W ] can be expressed
as
 
0 −ωz ωy
[W ] = [R] [Ṙ]T =  ωz 0 −ωx  (2.108)
−ωy ωx 0

Since [W ] is skew-symmetric, it can be recognized as the matrix representation of a


vector ω = ωx i + ωy j + ωz k used when expressing a cross product, [W ] = [ω̃]. We refer to
[W ] = [ω̃] as the angular velocity matrix.
The vector ω is defined as the angular velocity vector of the reference frame xyz with
respect to the XY Z frame, and ωx , ωy , and ωz are the instantaneous angular velocities
or components of the angular velocity vector. Thus, Equation (2.104) is verified for three-
dimensional motion.
It must be emphasized that, for three-dimensional rotations, the angular velocity vector
is a defined quantity and that it is not the derivative of another vector. For this reason,
the angular velocity vector is referred to as nonholonomic, a term that is associated with
expressions that cannot be integrated to another expression. The way we arrive at the
angular velocity vector is completely different from the derivation of the expression for
translational velocity or the rate of change of any defined vector.
When using the A and B frame notation, the angular velocity vector of the B frame
with respect to the A frame is written as Aω B , where the superscripts denote the frames
that are related. The angular velocity of the A frame with respect to the B frame is B ω A
62 Applied Dynamics

and B ω A = −Aω B . In addition, the derivative of a vector obtained in a certain reference


d
frame is denoted by a left superscript, such as B dt q or B v. It follows that the transport
theorem can be written as
A d B d A
q = q + ωB × q (2.109)
dt dt

2.8.3 Other Definitions of Angular Velocity


The definition of angular velocity in the preceding subsection is not the only way angular
velocity can be defined. Two additional definitions are presented here. Consider the frames
A and B. The angular velocity of frame B with respect to frame A is defined by
     
A B
ω = ω1 b1 + ω2 b2 + ω3 b3 = ḃ2 · b3 b1 + ḃ3 · b1 b2 + ḃ1 · b2 b3 (2.110)

This definition can be verified by analyzing the expressions for the rates of change of the unit
vectors. While the above definition is more abstract than the way we arrived at Equation
(2.108), it can be substituted more easily into mathematical operations that involve angular
velocity.
Yet another definition of angular velocity can be obtained from Euler’s theorem and
from the relationship between initial and rotated vectors, as given in Equation (2.84). Let
us rewrite that equation by replacing qi by q and by replacing qf by q0 , which gives

q0 = cos Φq + (1 − cos Φ) (q · n) n + sin Φn × q (2.111)

A special application of Equation (2.111) is for small rotations over small time intervals.
Indeed, as Φ approaches zero,

Φ → dΦ cos Φ → 1 sin Φ → dΦ q0 − q → dq (2.112)

which, when substituted into Equation (2.111), yields

dq = dΦ (n × q) (2.113)

Because the infinitesimal rotation is taken about a single axis at that particular instant,
division of dΦ by dt yields the angular velocity expression in the form

= ω ω = ωn (2.114)
dt
so that dq/dt = ω × q.

2.8.4 Additive Properties of Angular Velocity


Now that we have defined angular velocity as a vector, we can use the additive properties
of vectors and obtain the angular velocity of a reference frame by adding up the angular
velocities associated with the rotations that lead to that reference frame. As an illustration,
consider an initial frame XY Z and rotate it by an angle θ1 about the X axis to obtain
an X 0 Y 0 Z 0 frame. The angular velocity of the X 0 Y 0 Z 0 frame with respect to the initial
frame XY Z is recognized as simple angular velocity. Denote this angular velocity by ω 1
and express it as

ω 1 = θ̇1 I = θ̇1 I0 (2.115)

Next, rotate the X 0 Y 0 Z 0 frame about the Z 0 axis by an angle θ2 to obtain the xyz frame.
Kinematics Fundamentals 63

The angular velocity of the xyz frame with respect to the intermediate X 0 Y 0 Z 0 frame is
also simple angular velocity and it can be expressed as

ω 2 = θ̇2 K0 = θ̇2 k (2.116)

The angular velocity of the xyz frame with respect to the XY Z frame can then be expressed
as the sum of the two angular velocities

ω = ω 1 + ω 2 = θ̇1 I + θ̇2 K0 (2.117)

and we can use coordinate transformations to express the angular velocity in terms of the
unit vectors of XY Z, X 0 Y 0 Z 0 , or xyz. In general, it is more convenient to express the
angular velocity of a reference frame in terms of the unit vectors of that frame. For the
example above,

ω = θ̇1 I0 + θ̇2 k (2.118)

Since the transformation from X 0 Y 0 Z 0 to xyz is a 3-rotation, and from Equation (2.74) we
write I0 = cos θ2 i − sin θ2 j, thus the expression for angular velocity becomes

ω = θ̇1 (cos θ2 i − sin θ2 j) + θ̇2 k = θ̇1 cos θ2 i − θ̇1 sin θ2 j + θ̇2 k (2.119)

It is clear that ω cannot be expressed as the derivative of another vector, even though
the two components of the angular velocity, ω 1 and ω 2 , are differentiable when evaluated
individually. The situation does not change if ω is written in terms of the fixed reference
frame XY Z.
When using the A and B frame notations, if there is an additional reference frame, say,
D, we begin with the A frame, rotate it to obtain the D frame, and then rotate the D frame
to obtain the B frame. The angular velocities are related by
A
ωB = A
ωD + D
ωB (2.120)

As discussed earlier, we usually attach the moving reference frame to the body. The
angular velocity of the reference frame and the angular velocity of the body are then the
same. There are cases when it is preferable not to attach the moving reference frame to the
body. An important application is rotating axisymmetric bodies. Axisymmetric bodies are
analyzed at length in Chapters 9 and 11. A simple illustration is presented here.
Consider Figure 2.28a and the elbow-shaped pendulum that swings in the xy (or XY )
plane. The z and Z axes are along the same direction and the relation between the xyz
and XY Z coordinates is shown in Figure 2.28b. The disk at the end of the pendulum is
rotating with angular velocity Ω with respect to the pendulum. We attach the xyz frame
to the elbow and write the angular velocities of the disk and the reference frame as

ω elbow = ω frame = θ̇K = θ̇k ω disk = ω frame + ω disk/frame = θ̇k + Ωi (2.121)

The angular velocity of the disk is expressed in terms of coordinates of a reference


frame not attached to the disk. It is important to make the distinction between the angular
velocity of the reference frame and that of the body. From now on, the angular velocity of
a reference frame will be denoted by ω f .
We have defined angular velocity in a number of ways, discussed what it is physically,
and derived expressions for derivatives of vectors in moving reference frames. What we have
not done is to come up with a general way to quantify angular velocity as a function of
rotational parameters. This issue will be addressed in Chapter 9.
64 Applied Dynamics

 


 




 



  



FIGURE 2.28
a) Spinning disk on a rotating elbow arm, b) coordinate axes xyz and XY Z.

Example 2.10
Obtain the angular velocities of the double link in Figure 2.29 that is supported by a rotating
column.
The XY Z frame rotates with the column and the angular velocity of the column is ΩK.
Attaching an xyz frame to the first link, and noting that the link makes an angle of θ1 with
the vertical, the angular velocity of the first link becomes
ω 1 = ΩK + θ̇1 I = ΩK + θ̇1 i [a]
where K = sin θ1 j + cos θ1 k, so that the angular velocity of the first link becomes
ω 1 = θ̇1 i + Ω sin θ1 j + Ω cos θ1 k [b]
To obtain the angular velocity of the second link, observe that the angle θ2 is also
measured from the vertical, so that its value is independent of the orientation of link 1. The
angular velocity of the second link is found similarly, by replacing the number 1 with 2 in
Equation [b], as well as the unit vectors with their primed counterparts, with the result
ω 2 = ΩK + θ̇2 I = θ̇2 i0 + Ω sin θ2 j0 + Ω cos θ2 k0 [c]

2.9 Angular Acceleration and Second Derivatives


This section extends the developments of the previous section and obtains expressions for
angular acceleration and for other second derivatives.

2.9.1 Angular Acceleration


An important application of the transport theorem is the calculation of the derivative of
angular velocity, known as angular acceleration. Denoted by α , the angular acceleration is
defined as
d
α = ω (2.122)
dt
Kinematics Fundamentals 65

 





 
 
 
 


 

  




FIGURE 2.29
a) Double link on a rotating column, b) reference frames for first link.

Note that the time derivative here is being taken in the inertial (nonmoving) reference
frame. We need to make the distinction between the reference frame (or body) whose angular
velocity is considered and the coordinate axes used to express the angular velocity. Two
scenarios are possible:
1. The angular velocity components of the moving frame (or body) are expressed in terms
of the coordinates of the moving frame. This is the most widely encountered case. The
moving frame is attached to the body, as shown in Figure 2.30.








FIGURE 2.30
Reference frame attached to body.

Let us consider the moving xyz frame and express the angular velocity and angular
acceleration of the xyz frame as
ω = ωx i + ωy j + ωz k α = αx i + αy j + αz k (2.123)
Differentiation of the angular velocity results in
α = ω̇x i + ω̇y j + ω̇z k + ω × ω = ω̇x i + ω̇y j + ω̇z k (2.124)
66 Applied Dynamics

so that the components of the angular acceleration are time derivatives of the compo-
nents of the angular acceleration

αx = ω̇x αy = ω̇y αz = ω̇z (2.125)

Interestingly, we get the same result when the angular velocity is expressed in terms of
the coordinates of the inertial frame. Expressing ω as ω = ωX I + ωY J + ωZ K, it follows
that αX = ω̇X , αY = ω̇Y , αZ = ω̇Z .
2. The angular velocity components of the moving frame are not expressed in terms of the
coordinates of the moving frame. An example is shown in Figure 2.28. This approach is
used primarily when dealing with rotating axisymmetric bodies and in other cases when
it is convenient to do so.
Here, we need to distinguish between the angular velocities of two reference frames.
The notation ω f denotes the angular velocity of the reference frame used to express
the vector that is differentiated. The transport theorem for the angular acceleration
becomes

α = ω̇ ω rel + ω f × ω
ω = ω̇ (2.126)

Note that in this case αi 6= ω̇i (i = x, y, z). As an illustration, consider the elbow-shaped
rod and disk in Figure 2.28. The xyz frame is attached to the elbow-shaped rod, whose
angular velocity is ω elbow = θ̇k = ωk. The disk rotates about the x axis and the angular
velocity of the disk with respect to the collar is ω disk/elbow = Ωi. The angular velocity
of the disk can then be written as

ω disk = ω elbow + ω disk/elbow = ωk + Ωi (2.127)

To obtain the angular acceleration, observe that the angular velocity of the reference
frame whose coordinates are used to express ω disk is the xyz frame, so that ω f =
ω elbow = ωk and that this reference frame is not attached to the disk. Differentiating
the angular velocity term gives the angular acceleration as

ω disk = ω̇
α disk = ω̇ ω elbow + ω̇
ω disk/elbow (2.128)

Evaluation of the first term results in

ω̇ ω elbowrel + ω f × ω elbow = ω̇k + ωk × ωk = ω̇k


ω elbow = ω̇ (2.129)

The cross product term vanishes because ω f = ω elbow . The second term becomes

ω disk/elbowrel + ω f × ω disk/elbow
ω disk/elbow = ω̇
ω̇

= Ω̇i + ωk × (ωk + Ωi) = Ω̇i + ωΩj (2.130)

so that the angular acceleration of the disk is

α disk = Ω̇i + ωΩj + ω̇k (2.131)

For the general case involving more than one reference frame, say, A, C, F, B:
A
ωB = A
ω C + Cω F + F ω B (2.132)
Kinematics Fundamentals 67

the angular acceleration Aα B is obtained by differentiating each term individually, as shown


below.

A dA C A C A dC F C
ω = α ω = α F + Aω C × C ω F
dt dt

A dF B F
ω = α B + Aω F × F ω B (2.133)
dt
The above results are combined into
A B A C
α = α + C α F + F α B + Aω C × C ω F + Aω F × F ω B (2.134)

The transport theorem is the preferred approach for obtaining derivatives of angular
velocity, especially for complex problems of angular acceleration, and it is more adaptable
to implementation by digital computers.

2.9.2 Second Derivatives


Many times, it is necessary to take a second derivative or to take the derivative of an
expression in one reference frame that has been obtained by differentiation in another frame.
The transport theorem is suitable to use in such cases. The following notation will be used
when dealing with a second derivative:
   
A A d A d B B d B d
q̈ = q = q̈ q̈ = q = q̈rel (2.135)
dt dt dt dt

For a mixed derivative, application of the transport theorem gives


     
A d B d d Bd d
q = B q + Aω B × B q = q̈rel + ω × q̇rel (2.136)
dt dt dt dt dt

Note that when more than one derivative is taken in different reference frames, changing
the order of differentiation gives different results. This applies to differentiation with respect
to time, as well as to differentiation with respect to other variables. Be aware that a vector
may be a function of a certain variable in one reference frame and not in another.
Let us next evaluate the second derivative of a vector q. Equation (2.109) gives the first
derivative. Differentiating it gives
 
A A d B d d A B
q̈ = q +A ω × q + Aω B × A q̇ (2.137)
dt dt dt

Introducing Equation (2.136) to the above equation and collecting terms give
A
q̈ = Aα B × q + Aω B × Aω B × q + 2Aω B × B q̇ + B q̈

(2.138)

or

q̈ = α × q + ω × (ω
ω × q) + 2ω
ω × q̇rel + q̈rel (2.139)

A significant application of the above two equations will be discussed in the next section.
We end this section with an important reminder. It is crucial that we be able to distin-
guish between the reference frame in which a derivative is taken and the reference frame
whose coordinates are used to express a vector. Usually the two are the same; however,
68 Applied Dynamics

there are cases when they are not, especially when dealing with axisymmetric bodies. When
in doubt, we can use a straightforward, but not as intuitive, way of calculating a derivative
by differentiating each term individually. The derivative of q = qx i + qy j + qz k is

dq
q̇ = = q̇x i + q̇y j + q̇z k + qxω × i + qy ω × j + qz ω × k (2.140)
dt
in which ω is the angular velocity of the xyz coordinates, that is, the coordinate system in
which the vector is expressed. The same concept also applies to second derivatives.

Example 2.11

 
  



 

  
 
 
 
 
 

 

FIGURE 2.31
a) Two rotations of a coordinate system XY Z, b) the second rotation viewed from a different
vantage point.

Consider the coordinate transformation of a system shown in Figure 2.31a. Begin with
an inertial frame XY Z, rotate it by θ1 about X to obtain the frame x0 y 0 z 0 and rotate this
frame by θ2 about z 0 to obtain the xyz frame. The angular velocities associated with this
rotation sequence are
ω 1 = θ̇1 I ω 2 = θ̇2 k0 [a]
and the angular velocity of the xyz frame is ω = ω 1 + ω 2 .
Let us first find the angular acceleration by using the multiple frame approach:

α = ω̇ ω 2rel + ω f × ω 2
ω 1 + ω̇ [b]

in which ω f = ω 1 = θ̇1 I denotes the angular velocity with which the x0 y 0 z 0 frame is rotating.
Evaluating each term, and using the xyz coordinates to express α , and noting from Figure
2.31b that I = i0 = cos θ2 i − sin θ2 j and k0 = k, results in

ω 1 = θ̈1 I = θ̈1 (cos θ2 i − sin θ2 j)


ω̇ ω 2rel = θ̈2 k
ω̇

ω f × ω 2 = θ̇1 (cos θ2 i − sin θ2 j) × θ̇2 k = θ̇1 θ̇2 (− sin θ2 i − cos θ2 j) [c]


Introduction of Equation [c] into Equation [b] leads to
   
α = θ̈1 cos θ2 − θ̇1 θ̇2 sin θ2 i − θ̈1 sin θ2 + θ̇1 θ̇2 cos θ2 j + θ̈2 k [d]
Kinematics Fundamentals 69

Alternatively, we can obtain the angular acceleration by writing the angular velocity in
terms of the xyz coordinates and differentiate. Express ω as

ω = θ̇1 (cos θ2 i − sin θ2 j) + θ̇2 k [e]

and the individual components of the angular velocity are

ωx = θ̇1 cos θ2 ωy = −θ̇1 sin θ2 ωz = θ̇2 [f ]

The angular acceleration components are obtained by direct differentiation of the angular
velocity components

αx = θ̈1 cos θ2 − θ̇1 θ̇2 sin θ2 αy = −θ̈1 sin θ2 − θ̇1 θ̇2 cos θ2 αz = θ̈2 [g]

which, of course, is the same result as Equation [d].


For comparison purposes, let us obtain the angular acceleration in terms of the x0 y 0 z 0
coordinates. The angular velocities are ω 1 = θ̇1 i0 , ω 2 = θ̇2 k0 , ω = θ̇1 i0 + θ̇2 k0 , ω f = θ̇1 i0 .
Using Equation [b], the angular acceleration terms are obtained as

ω 1 = θ̈1 i0
ω̇ ω 2rel = θ̈2 k0
ω̇ ω f × ω 2 = θ̇1 i0 × θ̇2 k0 = −θ̇1 θ̇2 j0 [h]

We can also use the complete expression for ω and differentiate it to find the angular
acceleration. Recall that the angular velocity of the xyz frame is now being expressed in
terms of the x0 y 0 z 0 coordinates, so that
 
α = ω̇ ω rel + ω f × ω = θ̈1 i0 + θ̈2 k0 + θ̇1 i0 × θ̇1 i0 + θ̇2 k0 = θ̈1 i0 − θ̇1 θ̇2 j0 + θ̈2 k0 [i]

which is the same result as Equation [h].

2.10 Relative Motion


This section develops expressions for relative velocity and relative acceleration. The trans-
port theorem is applied to relate the motions of two points in moving reference frames.
Consider a rigid body that is undergoing translation and rotation. We can analyze the
motion by attaching a moving reference frame xyz to the body. The angular velocity and
angular acceleration of the frame then become the angular velocity and acceleration of the
body. Two points, B and P , lie on the body (Figure 2.32). Point B is fixed on the body and
point P may or may not move with respect to the body (hence, relative to the reference
frame). Also consider a nonmoving (inertial) coordinate system XY Z.
The angular velocity and angular acceleration of the body are denoted by ω and α,
respectively. The positions of points B and P are related by

rP = rB + rP/B (2.141)

where rB and rP are measured from a fixed reference point O and rP/B is the relative
position vector. This vector is similar to the vector q used in Equation (2.97) and in Equation
(2.139). Although we can substitute these equations directly into the relative velocity and
relative acceleration expressions, it is instructive to repeat the derivation in terms of rP/B .
Differentiation of Equation (2.141) leads to the relative velocity equation

vP = vB + vP/B (2.142)
70 Applied Dynamics

z y
P
rP/B
Z B
rP x

rB
O Y

FIGURE 2.32
Two points on a body.

The time derivative of the position vector rB , vB , is obtained by direct differentiation of


rB (or by measurement from the fixed reference point). The time derivative of the relative
position term rP/B can be obtained by using the transport theorem, as the vector rP/B
moves with respect to the reference frame. Applying Equation (2.97) yields

vP/B = ṙP/B = v(P/B)rel + ω × rP/B (2.143)

The first term on the right, v(P/B)rel , is the velocity viewed by an observer sitting (or a
measuring device located) on the body. This term vanishes if point P is fixed on the body.
The second term, ω × rP/B , is the velocity of point P with respect to point B due to the
rotation of the reference frame. The relative velocity expression thus becomes

vP = vB + ω × rP/B + v(P/B)rel (2.144)

The relative acceleration equation is obtained by differentiating Equation (2.144). Dif-


ferentiating the first term on the right side gives aB , as vB is measured from a fixed point
and so is its time derivative. Differentiation of the second term by means of the transport
theorem results in
d  
ω × rP/B = α × rP/B + ω × v(P/B)rel + ω × rP/B (2.145)
dt
and differentiation of the third term gives
d 
v(P/B)rel = a(P/B)rel + ω × v(P/B)rel (2.146)
dt
Combining the above equations, we get

aP = aB + α × rP/B + ω × ω × rP/B + 2ω
ω × v(P/B)rel + a(P/B)rel (2.147)

The term α × rP/B is due to the angular acceleration of the rotating frame, while
ω × ω × rP/B is the centripetal acceleration. For the special case of plane motion, the
centripetal acceleration takes the form

ω × ω × rP/B = −ω 2 rP/B

(2.148)

The term 2ωω × v(P/B)rel is the Coriolis acceleration.


If point P does not move with respect to the body (that is, with respect to the moving
Kinematics Fundamentals 71

frame), then the velocity and acceleration of P observed from the body are zero, v(P/B)rel =
0, a(P/B)rel = 0, and the relative acceleration expression reduces to

aP = aB + α × rP/B + ω × ω × rP/B (2.149)

When solving relative motion problems on a plane, we can use a scalar approach to
add the vectors involved by means of velocity polygons and we express the relative velocity
relationships as a sketch.

Example 2.12

a) b)

h3
0.6 m
g O
r
O h2 b3
B P B
h1
Extending
arm P b2
A

FIGURE 2.33
a) Robotic arm on a rotating shaft, b) coordinate frames.

Consider the robotic arm in Figure 2.33a mounted on a rotating shaft. The shaft is
rotating with a constant angular velocity of Ω = 0.1 rad/s. The angle φ that the robotic
arm makes with the shaft is varying according to the relationship φ = πt/12 rad, with
φ (t = 0) = 0. With a motion similar to that of an automobile antenna, a second arm
extends from the outer end of the arm according to the relationship r (t) = 5t2 cm.
Find the angular velocity and angular acceleration of the rod, as well as the acceleration
of the tip of the extending arm, at time t = 2 seconds.
This problem can be solved using a variety of approaches. In one approach, two relative
frames are used, one attached to the shaft (H frame, with axes h1 , h2 , h3 ) and rotating with
Ω, the other (B frame) attached to the arm and rotating with φ̇ with respect to the shaft,
as shown in Figure 2.33b. In the second approach, we deal with a single frame attached to
the robotic arm. For this problem, it is more convenient to use the single frame approach.
The following information is available:
π
Ω = 0.1 rad/s Ω̇ = 0 φ (2) = π/6 rad φ̇ (2) = rad/s φ̈ = 0
12
2
r (2) = 20 cm ṙ (2) = 20 cm/s r̈ (2) = 10 cm/s [a]
As the geometry and positions of the linkages are known, the velocity analysis can be
conducted. The pin joint to which the arm is connected is denoted by O. The angular
velocities of the reference frames are
A H H B π
ω = Ωh3 = 0.1h3 rad/s ω = φ̇h1 = h1 rad/s [b]
12
72 Applied Dynamics

where h3 = − cos φb2 + sin φb3 = − 3/2b2 − 0.5b3 . The angular velocity of the arm is
obtained by adding the individual angular velocities and has the form
A π π √
ωB = A
ω H + H ω B = 0.1h3 + b1 = b1 − 0.05 3b2 + 0.05b3 rad/s [c]
12 12
The relative velocity expression between point O and P is
A A
vP = vO + Aω B × rP/O + B vP [d]

in which A vO = 0 and
B

rP/O = OB + r (2) b2 = 80b2 cm vP = ṙ (2) b2 = 20b2 cm/s [e]

Carrying out the cross product in Equation [d] gives

A B
π √  π
ω × rP/O = b1 − 0.05 3b2 + 0.05b3 × 80b2 = −4b1 + 20 b3 cm/s [f ]
12 3
Adding Equations [d] and [e] gives the velocity of P as

A π
vP = −4b1 + 20b2 + 20 b3 cm/s [g]
3
Next, consider the acceleration and write
A A
aP = aO +A α B × rP/O +A ω B ×A ω B × rP/O + 2Aω B ×B vP/O +B aP/O [h]

where A aO = 0 and B aP/O = 10b2 . The angular acceleration is obtained by applying the
transport theorem to the angular velocity
A B A H H π
α = α + αB + A
ω H ×H ω B = 0 + φ̈h1 + 0.1h3 × b1
12
 √  π π √ π 2
= −0.05 3b2 + 0.05b3 × b1 = b2 + 3 b3 rad/s [i]
12 240 240
The individual terms are evaluated next and they result in

A B
 π √ π  √ π 2
α × rP/O = b2 + 3 b3 × 80b2 = − 3 b1 cm/s
240 240 80

5π 2 √
 
A π
ω B ×A ω B × rP/O = − √ b1 + 0.2 − b2 − 0.2 3b3
3 9
π √  10π 2
2Aω B ×B vP/O = 2 b1 − 0.05 3b2 + 0.05b3 × 20b2 = −2b1 − b3 cm/s [j]
12 3
Adding the respective terms, the acceleration is obtained as
√ π 5π 2 √
 
A π 10π 2
aP = − 3 b1 − √ b1 + 0.2 − b2 − 0.2 3b3 − 2b1 − b3 + 10b2 cm/s [k]
80 3 9 3

which can be further simplified.


Kinematics Fundamentals 73

2.11 Instantaneous Center of Zero Velocity


An important property of rigid bodies undergoing plane motion is that of an instant center.
The instant center is a powerful visual tool that gives valuable insight regarding the nature
of the motion. It can be used as an analysis as well as a design tool.
At any instant of motion, there exists an axis perpendicular to the plane of motion,
called the instantaneous axis of zero velocity, such that the body can be viewed as rotating
about that axis at that instant. The intersection of this axis and the plane of motion is
called the instantaneous center of zero velocity, or instant center.
In general, the instant center of a body is located by visual inspection. To establish the
location of the instant center, we need to know the velocities of two points on the body. If
the velocities are not in the same direction, we draw two lines, beginning at the points at
which the velocities are known and perpendicular to the velocities. Their intersection is the
instant center. This is the most commonly encountered case.
If the velocities of the two points are in the same direction, we again draw two lines:
one joining the points at which the velocities are known and the other joining the tips of
the velocity vectors drawn to scale. Their intersection is the instant center. Figure 2.34
illustrates these common ways of locating an instant center.

IC
A vA
B vB

B vB
A

vA
IC

FIGURE 2.34
Locating instant centers.

While the instant center has zero velocity, its location at every instant is different, and
its acceleration is not zero. Hence, the value of instant center analysis diminishes for kinetic
analysis.3
The definition of instant center here is for a single body. Section 3.9 extends the concept
of instant centers to linkages and mechanisms. This expanded definition is widely used for
mechanism and vehicle suspension system analysis.

Example 2.13
A rod of length L is sliding against two surfaces that are not perpendicular to each other,
as shown in Figure 2.35a. At the instant shown, the rod makes an angle of 45◦ with the
3 In Chapter 15, while studying vehicle suspension systems, sometimes one can make the assumption that
the instant center remains stationary for a short time period.
74 Applied Dynamics

 



 
 




    


  

FIGURE 2.35
a) Sliding rod, b) geometry and instant center.

horizontal. Also, point B is moving to the right with speed vB . Find the instantaneous
center of zero velocity and use it to calculate the velocity of point A.
Since points A and B are sliding on the surfaces with which they are in contact, the
instant center can be found by drawing perpendicular lines to the surfaces, as shown in
Figure 2.35b. The coordinates of the instant center can be found by geometry. The following
relationships hold:
√ √
◦ 2 ◦ 3
BD = AD = L cos 45 = L AD = AC cos 30 = AC [a]
2 2
from which we conclude that
r r
2 ◦ 1
AC = L CD = AC sin 30 = L [b]
3 6
so that
BC = BD + CD = 1.1154L [c]
The angular velocity of the rod is found from
vB vB
ω = = [d]
BC 1.1154L
and the velocity of A becomes
r
vB 2
vA = ωAC = L = 0.7320vB & [e]
1.1154L 3

2.12 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Beer, F.P., Johnston, E.R., Cornwell, P., Vector Mechanics for Engineers: Dynamics, 10th
Edition, McGraw-Hill, 2012.
Kinematics Fundamentals 75

Das, B.M., Kassimali, A., and Sami, S., Mechanics for Engineers: Dynamics, J. Ross, 2010.
Meriam, J.L., and Kraige, L.G., Engineering Mechanics: Dynamics, 7th Edition, Wiley,
2012.

2.13 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 2.3—Reference Frames: Single Rotation in a Plane


2.1 (E) The coordinates of a point viewed in the XY plane are r = 3I − 6J. By what angle
θ does this coordinate system needs to be rotated (about the Z axis) into an xy coordinate
system, so that the vector r will have the form r = 3ai + aj?
2.2 (M) The vectors r = 3i + 4j and q = 6I − 2J are defined in the xy and XY coordinate
systems, respectively. If xy is obtained by rotating XY by an angle of 45◦ clockwise, find
the angle between r and q.

Section 2.4—Column Vector Representation


2.3 (M) Given the vectors r = 3i + 4j − 5k and s = −2J − K, where the xyz coordinate
frame is obtained by rotating XY Z by 30◦ about the Z axis, obtain the cross product r × s
using column vector notation and express the final result using the xyz coordinates.
2.4 (M) Given the function S = 3q12 + 5q22 − 8q32 + 4q1 q2 − 12q1 q3 , calculate dS/d{q} and
the [D] matrix.
2.5 (M) We know that a vector crossed into itself gives the zero vector, r × r = 0. Yet
when considering the column vector representation of r, which is [r̃], and taking the matrix
product [r̃][r̃], the result is not zero. Why?

Section 2.5—Commonly Used Coordinate Systems


2.6 (E) An object moving along a straight line has an acceleration of a = v̇ = ẍ = 1 + 3x,
where x denotes position. If motion begins with initial conditions x0 = 2.5, v0 = 0.3,
calculate the velocity when xf = 5.

FIGURE 2.36
Figures for a) Problem 2.7, b) Problem 2.8.
76 Applied Dynamics

2.7 (M) The projectile in Figure 2.36a is fired from a height of 100 m with a speed of v0 = 40
m/s and making an angle of θ = 30◦ with the horizontal. Calculate the amount of time the
projectile stays in the air and the horizontal distance travelled by the projectile.
2.8 (C) A racquetball player hits the ball from a distance of 10 m from the front wall and
from a height of 1 m from the floor. The goal of the player is to select the velocity v0 and
launch angle θ such that the ball will hit the wall at a height less than 30 cm and speed less
than 15 m/s. Create a plot of v0 vs. θ and identify the ranges of these parameters so that
the ball will hit the wall at the desired height. Neglect air resistance.
2.9 (E) A projectile is launched with speed v0 = 10 m/s and angle θ = 30◦ . Calculate the
radius of curvature of the path when the projectile reaches its highest position.
2.10 (E) A vehicle is entering a turn with radius of curvature 120 m. Its speed is 80 km/h.
If the maximum acceleration that the wheels can sustain is 0.75g, how much (in g) can the
vehicle accelerate during the turn?
2.11 (M) The velocity and acceleration of a particle in the xyz coordinate system are
v = 3i + 4j − 6k m/s and a = −2i + 3k m/s2 . Using normal-tangential coordinates, calculate
the change in speed v̇ and radius of curvature ρ. Also, calculate the unit vector u that is
orthogonal to the plane of motion.

  
 

 






FIGURE 2.37
Figures for a) Problems 2.12 and 2.17, b) Problem 2.13.

2.12 (M) The airplane in Figure 2.37a is observed by means of cylindrical coordinates. The
plane is flying with constant velocity of v = 350I + 25K ft/sec. At the instant considered,
the plane is at an altitude of 3,000 ft and horizontal distance of 6,000 ft from the radar.
Calculate r, ṙ, r̈, θ̇, θ̈ at this instant. Note that at this instant the Y axis is the radial
direction r.
2.13 (M) The vehicle in Figure 2.37b is moving up with constant speed up a spiraled road
of constant radius R. In five full turns, the vehicle reaches height h. Express the position,
velocity, and acceleration of the vehicle in terms of polar coordinates and obtain the rela-
tionships between the unit vectors in polar coordinates and normal-tangential coordinates.
What is the radius of curvature of the path the vehicle follows?
2.14 (M) Pin A in Figure 2.38 is moving up with constant speed of 2 m/s. Calculate the
angular velocity and angular acceleration of the rod OB when θ = 60◦ .
2.15 (E) The position of a vehicle is described by the vector r = 3I − 4J + 8K. Express r
in terms of spherical coordinates.
2.16 (M) Consider Figure 2.15. Point P is moving according to the relationship θ (t) =
πt/100 rad, point P 0 on the x axis is 55 m away from O and the distance from P 0 to
P varies according to the relationship z = 12 − 6 sin(πt/12) m. Express r and its time
derivative, v in the spherical coordinate system.
Kinematics Fundamentals 77

 




FIGURE 2.38
Figure Problem 2.14.

2.17 (M) The airplane in Figure 2.37a is tracked by radar by means of spherical coordinates.
The plane is flying with constant velocity of v = 350I+25K ft/sec. At the instant considered,
the plane is at an altitude of 3,000 ft and horizontal distance of 6,000 ft from the radar.
Calculate the polar and azimuthal angles and their derivatives at this instant.

Sections 2.6 and 2.7—Moving Reference Frames and Selection of Rotation Pa-
rameters
2.18 (M) A coordinate system XY Z is first rotated about the Y axis by −30◦ , resulting
in the X 0 Y 0 Z 0 axes. Then, X 0 Y 0 Z 0 is rotated about the Z 0 axis by 67◦ to yield the xyz
coordinates. Find the coordinates of the vector r = −6I+3K in terms of the xyz coordinates.
2.19 (M) A coordinate system xyz is obtained by taking an XY Z coordinate system and
first rotating it about the Y axis by an angle of 15◦ counterclockwise to get the X 0 Y 0 Z 0
coordinates and then by rotating the X 0 Y 0 Z 0 coordinate system by an angle of 30◦ about the
X 0 axis to get the xyz coordinates. Find the resulting rotation matrix and the coordinates
in the XY Z system of a point whose position in the xyz coordinates is r = 2i − 3k.
2.20 (C) A coordinate system xyz is obtained by taking an XY Z coordinate system, and by
first rotating it about the Z axis by an angle of φ to get the X 0 Y 0 Z 0 coordinates. Then, the
X 0 Y 0 Z 0 coordinates system is rotated by an angle of θ about the X 0 axis to get the X 00 Y 00 Z 00
coordinates. The final rotation is about the Z 00 axis by ψ. Find the resulting rotation matrix
[R]. Evaluate [R] for φ = 30◦ , θ = 45◦ and ψ = −60◦ . Then find the coordinates in the
XY Z system of a point whose position in the xyz coordinates is r = −5i + 3j.
2.21 (D) Consider the box in Figure 2.39a. Denote the initial frame by XY Z. First rotate
the box about OA (X axis) by 30◦ . The resulting X 0 Y 0 Z 0 axes are then rotated by −10◦
about the Y 0 axis, yielding the X 00 Y 00 Z 00 axes. Finally, the X 00 Y 00 Z 00 axes are rotated by
18◦ about the Z 00 axis to yield the xyz coordinates. Calculate the rotation matrix that
relates the xyz to XY Z and the coordinates of point D in the initial XY Z frame after
these rotations.
2.22 (M) The rectangular box in Figure 2.39a is rotated counterclockwise by an angle of
30◦ about the axis passing through points O and D. Find the coordinates of point A after
this rotation in terms of the XY Z coordinates. Solve this problem using both a body-fixed
rotation sequence, as well as Equation (2.84).
2.23 (D) The rectangular box in Figure 2.39b is rotated counterclockwise by an angle of
30◦ about the axis passing through points O and B. Find the coordinates of point A after
78 Applied Dynamics
 



   
   
 
  



FIGURE 2.39
Figures for Problems 2.21–2.23.

this rotation in terms of the XY Z coordinates. Solve this problem using both a body-fixed
rotation sequence, as well as Equation (2.84).

a) b)
Z z
0.1 m
B2
A2
B
y b
c
t/2 G x
X
x 0.7 m
x x'
1 D t/2 (Roll axis)
0.5 m B1
A1
. 2
P y z

FIGURE 2.40
Figures for a) Problem 2.24, b) Problem 2.25.

2.24 (M) Consider the double-link robot mounted on a rotating base in Figure 2.40a with
the angles θ1 and θ2 measured from the vertical. Find the position of the tip of the robot
arm in terms of inertial coordinates when φ = 30◦ , θ1 = 60◦ and θ2 = −15◦ . The XY Z
coordinates are inertial. Solve this problem by using column vector notation.
2.25 (M) Figure 2.40b is representative of the roll motion of the vehicle. The xyz axes are
attached to the vehicle. The vehicle rotates about the roll axis x0 by angle φ. using the xyz
coordinates, calculate the vertical displacements of points A1 , A2 , B1 and B2 with respect
to G after the vehicle has rotated by φ = 10◦ . The roll axis x0 makes an angle of θ = 5◦
(rotation about y axis) with the x axis.
Kinematics Fundamentals 79

Section 2.8—Rate of Change of a Vector, Angular Velocity


2.26 (M) Consider Problem 2.24. Now, you are given that the angular velocity of the base is
φ̇ = 0.2 rad/s and the angular velocities of the arms are θ̇1 = −0.3 rad/s and θ̇2 = 0.32 rad/s.
Calculate the angular velocity of each arm and express it in terms of the xyz coordinates.
.
R P
v
B
.

FIGURE 2.41
Figure for Problems 2.27, 2.30, and 2.41.

2.27 (E) A disk of radius R shown in Figure 2.41 spins at the rate of θ̇ about an axle held
by a fork-ended horizontal rod that rotates itself at the rate of φ̇. Find the angular velocity
of the disk.

!" +" ,"


z z'
Z
-./01,!2 # z'

O
%&'
( O
 x' B

%&' (
X)*x'  x (
%&' $ C 
B
B
y' y) y'
$
C C
x' $

FIGURE 2.42
Figure for Problem 2.28.

2.28 (D) The robot arm in Figure 2.42a makes an angle of 40◦ with the rotating shaft. The
arm rotates about the y 0 axis with the relationship θ (t) = 20
π
cos 2t rad, as shown in Figures
2.42b, and c. The shaft has an angular velocity of ω1 = 0.5 rad/s. At the tip of the arm B,
there is another rod. A disk spins counterclockwise with ω3 = 7 rad/s about this rod. Find
the total angular velocity of the disk at t = 3 s.

Section 2.9—Angular Acceleration and Second Derivatives


2.29 (M) Consider Problem 2.24. Now, you are given that the angular velocity of the base
is φ̇ = 0.2 rad/s and φ̈ = −0.12 rad/s2 . The angular velocities of the arms are θ̇1 = −0.3
rad/s and θ̇2 = 0.32 rad/s, both angular velocities being constant. Calculate the angular
acceleration of each arm.
80 Applied Dynamics

2.30 (M) A disk of radius R shown in Figure 2.41 spins at the rate of θ̇ about an axle held by
a fork-ended horizontal rod that rotates itself at the rate of φ̇. Find the angular acceleration
of the disk.

a) b)

x
x
G P
Y
Pitch
y
y
z z Z

FIGURE 2.43
Figure for Problem 2.31. a) Airplane, b) rear view.

2.31 (M) The airplane in Figure 2.43a is moving with speed of 600 kph in a curved trajectory
ρ = 9000 m. Figure 2.43b shows the rear view of the airplane. At the same time, the airplane
is pitching upwards at the constant rate of ωy = 0.1 rad/s. The propeller is spinning
with the constant counterclockwise angular velocity of ωx = −4000 rpm. Find the angular
acceleration of the propeller.
2.32 (D) Consider Problem 2.28 and calculate the angular acceleration of the disk at t = 3.6
sec, given that ω1 and ω3 are both constant.
2.33 (E) The American football in Figure 2.44a is spinning about its axis of symmetry (z)
at the constant rate of 88 rpm. The axis of symmetry rotates about a horizontal axis (y)
at a rate of 10 rpm which decreases at the rate of −1.2 rpm/sec. Calculate the angular
acceleration in vector form and the magnitude of the angular acceleration.

a) b)
x z 25 cm
B
z
C

30 cm
2
y
O 1
y A

FIGURE 2.44
Figures for a) Problem 2.33, b) Problems 2.34, 2.42, and 3.18.
Kinematics Fundamentals 81

Section 2.10—Relative Motion


2.34 (M) When θ1 = θ2 = 60◦ in Figure 2.44b, then length of OC = 25 cm and the angle
between AB and BC becomes 36.87◦ . For this configuration, given that ωAB = 4 rad/sec
ccw and the velocity of the slider is 9 in/sec upwards, calculate the angular velocities ωOC
and ωBC .
2.35 (E) For the problem above, calculate the angular accelerations given that ωAB = 0 and
the velocity of the slider is constant.

0.8 m
a) b)
g
 L
R
h3

P
b2
h1, b1

h2

FIGURE 2.45
Figures for a) Problem 2.36, b) Problem 2.38.

2.36 (M) A bead of mass m is free to slide on a hoop of radius R, as shown in Figure
2.45a. The hoop rotates with a constant angular velocity Ω about the vertical axis. Find
the acceleration of the bead using a) a reference frame H attached to the hoop, with h3 in
the vertical direction and h1 perpendicular to the hoop, and b) a reference frame B attached
to the bead, with b1 b2 b3 obtained by rotating h1 h2 h3 about the h1 axis by θ.
2.37 (M) The pendulum of fixed length L in Figure 8.37 swings on the inclined plane xy.
The incline angle γ of the plane (between X and x axes) is being raised at the constant
rate of γ̇. The pivot point of the pendulum is at a distance of 1.2L from the y axis. Find
the velocity and acceleration of the pendulum.
2.38 (E) A spring pendulum is attached to a rotating shaft by an arm of length d = 0.8 m,
as shown in Figure 2.45b. At the instant shown, the shaft is rotating with angular velocity
Ω = 0.4 rad/s, θ = 30◦ , θ̇ = 0.3 rad/s. The length of the pendulum is L = 1.3 m and it is
getting shorter at the rate of 0.1 m/s. Find the velocity of the tip of the pendulum.
2.39 (E) Consider the previous problem and find the acceleration of the tip of the pendulum
if Ω = constant, θ̈ = 0.2 rad/s, L̈ = −0.1 m/s2 .
2.40 (M) The platform in Figure 2.46 is rotating with a constant angular velocity of ω = 0.2
rad/s. Pivoting on the platform is a tube oscillating according to the relationship θ (t) =
π
6 sin 2t rad. A particle slides inside the tube. Find the velocity and acceleration of the
particle at t = π seconds, at which point it is given that y = 40 cm, ẏ = −30 cm/s, and
ÿ = −4 cm/s2 .
2.41 (M) A disk of radius R shown in Figure 2.41 spins at the constant rate of θ̇ about an
axle held by a fork-ended horizontal rod that rotates itself at the nonconstant rate of φ̇.
An ant is crawling towards the center of the disk with constant speed v with respect to the
disk. Find the acceleration of the ant as a function of θ and when the ant is at the edge of
the disk, at point P .
82 Applied Dynamics

 
  

 
 
 



FIGURE 2.46
Figure for Problem 2.40.

Section 2.11—Instantaneous Center of Zero Velocity


2.42 (M) Solve Problem 2.34 by means of instant centers.

a) b)
3 kg
0.4 m
A y
g O P
2.5L
B
2
12 kg L
0.8 m 
1m 1 A
O x
2L
B

FIGURE 2.47
Figures for a) Problem 2.43, b) Problems 2.44 and 3.21.

2.43 (M) Consider the rod in Figure 2.47a and calculate its instant center. Given that rod
OA has a clockwise angular velocity of 0.6 rad/s, calculate the angular velocity of rod AB
and the velocity of point B.
2.44 (D) Calculate the location and coordinates of the instant center of the rod BP in
Figure 2.47b when θ1 = 30◦ . The incline is at an angle of ψ = 45◦ with the horizontal.
3
Kinematics Applications

3.1 Introduction
This chapter extends the kinematics formulations of the previous chapter to applications
of interest to practicing engineers and scientists. We begin with a special application of the
relative motion equations; motion with respect to the rotating Earth. Contact between two
bodies is discussed, together with the constraints contact creates.
A most important application of contact between two bodies is rolling. A simple kine-
matic model of a ground vehicle, known as the bicycle model, is introduced. Kinematic
differential equations, which are used in analysis and in simulation, are developed. The
chapter ends with an introduction to mechanisms.

3.2 Motion with Respect to the Rotating Earth


An interesting application of moving reference frames is the description of motion with
respect to the rotating Earth. Motion over short distances or with small velocities and in-
volving short time periods can be analyzed relatively accurately without considering the
rotation of the Earth. There are certain phenomena, however, that necessitate the inclusion
of the rotation of the Earth in the formulation. Any motion that takes place over large pe-
riods of time (days, weeks, . . . ), such as hurricane formation, cannot be modeled accurately
without considering the rotation of the Earth.
Consider a point B on the surface of the Earth, as shown in Figure 3.1a, and attach
a moving frame to the Earth, using an xyz coordinate system. The z direction is vertical,
the x direction is toward the north and the y direction is toward the west. The latitude is
denoted by λ.
The Earth is rotating about its own axis with constant angular velocity Ω. It takes about
365.25 days to orbit the sun, and the Earth rotates about its own axis at the rate of one
revolution per day. Both rotations are counterclockwise, which leads to an angular velocity
of
  
2π 1
Ω = 1+ = 7.2921 × 10−5 rad/s (3.1)
24(60)(60) 365.25
so that, considering Figure 3.1b, we can describe the angular velocity of the Earth in vector
form as

ω = Ω (cos λi + sin λk) (3.2)

The angular acceleration of the Earth, which is an extremely small quantity, is ignored
so that α = 0. Also ignored are i) the rotation axis of the Earth, which is not fixed and

83
84 Applied Dynamics

!" #"
$%&'()*%+,
 x 0$%&'("
x 0$%&'(" )5%2)
z 03,&'45!+" P z 03,&'45!+"

B
P   B )246)
y)01,2'" !P7B 
O  O !P7B

-./!'%&

FIGURE 3.1
Earth coordinate system: a) general orientation, b) side view.

exhibits a small wobbling motion, and ii) the inclination angle between the equatorial plane
(the plane of the equator) and the ecliptic plane (the plane generated by the orbit of the
earth around the sun).
Using these assumptions, the acceleration of a point, say B, on the surface of the Earth
becomes
aB = ω × (ω
ω × rB ) (3.3)
Consider now an off-planet point P and denote its position by r = rP/B = xi + yj + zk.
The velocity and acceleration of point P , viewed from the surface of the Earth, are vrel =
ẋi + ẏj + żk and arel = ẍi + ÿj + z̈k, respectively. The acceleration of this point can be
written as
a = aB + ω × (ω
ω × r) + 2ω
ω × vrel + arel (3.4)
Since the magnitude of r is much smaller than the radius of the Earth, rB , the term
ω ×ω
ω ×r can be ignored, as it is much smaller than aB = ω ×ω
ω ×rB . The relative acceleration
equation thus becomes
ω × vrel + arel
a = aB + 2ω (3.5)
Let us examine the terms on the right side of the above equation. Writing the position
vector rB as rB = re k, the acceleration of B becomes
aB = ω × ω × rB = Ω2 re sin λ cos λi − cos2 λk

(3.6)
Using an average radius for the Earth as 6761 km, Ω2 re = 3.33 cm/s2 , which represents the
centrifugal acceleration of point B. In most problems involving the rotation of the Earth,
the term sin λ cos λi is ignored, because the other velocities and accelerations involved in
the x and y directions are much higher. The component of aB in the z direction is usually
treated as contributing to the acceleration of gravity.
The term 2ω ω × vrel is the Coriolis acceleration. While the magnitude of this term is
small, its direction is always perpendicular to the velocity viewed from Earth, vrel . Hence,
this acceleration has the effect of changing direction. The dynamics associated with motion
with respect to the rotating Earth will be analyzed in Chapter 5.
Kinematics Applications 85

Example 3.1
Consider a baseball game in Yankee Stadium (latitude 40.827◦ ) where the batter hits the
ball with a speed of 120 mph at an angle of 38◦ with the horizontal, and towards the east
(ball travels directly over third base, straight down the middle of the ball park). Calculate
the Coriolis acceleration of the ball and the horizontal deflection of the ball due to the
Coriolis acceleration.
Using the Earth-based coordinates in Figure 3.1a, recalling that 60 mph is 88 ft/sec, the
velocity of the ball, immediately after it is hit by the batter, is

vrel = 176 (− cos 38◦ j + sin 38◦ k) = −138.7j + 108.4k ft/sec [a]

The angular velocity vector is

ω = Ω (cos 40.827◦ i + sin 40.827◦ k) = 7.2921 × 10−5 (0.7566i + 0.6538k) rad/sec [b]

The Coriolis acceleration is

ω × vrel = 2 × 7.2921 × 10−5 (0.7566i + 0.6538k) × (−138.7j + 108.4k)


aCor = 2ω

2
= (1.322i − 1.196j − 1.530k) × 10−2 ft/sec [c]
The Coriolis acceleration is indeed very small, despite the high speed of the ball. However,
while the ball does not have a velocity component in the x direction when it leaves the bat,
the Coriolis acceleration does have a component in the x direction, which causes the ball to
change direction.
Assume the ball spends six seconds in the air, and also assume that the Coriolis accel-
eration remains constant during that time. The deflection in the x direction can then be
approximated as
1 1.3
x ≈ −0.5ax t2 = − × × 62 = −0.23 ft [d]
2 100
While this deflection is not small, it is negligible when compared to the ∼ 440 ft that the
ball travels in the −y direction, the change in direction caused by the spin of the ball, as
well as wind. As will be shown in Chapter 5, the Coriolis deflection is opposite in direction
to the Coriolis acceleration.

3.3 Contact
Dynamical systems are composed of components that come into and out of contact with
each other. The contacting components can translate or rotate (or both) with respect to
each other. Contact may take place in many forms:
• Joints. A joint holds two points on two (or more) bodies together and permits relative
motion of the joined bodies in certain specific directions. For example, a pin joint, such
as the ones shown in the section on mechanisms, permits the two bodies it joins to have
a rotation with respect to each other about a direction perpendicular to the pin axis,
while preventing relative motion (translation or rotation) in any other direction. Such
joints and the kinematic relationships they lead to are discussed later on in this chapter.
86 Applied Dynamics

• Impact. The contacting bodies are initially separate. They come into contact at a point
(or along a line or on a plane) for a very short period of time and then separate from
each other. Examples include the bouncing of a ball and collisions of objects.
• Rolling and Sliding. Here, the bodies in contact move with respect to each other in a
way that two different points on each of the two bodies are in contact with each other
continuously. The rolling of a disk on a plane or the sliding of a body over another body
are two typical examples.

. O1

Body 2
1

.
C2
C1
Common tangent (t)

t
2
Body 1 O2 .
Common normal (n)
n

FIGURE 3.2
Contact of two bodies: common tangent and common normal.

A typical configuration of two bodies in contact with each other is shown in Figure 3.2.
The contacting points on the two bodies 1 and 2 are referred to as C1 and C2 , respectively.
Contact between two bodies is characterized by a common tangent (t) and a common normal
(n).
Consider the case when the contours of the contacting bodies at the contact point are
smooth (continuous, with continuous derivatives, hence, no corners). The common normal
connects the contact point and the centers of curvature of the contacting bodies (points O1
and O2 in Figure 3.2). Their radii of curvature are ρ1 and ρ2 . The plane tangent to each
body at the contact point is defined as the plane of contact or contact plane. The contact
plane is perpendicular to the common normal and the line of contact lies on this plane.
When a sharp-edged body comes in contact with a body that has a smooth contour, the
tangent to the smooth-contoured body defines the common tangent. The common normal
goes through the contact point and the center of curvature of the body with the smooth
contour. When two sharp edges come into contact, one has to make a reasonable assumption
regarding the common normal and common tangent.

3.4 Rolling
Rolling is a special case of contact between two bodies, where a continuous sequence of
points on one of the bodies is in continuous contact with a continuous sequence of points on
the other body. For rolling to take place, the contacting bodies must have smooth contours
Kinematics Applications 87

and the radius of curvature for both bodies must exist at each contact point on both bodies.
This section discusses rolling of a body in plane motion. Three-dimensional rolling will be
discussed in Chapter 9.

3.4.1 General Formulation

FIGURE 3.3
Different types of rolling: a) over a flat surface, b) over a curved surface.

Rolling can occur in a variety of ways; two are depicted in Figure 3.3. The most common
is for a body to roll over a fixed surface, such as the rolling of a disk over a plane. As discussed
in the previous section, the contacting bodies define the common tangent plane (or line)
and the common normal. Two bodies in contact that do not have a common normal cannot
roll over one other.
Consider Figure 3.2 and that the two bodies are rolling with respect to each other.
Because there is continuous contact, the contact points cannot move with respect to each
other in the normal direction. This constraint can be expressed as

(vC1 − vC2 ) · n = 0 (3.7)

in which n is the unit vector along the common normal. Note that −n is also a unit vector
along the common normal as, unlike the case of normal-tangential coordinates, the common
normal does not have a defined positive direction.
A special case of rolling that is of interest, especially in the case of disks and wheels,
is that of rolling without slipping. Here, the contacting points on the two bodies have the
same velocity, that is, they do not move with respect to each other. Roll without slip is
characterized by the relationship

vC1 = vC2 (3.8)

and for the case where one of the bodies is fixed, as in the rolling of a disk over a fixed
surface, the velocity of the contact point becomes zero. For roll without slip over a fixed
surface, the contact point becomes an instant center so that vC1 = 0.
Consider, for example, Figure 3.3a, which is representative of a train wheel. When the
wheel is rolling without slipping, point C is an instant center and the velocity of the center
of the wheel becomes

vG = RΩI (3.9)
88 Applied Dynamics

The bottom point of the wheel, D, does not have zero velocity. What causes roll without
slip is the friction between the contacting points. The effects of the friction force on rolling
will be discussed in Chapter 5.

Example 3.2

FIGURE 3.4
Rolling sphere.

A sphere (or disk) of radius r rolls without slip inside a circular surface with radius R,
as shown in Figure 3.4. Calculate the angular acceleration of the sphere at t = 2, given that
the angle θ varies according to the relationship θ = 0.3 sin 4t.
Instant centers relate the angular velocity of the disk to θ̇. While point C is the instant
center for the sphere, the center of the sphere G can also be viewed as rotating about O.
The velocity of G can be written for the two instant centers of rotation as

vG = rω = − (R − r) θ̇ = −1.2 (R − r) cos 4t [a]

Differentiating the above equation gives


R−r R−r
ω̇ = − θ̈ = 4.8 sin 4t [b]
r r
The value at t = 2 becomes
R−r R−r
ω̇ (2) = 4.8 sin 8 = 4.7489 [c]
r r

3.4.2 Rolling Constraints, Wheel on an Axle


Chapter 1 discussed the concept of degrees of freedom and constraints. The rolling with-
out slipping of a disk creates three constraints: contact between the rolling body and the
body over which it is rolling, prevention of relative motion along the common normal, and
prevention of spinning and sliding on the plane on which the disk rolls.
Consider the rolling disk in Figure 3.5a. The xyz coordinates move with the disk. The xz
plane denotes the plane of the disk. The disk is always upright.1 The xyz axes are obtained
1 In essence, we do not consider nutation or lean of the disk. We will discuss general three-dimensional

rolling in Chapter 9.
Kinematics Applications 89

by rotating the disk about the Z axis of the inertial XY Z coordinate system by the heading
angle ψ, as shown in Figure 3.5b. Hence, the Z and z axes coincide and they are in the
vertical direction. The contact point of the disk travels on the xy plane, which is the same
as the XY plane.

a) b)
Side view Top view
v
z X
y x X
G x

.
R x
Y
C y G
Slide slip y
Spin slip

FIGURE 3.5
Rolling of an upright disk and slip conditions: a) side view, b) top view.

The angular velocity of the disk is


ω = Ωj + ψ̇k (3.10)
where Ω is the spin rate and ψ̇ is the change in the heading angle.
When the disk slips, the contact point C will have a nonzero velocity in the XY (xy)
plane. Slip in the x direction is called spin slip and is due to excessive spin of the disk, as
when a driver floors the accelerator. In the presence of spin slip, vG · i 6= RΩ.
Slip in the y direction is called slide slip or sliding. The no slip constraints are enforced
by means of friction between the disk and the surface on which it travels.
Writing the velocity of the contact point C as vC = vCx i + vCy j, the velocity of the
center of mass becomes
vG = vC + ω × rG/C = vCx i + vCy j + ω × rG/C

 
= vCx i + vCy j + Ωj + ψ̇k × Rk = (vCx + RΩ) i + vCy j (3.11)

When there is no spin slip, vCx = 0, and when there is no slide slip, vCy = 0. The no
sliding condition and no spin slip condition, vC = 0, leads to the relation vG = RΩi. It
follows that we can write the following constraint relations for the velocity of the rotation
center G
vG · j = 0 vG · k = 0 (3.12)
The first constraint above, vG ·j = 0, is a special type of constraint, known as a nonholo-
nomic constraint. Such a constraint can be expressed in only velocity form and it cannot be
integrated to a constraint in terms of displacement variables. To understand this better, let
us express the velocity of the center of the disk in terms of the inertial coordinate system
XY . The position vector, in terms of the fixed coordinates XY Z, is rG = XG I + YG J.
Differentiating this relationship gives
vG = ẊG I + ẎG J = ẊG (cos ψi − sin ψj) + ẎG (sin ψi + cos ψj)
90 Applied Dynamics
   
= ẊG cos ψ + ẎG sin ψ i + ẎG cos ψ − ẊG sin ψ j (3.13)

Imposition of the constraint vG · j = 0 leads to the constraint equation


ẎG cos ψ − ẊG sin ψ = 0 (3.14)
The above relationship cannot be integrated to one in terms of XG , YG , and ψ. While
three variables are needed to describe the position and orientation of the disk, the derivatives
of these variables are related to each other by the constraint equation. It follows that we
can describe the velocity of the center of the disk by two independent velocity variables.
Such variables are known as generalized speeds or quasi-velocities. For the disk at hand, it
is convenient to select the speed of the wheel center, vG , and ψ̇ as the velocity variables.

z .
x R
y
t/2
F vA
d D
A

FIGURE 3.6
Wheels on an axle.

Attaching a disk (or wheels) to a rod (or axle) and imposing two no-slipping conditions
(no slide slip and no spin slip) converts the wheel on an axle assembly into a vehicle. The
invention of the wheel on an axle, estimated to have taken place about 6000 years ago,
changed world history, making it easier to transport payloads and people.
Consider now an axle with two wheels attached to it, as shown in Figure 3.6. Assume
that both wheels roll without slipping or sliding. The centers of both wheels, points F and
D, have velocities only in the x direction. Noting that the angular velocity of the axle is θ̇k,
we can use relative velocity equations to show that the velocity of any point on the axle is
in the x direction.
The velocity vA (vA = vA i) and rate of change of the heading angle ψ̇ are suitable to
describe the velocity of any point on the axle. The angular velocity of the axle is θ̇k, and
defining the distance between the two disks as t and calling it track, the velocities of the
centers of the wheels become
 
vF = vA + θ̇k × (t/2) j = vA − θ̇t/2 i
 
vD = vA + θ̇k × (−t/2) j = vA + θ̇t/2 i (3.15)

Assuming that the wheels are rolling without slip, the spin rate of each wheel is
   
vA − θ̇t/2 vA + θ̇t/2
ΩL = ΩR = (3.16)
d d
Kinematics Applications 91

where d is the wheel radius, and L and R denote left and right, respectively, indicating the
well-known fact that the two wheels have different angular velocities when a vehicle turns.
This was noticed early on in the automotive industry and led to the development of the
differential.

Example 3.3—Importance of Wheel Alignment


The discussion above is a reminder of the importance of wheel alignment. The axle in Section
3.4.2 moves because the wheels are parallel to each other. This can be demonstrated by an
instant center analysis.

FIGURE 3.7
Top view of axle: a) aligned wheels, b) unaligned wheels.

Consider first a properly aligned wheel and axle system, whose top view is shown in
Figure 3.7a. The instant center IC is located by first drawing the velocities of the wheels
to scale and then connecting the tips of the velocity vectors. The instant center is the
intersection of the line joining the tips of the velocities with the line going through the axle.
Next, consider the case where the wheels are not in alignment, shown in Figure 3.7b.
Without loss of generality, we assume that wheel D is in alignment and wheel F is not.
Because the two wheel velocities are no longer in the same direction, the instant center is
located by drawing perpendicular lines to the two velocities and finding where they intersect.
In this case, the instant center turns out to be at point F .
The instant center has zero velocity. But wait, point F is supposed to be moving and
hence should have a nonzero velocity. It follows that a system of unaligned wheels cannot
exist when both wheels are turning and the no-sliding condition is enforced. If friction is
sufficient to enforce the no-sliding condition and the wheels are not aligned, the left wheel
won’t turn and the axle will rotate about point F .
In reality, the friction forces are not strong enough to enforce the no-sliding condition at
F and the wheel and axle assembly moves in a way that there is rolling and sliding. Such
motion leads to important problems:
1. Since there is sliding at F and possibly also on the right wheel, the sliding of the tires
deteriorates the capability to steer and to control the vehicle. This leads to reduction of
stability and potentially dangerous driving conditions.
92 Applied Dynamics

2. The sliding friction forces result in faster and uneven wear of the tires, as well as increase
in tire temperatures. It is always a good idea to have a wheel alignment done after
replacing tires that have worn out unevenly or faster than expected.
On the other hand, in a vehicle having the wheels pointing towards each other by
the same amount, shown in Figure 3.8a for the front wheels and known as toe-in, is less
dangerous than when the wheels point to against each other. This latter situation, known as
toe-out and shown in Figure 3.8b, leads to a less stable ride, especially when tire flexibility
is considered (see Chapter 15.) Most vehicles are designed with a small amount of toe-in
for added stability.

!" #"

v v

FIGURE 3.8
Unaligned wheels (figure not to scale): a) toe-in, b) toe-out.

3.5 Bicycle Model of a Car


Modeling a four-wheeled vehicle is a complex task, especially when we include tire flexibility
effects. A simpler model that is frequently used is the bicycle model, or single-track model,
in which the vehicle is approximated by a rigid body, a front tire that is steered and a rear
tire that rolls on an axle. Figure 3.9 shows the top view. The wheelbase of the vehicle is L.
The model should more correctly be called a tricycle model because the lean of the vehicle
is not considered.
Let us count the degrees of freedom. The body of the car, without wheels, has three
degrees of freedom (rigid body moving on a plane). Select the three position variables as
coordinates of the center of mass XG , YG , and rotation angle θ. The steer angle δ adds one
more d.o.f. for a total of four. Each tire takes away a degree of freedom since there is no
sliding at A and B, resulting in a two-degrees-of-freedom system.
Three coordinate systems, XY Z, xyz and x0 y 0 z 0 , depict the geometry. The XY Z coor-
dinates are inertial. The xyz coordinates move with the body of the vehicle, while x0 y 0 z 0
are attached to the front wheel. The Z, z 0 , and z axes, which are not shown, denote the
vertical. In the absence of sliding, the wheels move in the direction of their heading and
their velocities are vA = vA i, vB = vB i0 . Thus, the no sliding constraints at A and B can
be expressed as

vA · j = 0 vB · j0 = 0 (3.17)

Let us next locate the instant center of the vehicle. Since the velocity directions of A
and B are known, we can draw perpendicular lines to the two velocities, and where the
Kinematics Applications 93



 

 



 

 

 

FIGURE 3.9
Bicycle model of a car: top view.

lines intersect is the instant center. For a vehicle, the instant center is frequently called the
turn center. The distance h from the instant center to the rear tire is the turn radius for
the rear wheel. The distance d from the instant center to the front tire is the turn radius
for the front wheel.
The information above can be used to relate the angular velocity of the vehicle to the
steer angle. From Figure 3.9, the steer angle is related to the wheelbase by tan δ = L/h.
When the steer angle is small, we can use the small angles assumption, tan δ ≈ δ, so that
L
δ = (3.18)
h
The distance h from the rear wheel A to the instant center is obtained as h = L/δ. The
speed of point A can be expressed as vA = hθ̇, and combining the two relationships leads
to an expression relating the angular velocity θ̇ of the vehicle to the steer angle δ as
vA δ
θ̇ = (3.19)
L
The equations of motion of the bicycle model will be considered in Chapter 5. Consider
next the lateral acceleration of a point, say, of point A. For point A, and considering normal-
tangential coordinates, the x direction is the tangential direction and the y direction is the
normal direction. It follows that the normal acceleration of point A is the lateral acceleration
and it has the form
2
vA v2 δ
aAn = aA · j = = A (3.20)
h L
The front wheels track a larger path than the rear wheels because the radius of curvature
h for the rear wheels is smaller than d, the radius of curvature for the front wheels. This is
not an important issue with passenger cars, but for trucks and buses it is, especially when
taking turns at lower speeds. Indeed, when a truck or bus takes a turn or exits a highway,
the driver must keep in mind the smaller path tracked by the rear wheels so that the rear
wheels will not hit the curb or leave the road surface. Specifications for the width and
curvature of highway entry and exit lanes take into consideration this difference in paths.
This section concludes with a few definitions pertaining to steering:
94 Applied Dynamics

• Steering wheel turn is the number of revolutions a steering wheel can make when
going from lock to lock (from as much steering in one direction to the other). In most
passenger vehicles, steering wheels can make three revolutions.
• Steering ratio is the ratio between the rotation angle of the steering wheel to the
rotation angle of the wheels. For most cars, this ratio is between 12 and 20. The steering
ratio is smaller for sporty and race cars. Some vehicles, such as those equipped with
rack-and-pinion steering, use a variable steering ratio. The steering ratio is larger in the
middle of the rack and becomes smaller as the steering wheel is turned more.
• Maximum steer angle is the largest steer angle a wheel can assume. For three revo-
lutions of the steering wheel and a steer ratio of 16, the maximum steer angle becomes

360 × revolutions/2 3 1
= 360 × × ≈ 34◦
steer ratio 2 16

• Minimum turn circle is the smallest circle that can be traversed by a vehicle (e.g.,
as the vehicle makes a U-turn). The diameter of this circle is calculated by 2L/ sin δmax .
For passenger vehicles, the minimum turn circle has a diameter around 35 ft. Many
people colloquially (and mistakenly) refer to the diameter of the turn circle as the turn
radius.

Example 3.4
To illustrate the substantial effect of steer angle on the vehicle angular velocity (yaw),
consider a vehicle with a wheelbase of L = 2.7 m, a speed of vA = 90 kph = 25 m/s, and a
steer angle of 1◦ = 1/57.296 rad, which is quite small. From Equation (3.19), the angular
velocity is
vA δ 1 1
θ̇ = = 25 × × = 0.1616 rad/s = 9.26◦ /s [a]
L 2.70 57.296
This is a high angular velocity for a small steer angle. Extreme levels of precision and
accuracy need to go into the design and manufacture of steering system components. A
small rotation in the steering wheel should accurately translate to a proportional value of
the steer angle, especially at high speeds. Loss of accuracy in steering, either due to wear
and tear or malfunctioning components, has a detrimental effect on the stability of a vehicle.
Next, calculate the lateral acceleration of point A. Using Equation (3.20) the lateral
acceleration is
v2 δ 1 1 2
aAn = A = 252 × × = 4.04 m/s [b]
L 57.296 2.7
which is larger than 0.4g. This value is nearing the limit lateral acceleration of most pas-
senger vehicles (∼ 0.7g). From Equation [b], the lateral acceleration is linearly proportional
to the steer angle, so that if in this problem the steer angle is doubled to 2◦ , the lateral
acceleration would be almost 1g. A small change in the steer angle can result in substantial
differences in lateral acceleration.
The turn radius (correct definition of the term turn radius used here) d of the rear wheel
and turn radius h of the front wheel can be calculated as
L 2.7 L 2.7
d = = = 154.71 m h = = = 154.68 m [c]
sin δ sin 1◦ tan δ tan 1◦
These two quantities are very close to each other, justifying the small angle assumptions
used earlier.
Kinematics Applications 95

Example 3.5
Calculate the turn radii of the front and rear wheels for a truck with a wheelbase of L = 9
m, given that the maximum value of the steer angle is δmax = 33◦ .
The steer angle of δmax = 33◦ is not small, so the small angle assumption is not used
here. The turn radii become
L 9 L 9
d = = = 16.525 m h = = = 13.859 m [a]
sin δmax sin 33◦ tan δmax tan 33◦
so that the turn diameters are 16.525 × 2 = 33.050 m for the front wheels and 13.859 × 2 =
27.718 m for the rear. This example demonstrates that there can be quite a difference
between the paths tracked by the front and rear wheels when the steer angle is large.

3.5.1 Where Is the Instant Center of a Car?


Consider the four-wheel model of a car (with proper wheel alignment). Each wheel rolls
without slipping and sliding, so each wheel moves in the direction of its heading. The steer
angles of the front left and right wheels are the same. To find the location of the instant
center, we draw perpendicular lines to each wheel, as shown in Figure 3.10.

 vB 

FIGURE 3.10
Instant center(s)? of a car.

There are three lines and the three lines do not meet at one point. So, where is the instant
center? It turns out that the turning motion of a four-wheeled vehicle is not possible without
sliding if the front tires have the same steer angle. The system becomes overspecified.
Counting the degrees of freedom helps understand this point.
The body of the car, without wheels, has three degrees of freedom, describing the two
translations and the rotation. The steering adds one more (δ) for a total of four. Each
front tire takes away a d.o.f. The no-sliding condition of the two rear tires, when perfectly
parallel to each other and perpendicular to the axle, reduces one additional d.o.f. so the
system is effectively down to one d.o.f. This degree of freedom is the translational motion
of the vehicle when the steer angle is zero, that is, moving along a straight line. Recall that
the simplified car (bicycle model) has two d.o.f.
In order for the turning motion of a vehicle to take place, the no sliding constraint
must be violated in one or both front tires. This is what happens in practice, as the lateral
acceleration due to turning overcomes the sliding friction forces.
96 Applied Dynamics

FIGURE 3.11
Ackermann steering.

The sliding discussed above was noticed in the 19th century, because carriage tires would
leave tire marks as they turned. The sliding problem disappears when the steer angles of
the front wheels are selected so that the three lines mentioned above intersect at one point,
as shown in Figure 3.11. This is known as the Ackermann steering condition.
Denoting the track of the vehicle, which is the distance from the two rear tires,2 by t
and the inside wheel (the one closer to the center of rotation) and outside wheel angles by
δo and δi , respectively, we can observe from Figure 3.11 that
L L
tan δi = tan δo = (3.21)
h h+t
Rewriting the steer angles in terms of cotangents (1/tangent) and subtracting the cotan-
gents gives
h+t h t
cot δo − cot δi = − = (3.22)
L L L
which is the Ackermann steering condition. The inside front wheel has a larger steer an-
gle than the outside front wheel in Ackermann steering. This difference in steer angles is
achieved by designing steering systems as linkages.3 There are several types of such designs.
No one design can satisfy the Ackermann condition exactly for all values of the steer angles,
but some designs come very close for a range of steer angles.
A trapezoidal steering design is shown in Figure 3.12 in the form of a four-bar linkage.
The arm length a and angle β are selected such that the Ackermann condition is satisfied
for as large a range of the steer angles as possible. A commonly used way to accomplish
this is by selecting the angle β such that the extensions of the arms meet at the center of
the rear axle, so that tan β = t/2L.
Modern cars do not use pure Ackermann steering, partly because it is based on a kine-
matic model, and hence, it ignores important dynamic and compliance effects associated
with tires, especially at high speeds. Some vehicles, especially in racing circles, prefer to
2 Some vehicles have different track values in the front and rear.
3 Linkages are studied in Section 3.8.
Kinematics Applications 97

FIGURE 3.12
Trapezoidal linkage for Ackermann steering.

have reverse Ackermann steering, where the steer angle of the outside wheel is larger than
the inside wheel. Table 3.1 summarizes the neutral, Ackermann, and reverse Ackermann
steering conditions.

TABLE 3.1
Steering types

Type of Steer Steer Angles


Neutral δi = δo
Ackermann δi > δo
Reverse Ackermann δi < δo

Example 3.6
A vehicle has a wheelbase of L = 9 ft and track of t = 5 ft. For a turn radius of h = 250 ft,
calculate the steer angles for the inside and outside wheels.
From Equation (3.21)
 
L 9
δi = tan−1 = tan−1 = 2.0618◦
h 250
 
L 9
δo = tan−1 = tan−1 = 2.0214◦ [a]
h+t 255
This is yet another example of the tremendous precision that is required of steering
mechanisms. The two steer angles differ from each other by only 2%.

Example 3.7—Instant Center(s) of Multi-Axle Vehicles


Consider the bicycle model of a multi-axle vehicle, such as a truck, characterized by two
rear axles and one front axle, as shown in Figure 3.13. Drawing the lines perpendicular to
the velocities of the wheels, it is clear that the perpendicular lines for the two rear tires are
parallel to each other. The truck will not have an instant center or be able to take a turn,
as long as the no-sliding assumption holds true.
98 Applied Dynamics

FIGURE 3.13
Bicycle model of multi-axle vehicle.

If in Figure 3.13 the rearmost axle (A) does not slide, then the velocity of the second
rear axle (D) becomes

vD = vA + ω × rD/A = vA i + ωk × ei = vA i + eωj [a]

Axle D has the same velocity as point A in the x direction but D also has a velocity
component in the lateral direction of magnitude eω, which is indicative of the sliding that
occurs at D. Conversely, if axle D does not slide, axle A will slide with a lateral velocity of
eω. If we increase the distance between the axles, that is, as e becomes larger, the amount
of slip increases. The further the rear axles are from each other, the more sliding will occur,
which is the reason why most trucks with two rear axles have the axles close to each other.
An approximate way to deal with vehicles with two rear axles is to assume that there is
no sliding at the midpoint between the axles (point E in Figure 3.13). In other words, we
assume that both rear axles will slide a certain amount.

3.6 Kinematic Differential Equations


Kinematic differential equations are a useful tool for analyzing the kinematics of bodies,
especially those subjected to nonholonomic constraints. The kinematic differential equations
give the rates of change of the displacement coordinates in terms of the velocity variables.
They are in the form of first-order ordinary differential equations, which can be integrated
to find the time histories of the position variables.
Kinematic differential equations do not, however, represent the complete description of
the motion. The kinetics of the system, as described by the equations of motion, is also
needed. The kinematic differential equations can be used on their own as a design tool
when it is relatively easy and straightforward to produce changes in the velocity variables.
Let us consider the bicycle model in Figure 3.9 and obtain the kinematic differential
equations. The coordinates of the center of mass G are XG and YG , and the body and steer
angles are θ and δ, respectively. It is assumed that the tires do not slide. Hence, the velocity
of point A is along the x axis, and the velocity of point B is along the x0 axis. The two
Kinematics Applications 99

nonholonomic constraints are

vA · j = 0 vB · j0 = 0 (3.23)

As discussed earlier, while the variables XG , YG , θ, and δ are independent of each other,
the nonholonomic constraints force their derivatives to be dependent on each other. Assume
that the steer angle is completely controllable by the driver. The velocity of the rear axle
(point A) is expressed as

vA = vG + ω × rA/G (3.24)

where rA/G = −ci. From Figure 3.9, we can write I = cos θi − sin θj, J = cos θj + sin θi, so
that the velocity of A becomes

vA = ẊG I + ẎG J + θ̇k × −ci

   
= ẊG cos θ + ẎG sin θ i + ẎG cos θ − ẊG sin θ − cθ̇ j = vA i (3.25)

leading to the nonholonomic constraint

ẎG cos θ − ẊG sin θ − cθ̇ = 0 (3.26)

Equation (3.25) can be solved for ẊG and ẎG as

ẊG = vA cos θ − cθ̇ sin θ ẎG = vA sin θ + cθ̇ cos θ (3.27)

From Equation (3.19), θ̇ = vA δ/L, so the kinematic differential equations can be written as
c c 1
ẊG = vA cos θ − sin θvA δ ẎG = vA sin θ + cos θvA δ θ̇ = vA δ (3.28)
L L L
Using XA , YA , and θ as the motion variables, we can show that the kinematic differential
equations have the form
1
ẊA = vA cos θ ẎA = vA sin θ θ̇ = vA δ (3.29)
L

3.7 Topspin and Backspin


Objects that move through the air (or any fluid) encounter a resistive force called drag
force. If the object has rotational motion as well as translational motion, depending on the
sense of rotation, the drag force acting on the body will have different values along the body.
These effects are considered in Chapter 4. Here, the terminology associated with rotation
of balls is introduced.
Consider two balls moving horizontally, viewed from the side, as shown in Figure 3.14.
The ball in Figure 3.14a has a clockwise spin, so that ω = ωJ. The velocity of point B, the
highest point on the ball, becomes

vB = vG + ω × rB/G = vI + ωJ × RK = (vG + Rω) I (3.30)

indicating that the velocity of point B is higher than the velocity of the center of the ball.
100 Applied Dynamics

 
 
 



 
 
 

 

FIGURE 3.14
Balls moving in air (side view): a) topspin, b) backspin.

Similarly, the velocity of point A, vA = (vG − Rω) I, is lower than the velocity of the center
of the ball. This type of motion is referred to as topspin.
By contrast, the ball in Figure 3.14b has a counterclockwise angular velocity and the
velocities of points A and B can be shown to be

vA = (vG + Rω) I vB = (vG − Rω) I (3.31)

and thus the top of the ball has a lower speed than the bottom. Such motion is referred to
as backspin.
Topspin and backspin of balls affect the motion of the ball during travel in the air,
changing the magnitude of the drag force, as well as when making contact with other
bodies, such as in shooting a basketball with a backspin; hitting a ball in tennis; throwing
a curveball, slider, or screwball in baseball; or when playing billiards.

3.8 Mechanisms
Mechanisms constitute an important application of kinematics for the purpose of transmit-
ting motion. Windshield wipers and steering and suspension systems in vehicles, pumps,
pistons of engines, door and casement window closers, door hinges, folding chairs, robots,
garden shears, power transmission in trains are all examples of mechanisms transmitting
motion from a source to an output. The history of mechanisms is full of ingenious as well
as failed designs. Figure 3.15 shows some mechanisms.
A mechanism is broadly defined as an assembly of components connected by movable
joints that transmit motion in a predetermined fashion. Mechanisms contain one or more
kinematic elements, such as linkages, cams, gears, belts, or chains. Many mechanisms can be
analyzed just by kinematic principles, as the forces they transmit are usually much smaller
than the forces their components can withstand. This section considers mechanisms that
transmit plane motion.
Kinematics Applications 101

FIGURE 3.15
Examples of mechanisms: a) door closer, b) locomotive wheels. Source: Wikimedia
Commons, http://commons.wikimedia.org/wiki/File:Mechanical door closer.jpg and
http://commons.wikimedia.org/wiki/File:Dinky the steam engine main drive wheel.jpg
(Last accessed August 1, 2014).

3.8.1 Links and Joints


Mechanisms are assembled using links and joints. A link is a component that possesses at
least two nodes, which are used as points of attachment to other links or to a body. The
connection between links is accomplished by means of joints. A collection of several links
connected by joints is called a kinematic chain or linkage.
Linkages can be classified into two types: open and closed. In an open linkage, also
known as an open chain, at least one of the links is connected to only one joint. In a closed
linkage, all links are connected to at least two other links. Figure 3.16 depicts open and
closed links.

a) b)
3
4 3

2 4
Links 2
Joints

1 (Ground) 1 (Ground)

FIGURE 3.16
a) Open and b) closed linkages.

The most widely used joint is the revolute joint. A revolute joint is also referred to as a
pin joint or turning pair. For plane motion, adding one link to another by a revolute joint
adds one d.o.f. to the system. We can also view the effect of joining two links by a revolute
joint as taking away two degrees of freedom from the combined d.o.f. of the independent
links connected by the revolute joint.
Consider plane motion and the link in Figure 3.16a. There are three links connected by
revolute joints. It is customary to consider the fixed surface to which link 2 is attached as
102 Applied Dynamics

a link itself. This link is referred to as the grounded link and is usually denoted as link
number 1. Link 2 in Figure 3.16a and each added link (links 3 and 4) add one d.o.f., for a
total of three.
The linkage in Figure 3.16b differs from Figure 3.16a through the addition of a pin joint,
which connects link 4 to the grounded link (link 1). The pin joint imposes two constraints,
so that the linkage in Figure 3.16b has one degree of freedom.
Another classification of links is by the number of nodes (attachments for joints) they
have. As shown in Figure 3.17, a binary link has two nodes. A ternary link has three and a
quaternary link has four nodes.

  

FIGURE 3.17
a) Binary link, b) ternary link, c) quaternary link.

Another widely-used joint is the prismatic or sliding joint, as shown in Figure 3.18. As
with a revolute joint, a prismatic joint also adds one d.o.f. to the component to which it is
attached. A common use of prismatic joints is to connect the sliding element to a link by a
revolute joint. The resulting joint is called a cylindrical pair.

FIGURE 3.18
Prismatic joint.

There are several other types of joints including those that involve three-dimensional
motion such as a joystick or a screw, as well as ball-and-socket joints. Such joints will be
discussed in Chapter 9.

3.8.2 Degrees of Freedom: Gruebler’s Equation


The number of degrees of freedom of a mechanism indicates the different number of ways
the mechanism can move and hence the number of controls needed to manage its motion.
For the most part, mechanisms are designed so that they have one degree of freedom. By
Kinematics Applications 103

providing a specified motion to one of the links, we can obtain a desired motion from the
other links.
We calculate the number of degrees of freedom for planar linkages using Gruebler’s
equation, which states
d.o.f. = M = 3n − 2J − 3G (3.32)
where n is the number of links, J is the number of joints, and G is the number of grounded
links (G is always 1 for open as well as closed links). In mechanism study, the number of
d.o.f. is also referred to as the mobility and is denoted by M .
As an example, let us calculate the mobility of the mechanism shown in Figure 3.19.
There are seven mobile links and one grounded link, referred to as link 1, for a total of
n = 8. There also are 10 joints. Note that one of the joints on the ternary link (link 5) is a
double joint, counting for two joints and link 8 is connected to the ground by a prismatic
joint. The mobility is
M = (3 × 8) − (2 × 10) − (3 × 1) = 1 (3.33)

    








   



FIGURE 3.19
A linkage.

Note that there are exceptions to Gruebler’s equation, which does not consider the sizes
and shapes of the links. In these exceptions, Gruebler’s equation predicts zero mobility, yet
the mechanism still moves. The mechanisms in Figure 3.20a and Figure 3.20b each have
one d.o.f. when certain geometrical relationships are satisfied. In Figure 3.20a, the exception
occurs when L2 = L3 = L4 , and all links are all parallel to each other, making one of the
links redundant, so that the redundant link does not affect the kinematics of the other links.

Figure 3.20b represents two spur gears, which can be considered a mechanism with three
links (the two gears and the grounded link) and three joints (the two pin joints and the
point of contact between the links), resulting in a mobility of M = 3 × 3 − 2 × 3 − 3 × 1 = 0.
Such a linkage is represented in Figure 3.20c. The exception occurs when there is no sliding
between the two gears (which in ensured by the gear teeth) and r1 + r2 = L, where r1 and
r2 denote the pitch radii of the two gears.4
4 Figure 3.20c is proof that the triangle is the most stable structural member, with a mobility of zero. A

truss is, in essence, a collection of triangular structural members.


104 Applied Dynamics

  

 
   




FIGURE 3.20
Exceptions to Gruebler’s equation: a) redundant link, b) spur gear. c) Triangular member,
or truss, which has zero mobility.

In a mechanism designed with redundant components, the dimensions of the components


must be extremely accurate; otherwise the mechanism will not move or it will deform during
motion.

3.8.3 Four-Bar Linkage and Slider-Crank Mechanism

 





 

 

FIGURE 3.21
a) The four-bar linkage and b) the slider-crank mechanism.

Shown in Figure 3.21a and Figure 3.21b, the four-bar linkage and slider-crank mech-
anisms are two of the most important linkages in the study of kinematics. Each has one
degree of freedom. If we examine complex mechanisms, we can see that they are composed
of two or more four-bar linkages or slider-cranks connected in a special way. For example,
the mechanism in Figure 3.19 can be viewed as the combination of two four-bar linkages
(1-2-3-4 and 1-4-5-6) and a slider-crank (1-5-7-8, considering that the motion of the ternary
link, link 5, is completely defined by the two four-bar mechanisms).
Mechanisms find different uses depending on the link sizes and the link that is grounded,
as shown in Figure 3.22. Figure 3.22a is a vise grip and Figure 3.22b is a hand pump.
Kinematics Applications 105

 

 

 


FIGURE 3.22
Uses of mechanisms: a) vise grip (four-bar), b) hand pump (slider-crank).

3.9 Instant Center Analysis for Linkages


Section 2.11 demonstrated that a body undergoing plane motion has an instantaneous
center of zero velocity (instant center), about which the body rotates at that instant. Here
we extend the concept of instant center to linkages and mechanisms.
The most general definition of an instant center (IC) is as follows: An instant center is
a point common to two bodies in plane motion, and it has the same instantaneous velocity
in each body. The instant center of two bodies can lie on one body, both bodies, or outside
both bodies. The two bodies do not need to be connected to each other. The number of
instant centers is n(n − 1)/2, where n is the number of linkages. An instant center can have
nonzero velocity.
The definition of instant center that we studied earlier in Section 2.11 is the instant
center between the body under consideration and the grounded link, that is, a fixed surface
from which the motion is measured. As the grounded link has zero velocity, so does the
instant center. An IC that has zero velocity is a center of rotation, and it thus becomes a
useful tool for velocity analysis.
Instant centers have a useful property. Known as Kennedy’s rule, this property facilitates
the finding of ICs, and it can be stated as follows. Any three bodies on the same plane have
three ICs (for n = 3, n (n − 1) /2 = 3), and the three ICs lie on the same line. The three
bodies do not need to be connected.
The following notation is widely used when dealing with linkages and their instant
centers. The linkages are denoted by 1, 2, 3, . . . . The grounded link is labeled as link no. 1.
The instant center of links j and k is denoted by Ijk (or Ikj ), with the smaller index usually
placed first. The joints attached to grounded links are denoted by O1 , O2 , . . . . Other joints
are denoted by J1 , J2 , . . . .

3.9.1 Locating Instant Centers in Linkages


Consider the four-bar linkage in Figure 3.23 and label the links and joints according to
the convention above. To find the instant centers, we begin with what is known from the
geometry and find the rest of the ICs by means of Kennedy’s rule.
There are six ICs (n = 4 =⇒ n (n − 1) /2 = 6) of which four can be located from the
geometry. These four ICs (I12 , I23 , I34 , I14 ) are the four joints. ICs I12 and I14 have zero
velocity as they are associated with the grounded link (link 1).
106 Applied Dynamics

I13
3 v3
I23 L3
I34
3
L4
L2 2 v2
2 4 4
I24 I12
I14
1

FIGURE 3.23
Four-bar linkage and its instant centers.

We need to locate I13 and I24 . Consider I13 first and Kennedy’s rule. Links 1, 2, and
3 have three common ICs I12 , I23 , I13 and these ICs lie on a straight line. Similarly,
associated with links 1, 3, and 4 are three ICs: I13 , I34 , I14 . These ICs also lie on a
straight line. Two lines are drawn next, one going through I12 , I23 and the other going
through I14 , I34 . Their intersection is I13 . Similarly, I24 is found as the intersection of two
lines, one going through I12 , I14 and the other through I23 , I34 . Note that I13 has zero
velocity.
Next, consider the slider-crank mechanism in Figure 3.24. The procedure for finding the
instant centers is similar to the four-bar linkage. The difference is in the treatment of the
prismatic joint. As link 4 slides over a straight line and the radius of curvature of a straight
line is at infinity, instant center I14 is at infinity and it is located by drawing a perpendicular
to the line over which the slider moves.

I13

I24
I14 (at )
I23 =
3
= I34
2 v
2 v
I12 4
1

FIGURE 3.24
Slider-crank mechanism and its instant centers.

After locating I12 , I23 , I34 , I14 from the geometry, we proceed to find ICs I13 and I24 .
I13 is located as the intersection of the two lines, one going through I12 , I23 and the other
through I14 , I34 . Note that I14 is at infinity and I13 lies on the line that was drawn to mark
the location of I14 .
Finding I24 requires the intersection of two lines, one going through I12 , I14 and the
Kinematics Applications 107

other through I23 , I34 . The line going through I23 , I34 is easy to draw. But how about the
line going through I12 , I14 ? Since I14 is at infinity we draw a line that goes through I12
which is parallel to the line (connecting I34 and I13 ) that was drawn as perpendicular to
the motion of the slider. I24 is the intersection of those two lines, as shown in Figure 3.24.
In a mechanism that has more than one degree of freedom, we cannot find all the ICs
by means of the geometry alone. Additional information about the velocities is needed.
For mechanisms that have several links, we can make use of an instant center locating
tool, shown in Figure 3.25, that facilitates locating instant centers. A circle is drawn, and
the links are marked on the circumference. Once the location of an IC of two links is
determined, a solid line is drawn between the numbers of the two links.

FIGURE 3.25
Diagram for locating instant centers.

For the four-bar linkage above, instant centers I12 , I23 , I34 and I14 are located by inspec-
tion and solid lines are drawn on the locating tool. The diagram then provides suggestions
as to which sets of points to join so that we can locate the other ICs. Drawing a dashed
line between points 1 and 3, we can form two triangles, 1-2-3 and 1-3-4. We draw two lines:
one joining ICs I12 and I23 and the other connecting ICs I14 and I34 . The intersection of
these two lines is I13 . The procedure is repeated for I24 .
The instant center locating tool becomes more useful when there are more than four
links. Note that for mechanisms with several links it is not always necessary to calculate
the locations of all of the instant centers. We usually find only the ICs that are needed for
analysis or design.

Example 3.8
Locate the instant centers of the mechanism in Figure 3.26.
This mechanism has six linkages and seven joints, so its mobility is

M = 3n − 2J − 3 = (3 × 6) − (2 × 7) − 3 = 1 [b]

The number of instant centers is n (n − 1) /2 = 15. Instant centers I12 , I23 , I34 , I15
and I56 are located by observation, as they are located at the pin joints. Also, I14 and I46
are at infinity. A total of seven ICs are identified by visual inspection and the remaining
eight ICs need to be located. The instant center locating tool, where we draw solid lines
between the instant centers that are located, as shown in Figure 3.27, comes in handy.
To locate I13 we draw a dashed line between links 1 and 3 on the IC locating tool.
Triangles 1-3-4 and 1-2-3 each have two solid lines, indicating that we can locate I13 as the
intersection of the extensions of two lines, one line joining I12 and I23 and the other joining
108 Applied Dynamics



FIGURE 3.26
Six-bar mechanism.

 

 

FIGURE 3.27
Kinematic tool for six-bar mechanism.

I14 and I34 . The result is shown in Figure 3.28. Note that some of the instant centers are
not shown in Figure 3.28. For example, I26 is not shown, as it extends quite far out of the
page. Also, lines that need to be drawn to locate I36 and I25 are not shown.
The other instant centers are found in a similar way. Table 3.2 shows the instant centers
that need to be connected in order to find the remaining instant centers. As we progress in
locating the unknown instant centers, we may have several choices of linkage triples. Also,
using ICs with zero velocity facilitates the process.

3.9.2 Velocity Analysis Using Instant Centers


One can calculate velocities of joints or of points on links, as well as angular velocities
of links, by means of instant centers. For a one d.o.f. mechanism, all that is needed is a
single piece of velocity (or angular velocity) information, such as the velocity of a joint or
angular velocity of a link. We can move from one IC to another to determine the other
velocities. ICs associated with grounded links are particularly useful, as they are centers of
rotation. It usually is not necessary to locate all the instant centers for a velocity analysis.
The procedure is best described by means of an example.
Kinematics Applications 109



 
  
 


   


 







FIGURE 3.28
Instant centers of the six-bar mechanism.

Example 3.9
For the four-bar linkage in Figure 3.23, given that the angular velocity of link 2 is ω2
clockwise, find the angular velocity of link 4.
Denote the link lengths by Lj and velocities of the joints by vj , (j = 1, 2, 3, 4). The
velocity of joint J2 , which is I23 , can be found by
v2 = L2 ω2 [a]
where L2 is the length of link 2. Next, make use of I13 and write the velocity of I23 as
v2 = I13 I23 ω3 [b]
in which I13 I23 denotes the distance between I13 and I23 . This distance can be calculated
using geometry. It should be noted here that before we conduct an instant center analysis, we
should know (or should have calculated) all distances and angles from the given information.
Using the instant center I13 , which has zero velocity, the angular velocity of the third
link can be calculated as
v2 L2 ω2
ω3 = = [c]
I13 I23 I13 I23
and it is counterclockwise. Since the angular velocity of link 3 is now known, the velocity
of joint J3 can be found using
v3 = I13 I34 ω3 [d]
where, again, I13 I34 is found from the geometry. The velocity of joint J3 can also be found
from the angular velocity of link 4 using v3 = L4 ω4 . Combining this relationship with the
above equation yields the angular velocity of link 4 as
v3 I13 I34 L2 ω2
ω4 = = ×  [e]
L4 L4 I13 I23
110 Applied Dynamics

TABLE 3.2
List of link triples used to locate instant centers

To Find Use Link Triples And Connect


I13 1-2-3, 1-3-4 I12 with I23 and I14 with I34
I45 1-4-5, 4-5-6 I14 with I15 and I46 with I56
I16 1-4-6, 1-5-6 I14 with I46 and I15 with I56
I24 1-2-4, 2-3-4 I12 with I14 and I23 with I34
I35 1-3-5, 3-4-5 I13 with I15 and I34 with I45
I36 3-5-6, 1-3-6 I35 with I56 and I13 with I16
I26 1-2-6, 2-3-6 I12 with I26 and I23 with I36
I25 1-2-5, 2-3-5 I12 with I15 and I23 with I35

where L4 is the length of link 4, or L4 = I14 I34 . The angular velocity of the fourth link is
clockwise. This makes sense, as a clockwise rotation of link 2 causes a clockwise rotation of
link 4.
The lengths between the links and the instant centers can be calculated from the geom-
etry.5 If the linkage is drawn to scale, we can measure the distances involved by a ruler and
obtain a ballpark estimate of the angular velocities and velocities. This feature of velocity
analysis by instant centers is useful in design.

3.10 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Beer, F.P., Johnston, E.R., and Cornwell, P., Vector Mechanics for Engineers: Dynamics,
10th Edition, McGraw-Hill, 2012.
Erdman, A.G., Sandor, G.N., and Kota, S., Mechanism Design, Vol. 1, 4th Edition, Prentice-
Hall, 2001.
Meriam, J.L., and Kraige, L.G., Engineering Mechanics: Dynamics, 7th Edition, Wiley,
2012.
Myszka, D.H., Machines and Mechanisms, 4th Edition, Prentice-Hall, 2011.
Norton, R.L., Design of Machinery: An Introduction to the Synthesis and Analysis of Mech-
anisms and Machines, 4th Edition, McGraw-Hill, 2008.

3.11 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.
5 If not, first use the geometry and calculate all link lengths and angles.
Kinematics Applications 111

Section 3.2—Motion with Respect to the Rotating Earth


3.1 (E) A race car is traveling eastward with a speed of 125 kph in Highland Park, New
Jersey. Calculate the magnitude of the Coriolis acceleration. Assuming the Coriolis accel-
eration is the only acceleration of the vehicle, calculate the deflection due to the Coriolis
acceleration that is generated in ten seconds of driving.
3.2 (E) Every other year (or so) a second is added to standard time to account for the
slowing of the angular velocity of the Earth. Calculate the angular deceleration (assume it
is constant) of the Earth if we add one second to standard time every two years.
3.3 (C) An experimental vehicle travels from the Equator to the South Pole along a contin-
uous railway track. The vehicle moves at a constant speed of 350 kph relative to the Earth.
Calculate and plot the Coriolis acceleration aCoriolis as a function of latitude angle λ.

Section 3.4—Rolling

Planet
a) b)
B P
Gear D
r Arm

R+r C B A
G
arm
O O

R 3 rad/s

FIGURE 3.29
Figures for Problems a) 3.4 and b) 3.5.

3.4 (M) The planetary gear in Figure 3.29a can be viewed as the inverse of Example 3.2.
The planet (smaller gear is rotating clockwise with angular velocity ωP = 3 rad/s cw. The
dimensions are R = 10 cm, and r = 1.3 cm. Obtain the relationship between the angular
velocity of the planet and the angular velocity of the arm connecting the gears when i)
the gear (large gear) is stationary, and ii) when the gear is rotating counterclockwise with
angular velocity ωG = 0.3 rad/s.
3.5 (D) The arm OB in Figure 3.29b rotates with constant angular velocity ωarm = 3 rad/s
ccw, with point O stationary. Find the following using instant centers: i) angular velocity
of gear B if gear D is fixed, ii) angular velocity of gear B if D is rotating with a clockwise
angular velocity of 2.5 rad/s.
3.6 (M) The disk in Figure 3.30 of radius R rolls without slipping with constant angular
velocity Ω. Carved inside the disk is a slot and a mass moves inside the slot. Denoting the
position of the mass inside the slot by s, calculate the velocity and acceleration of the mass
as a function of θ.
112 Applied Dynamics


 

  
 


  

FIGURE 3.30
Figure for Problem 3.6.

Section 3.5—Bicycle Model of a Car


3.7 (M) Consider the bicycle model in Example 3.4. Find the acceleration of point B,
considering that point A is accelerating with 0.1g and the angular acceleration of the vehicle
is zero.
3.8 (M) Consider the bicycle model of a vehicle (Figure 3.9) of wheelbase 10 ft. Point A,
the rear axle, is moving with speed 50 mph, which is increasing at the rate of 2 ft/sec2 . The
steer angle is δ = 0.9◦ and is constant. Find the acceleration of point B and the radius of
curvature at B.
3.9 (E) A vehicle of wheelbase 2.9 m and track 1.22 m is taking a turn. If the turn radius
for the center of the rear axle is 80 m, calculate the steer angles for the inside and outside
front wheels if the Ackermann steering condition is to be satisfied.

A B
v x
r f
L

FIGURE 3.31
Figure for Problem 3.10.

3.10 (M) Rear wheel steering is a relatively new technology, where the rear wheels are
steered in addition to the front wheels. The steer angle for the rear wheel is usually computer
controlled. Rear wheel steering for the bicycle model is shown in Figure 3.31. Given a vehicle
with wheelbase L = 2.85 m and front steer angle of δf = 5◦ , calculate the turn radius for
the rear wheel when the rear steer angle is a) δr = 3◦ , b) δr = −3◦ . Compare the results you
get with the turn radius when δf = 8◦ and δf = 2◦ , with δr = 0◦ . Comment on when you
would prefer a negative rear steer angle and when a positive rear steer angle is preferred.
3.11 (M) Figure 3.32a shows the top view of a trailer and Figure 3.32b shows the side view.
Both the front and rear vehicles are modeled by assuming that all wheels roll without slip
or slide. Locate the instant centers of the vehicle and of the trailer for when a = 0 and
calculate the angular velocity of the trailer, given the speed vA and steer angle δ of the
front vehicle.
Kinematics Applications 113

FIGURE 3.32
Figure for Problems 3.11, 3.12, 3.14, and 8.9.

3.12 (M) For the problem above, derive the condition for when the turn radius of D is larger
than the turn radius for A. Do this for when LR = 2LF .

Section 3.6—Kinematic Differential Equations


3.13 (E) Derive the kinematic differential equations in Equation (3.29).
3.14 (D) Figure 3.32a shows the top view of a trailer and Figure 3.32b shows the side view.
Both the front and rear vehicles are modeled by assuming that all wheels roll without slip
or slide. Treat the steer angle as a known quantity (driver-provided input) and determine
the number of degrees of freedom and constraints. Then, using XA , YA , XD , YD , θF and
θR as position variables, write the kinematic differential equations using vA as the velocity
variable.
3.15 (C) Numerically integrate the kinematic differential equations in Equation (3.29) to
simulate parallel parking of a vehicle. The parallel parking maneuver should consider a
parking spot of length 4.1 m., width 1.7 m., and you should make sure that the front of the
vehicle does not hit the car in front of it. The vehicle is of wheelbase 3.4 m and width 1.4
m. You can use a profile like a sinusoidal profile, or another profile of your choice. You can
use a constant vehicle speed vA that is small, such as 0.2 m/s. When parked, the vehicle
should be parallel to the curb (less than 5◦ angle with the curb) and be at a distance not
wider than 0.25 m from the curb.

Section 3.8—Mechanisms
3.16 (E) Show that a mechanism with an odd number of links can never have one degree of
freedom.
3.17 (E) Calculate the number of degrees of freedom of the mechanism in Figure 3.33.
Identify the four-bar and slider-crank components of the mechanism.
3.18 (E) Calculate the number of degrees of freedom of the mechanism in Figure 2.44.

Section 3.9—Instant Center Analysis for Linkages


3.19 (E) Locate all of the instant centers of the mechanism in Figure 3.34 geometrically.
114 Applied Dynamics

3
L=20 5
L=25
2 6
L=12
4
40° L=18 10

1
L=16

FIGURE 3.33
Figure for Problems 3.17, 3.20, and 3.23.

3 4

2
1 (ground)

FIGURE 3.34
Figure for Problem 3.19.

3.20 (M) Locate all of the instant centers of the mechanism in Figure 3.33 geometrically.
3.21 (M) Locate all of the instant centers of the mechanism in Figure 2.47b geometrically.
3.22 (M) Consider the slider-crank mechanism in Figure 3.24. Given the length of the crank
as L and slider arm length of 2L, calculate the velocity of the slider as a function of the
crank angular velocity ω ccw, when the crank is at an angle of 30◦ with the horizontal.
3.23 (D) Given that link 2 in the mechanism in Figure 3.33 is rotating with angular velocity
ω, calculate the velocity of the slider.
4
Kinetics Fundamentals

4.1 Introduction
This chapter discusses basic principles of dynamics. The primary interest is in particle kinet-
ics in two or three dimensions and in planar rigid body dynamics. We develop expressions
for linear and angular momentum and then state the translational and rotational laws of
motion.
We analyze forces that act on bodies, such as friction, gravity, contact, aerodynamic,
springs and dampers. The following chapter takes the concepts developed here and applies
them to a variety of systems.
While the expressions for momenta and laws of motion are valid for two- or three-
dimensional motion, the examples considered in this and the next chapter are primarily
for two-dimensional motion, especially for rigid bodies. Three-dimensional motion of rigid
bodies will be discussed in Chapter 11.

4.2 Rigid Body Geometry


In dynamics, components of a system are modeled as particles, rigid bodies, or deformable
bodies. A particle is defined as a body with no physical dimensions whose mass is con-
centrated at one point. A particle undergoes only translational motion; thus, bodies whose
motion is primarily translational can be modeled as particles. An idealized particle cannot
have a rotation. By contrast, a rigid body has physical dimensions and a shape. The rigidity
assumption states that the shape of the body does not change under the action of forces.
To describe the characteristics of a particle we only need to know its mass. For a rigid
body, additional information is needed on how the mass is distributed. This section discusses
rigid body geometry and the distribution of mass on a body.

4.2.1 Center of Mass


Consider the system of N particles in Figure 4.1. A reference point, or origin, O is intro-
duced, from which displacements are measured. The mass of the i-th particle is mi and
position of particle i is denoted by ri (i = 1, 2, . . . , N ). The center of mass of the system is
denoted by G and its location is defined by the vector rG as

N
1 X
rG = mi ri (4.1)
m i=1
P
in which m = mi is the total mass of the system.

115
116 Applied Dynamics
mi
m(
mN
G
!i m)
!G
!)
O !"#$%&'

FIGURE 4.1
System of particles.

dm

r G

rG
O (fixed)

FIGURE 4.2
A rigid body.

Consider now a rigid body, such as the one in Figure 4.2. The rigid body can be viewed
as a system of particles where the number of particles approaches infinity. Replacing N in
Equation (4.1) with ∞, the summation with integration, and mi with the differential mass
dm, leads to the expression for the center of mass for the rigid body as
Z
1
rG = rdm (4.2)
m body
R
The total mass is m = body dm. The position of a differential mass with respect to the
center of mass is r = rG + ρ . Introduction of this expression into the above equation gives
Z Z Z
1 1 1
rG = rdm = (rG + ρ ) dm = rG + ρ dm (4.3)
m body m body m body

leading to the conclusion that


Z
ρ dm = 0 (4.4)
body

This equation indicates that the weighted average of the position vector measured from the
center of mass is zero.
For a body with a complex geometry, we can calculate the location of the center of mass
by separating the body into components whose individual centers of mass can be calculated
more easily. Denoting the mass and the center of mass location of each individual component
Kinetics Fundamentals 117

by mi and rGi , respectively, the expression for the center of mass becomes
N
1 X
rG = mi rGi (4.5)
m i=1
P
where m = mi .

Example 4.1
Calculate the center of mass of the composite shape in Figure 4.3. The rod weighs 2 lb and
the disk would weigh 3 lb in the absence of the hole.
y

Rod

4" 2 lb

2"

O x
Hole
3"
1.5"
3 lb

FIGURE 4.3
Composite shape.

It is convenient to split the composite shape into three parts: the disk without the hole,
which has a weight of 3 lb and radius 3 in.; the hole (negative mass) of radius 1 in.; and
the rod, which is of length 4 in. and weight 2 lb. Select a coordinate system xy and place
a reference point O at the center of the disk. Table 4.1 shows the mass (weight) and the
center of mass location of each shape. The negative mass of the hole is obtained by noting

TABLE 4.1
Mass and center of mass location

Object Weight (lb) Mass (slug) Center of Mass (in.)


Disk without Hole 3 3/32.17 0i+0j
Hole −1/3 −1/96.51 1.5 i + 0 j
Rod 2 2/32.17 0i+5j

that the radius of the hole is one third the radius of the disk, so the hole removes 1/9 of
the mass of the disk, 3 × 1/9 = 1/3 lb. The total weight is
1
W = 3− + 2 = 4.667 lb [a]
3
118 Applied Dynamics

Using Equation (4.5), the location of the center of mass becomes


 
1 1
rG = − × 1.5i + 2 × 5j = −0.1071i + 2.1429j in. [b]
4.667 3

4.2.2 Mass Moment of Inertia


The center of mass represents the point at which the entire mass of a body can be viewed as
being concentrated for the purpose of describing translational motion. It does not provide
any information regarding the shape of the body or how the mass is distributed about the
center of mass. The mass moment of inertia quantifies the mass distribution of the body.
The mass moment of inertia also represents the resistance of a body to rotational motion.

q dm
D
y , 
dy 
rG/D
x
G
dx x

FIGURE 4.4
Rigid body in plane motion.

Consider the rigid body in Figure 4.4 and an arbitrary point D on the body. Also consider
a differential element on the body, so that the position vector from D to the differential
element is q = rG/D +ρρ. The xy coordinate system is attached to the body, so it represents
a moving coordinate system.
The interest is in the distribution of the mass about an axis perpendicular to the plane
of motion and going through point D. From elementary dynamics, we know that the mass
moment of inertia of a point mass m about a point D at a distance d from it is ID = md2 ,
as shown in Figure 4.5. The mass moment of inertia of the differential mass in Figure 4.5
about point D is denoted by dID and defined as

dID = q · q dm (4.6)

The mass moment of inertia of the entire body about point D is obtained by integrating
the above expression over the body
Z
ID = q · q dm (4.7)
body

Introduction of the expression for q in terms of the center of mass location to the above
equation gives
Z

ID = rG/D · rG/D + rG/D · ρ + ρ · rG/D + ρ · ρ dm (4.8)
body
Kinetics Fundamentals 119

"

d m

!
ID = md 2
D

FIGURE 4.5
Mass moment of inertia of point mass.

Let us examine each term in the above equation. The vector rG/D is not related to the
integration and to dm; thus, the first term in the above equation becomes mrG/D ·rG/D . The
secondR and third terms vanish because of the definition of the center of mass in Equation
(4.4), body ρ dm = 0. Considering the xyz coordinate system moving with the body, where
the xy plane is the plane of motion, the position vector from the center of mass to the
differential element can be written as ρ = ρx i + ρy j. The vector from D to the center of
mass is rG/D = dx i+dy j. Using these definitions, we can express the mass moment of inertia
about point D as
Z
2 2
ρ2x + ρ2y dm
 
ID = m dx + dy + (4.9)
body

When point D is selected as the center of mass, dx = dy = 0 and the mass moment of
inertia about an axis perpendicular to the plane of motion and going through the center of
mass G becomes
Z
ρ2x + ρ2y dm

IG = (4.10)
body

The mass moment of inertia IG is a positive quantity and it is a measure of how the mass
is distributed about the center of mass. However, it does not describe the actual shape of
a body. Two bodies with the same mass and the same mass moment of inertia can have
totally different shapes.
Introducing Equation (4.10) into Equation (4.9), the mass moment of inertia of a body
about an arbitrary point D can be written as

ID = IG + m d2x + d2y = IG + md2



(4.11)
q
in which d = d2x + d2y is the distance between D and G. The above equation is known as
the parallel axis theorem and is a useful tool for calculating the mass moment of inertia.
Many times, we know the mass moment of inertia of a common shape about a certain point,
and we use the parallel axis theorem to find the mass moment of inertia about another point.
The following are of interest:
• The parallel axis theorem relates the mass moment of inertia about the center of mass
to the mass moment of inertia about another point. It cannot be used to relate the mass
moments of inertia about any two points.
• Mass moment of inertia has its lowest value when calculated about the center of mass.
120 Applied Dynamics

For a composite body we can separate the body into a series of bodies whose masses,
centers of mass, and mass moments of inertia are known, or are easy to calculate. The
respective mass moments of inertia are then added to find the mass moment of inertia of
the composite body.
Mass moments of inertia of commonly encountered shapes are give in the Appendix.

Example 4.2
Calculate the centroidal (about the center of mass) mass moment of inertia of the composite
body in Example 4.1.
As in the previous example, it is convenient to break the body into three parts, calculate
the mass moments of inertia of each part about its center of mass, and use the parallel axis
theorem to add the moments of inertia. For the disk,

Idisk = IGdisk + mdisk d2disk [a]

where  2
1 1 3 3
IGdisk 2
= mdisk Rdisk = = 29.142 × 10−4 slug · ft2 [b]
2 2 32.17 12
4.6033 2
d2disk = |rG − 0|2 = 0.1072 + 2.14292 = 4.6033 in.2 = ft [c]
144
We do the same for the hole and the rod, with the result

Ihole = IGhole + mhole d2hole [d]

where  2
1 1 1
IGhole = − = −0.3598 × 10−4 slug · ft2 [e]
2 96.51 12
2
d2hole = (1.5 + 0.1071) + 2.14292 = 7.1747 in.2 [f ]
The mass moment of inertia of the rod is

Irod = IGrod + mrod d2rod [g]

where  2
1 2 4
IGrod = = 5.7565 × 10−4 slug · ft2
12 32.17 12
2
d2rod = 0.10712 + (5 − 2.1429) = 8.1747 in.2 [h]
Table 4.2 gives the calculated information. The individual moments of inertia are added to
calculate the total mass moment of inertia.

4.3 Linear Momentum and Angular Momentum


Linear and angular momenta are two fundamental quantities that describe the motion
characteristics of bodies. The linear momentum of a particle is defined as its mass times its
velocity

p = mv (4.12)
Kinetics Fundamentals 121

TABLE 4.2
Mass moments of inertia of composite body

Object Wt. (lb) IG × 104 (slug·ft2 ) d2 (in.2 ) I × 104 (slug·ft2 )


Disk without Hole 3 29.142 4.6033 58.953
Hole −1/3 −0.3598 7.1747 −5.5765
Rod 2 5.7511 8.1747 41.050
Entire Object 4.667 94.426

The linear momentum of a body denotes the tendency of that body to continue its
translational motion. The linear momentum of a rigid body can be written as the sum of
the linear momentum of each particle that constitutes the rigid body. Consider Figure 4.4
and write the velocity of a differential element in terms of the velocity of the center of mass
as

v = vG + ω × ρ (4.13)

in which the vector ρ connects the center of mass and the differential element dm. The
linear momentum of the differential element is dp = dmv. Integration of this expression
over the entire body gives the linear momentum of the body as
Z Z
p = v dm = (vG + ω × ρ ) dm (4.14)
body body
R
From the definition of the center of mass and Equation (4.4), body
ρ dm = 0, and the
linear momentum of a rigid body becomes

p = mvG (4.15)

Linear momentum is an absolute quantity, and it is independent of the coordinate system


used to describe the motion. The above definition of linear momentum is valid for plane as
well as three-dimensional motion.
Analogous to linear momentum and translational motion, angular momentum describes
the tendency of a body to continue to rotate. Because a rotation takes place about an axis
or about a point, angular momentum is defined about a point, so that its value depends on
the point about which it is calculated.
Consider the point mass in Figure 4.5, which has velocity v and linear momentum
p = mv. The angular momentum about point D is denoted by HD and defined as the
moment of the linear momentum vector of the body about that point, that is, the cross
product between the position vector from D to mass (d) and the linear momentum vector.
Thus,

HD = d × p = d × mv (4.16)

To extend this definition to rigid bodies consider a rigid body, such as the one in Figure
4.4, a point D on the body, and a differential element of mass dm. The velocity of the
differential element is v = vG + ω × ρ , and its linear momentum is

dp = vdm = dm (vG + ω × ρ ) (4.17)

The angular momentum of the differential element about point D is denoted by dHD and is
122 Applied Dynamics

defined as the cross product of the position vector q between D and the differential element
with the linear momentum

dHD = q × dp = q × dm (vG + ω × ρ ) (4.18)

The angular momentum of the body is calculated by integrating the above expression
over the entire body
Z
HD = q × (vG + ω × ρ ) dm (4.19)
body

The position vector from D to the differential element can be expressed in terms of the
center of mass as q = rG/D + ρ . Introducing this to the definition of angular momentum
gives
Z
HD = (rG/D + ρ ) × (vG + ω × ρ ) dm
body

Z

= rG/D × vG + rG/D × (ω
ω × ρ ) + ρ × vG + ρ × (ω
ω × ρ ) dm (4.20)
body

Both rG/D and vG are not related to the integration variable. The first term in the
above equation reduces to mrG/D × vRG . The second and third terms vanish because of
the definition of the center of mass, body ρ dm = 0, so that the expression for angular
momentum about D reduces to
Z
HD = mrG/D × vG + ρ × (ω
ω × ρ ) dm (4.21)
body

The above expression for the angular momentum about D is valid for both plane as
well as three-dimensional motion. For plane motion, the second term in the above equation
can be simplified. Consider a set of coordinates xyz attached to the body and denote the
plane of motion as the xy plane, so that the angular velocity is ω = ωk. We can write
ρ = ρx i + ρy j. It follows that ω × ρ is on the xy plane and ρ × (ω
ω × ρ ) is in the z direction.
Evaluation of the cross product gives

ρ × (ω
ω × ρ ) = ρ × (ωk × (ρx i + ρy j))

= (ρx i + ρy j) × (ωρx j − ωρy i) = ω ρ2x + ρ2y k



(4.22)

with the result


Z Z
ω ρ2x + ρ2y dm k = IG ωk

ρ × (ω
ω × ρ ) dm = (4.23)
body body

where IG = body ρ2x + ρ2y dm is the mass moment of inertia of the body about the center
R

of mass. Hence, as described in Figure 4.6, the general expression for angular momentum
about an axis perpendicular to the plane of motion and going through point D becomes

HD = mrG/D × vG + IG ωk (4.24)

The interest is in the case where point D is selected as the center of mass, point G. In
Kinetics Fundamentals 123


m, IG

G vG

rG/D
D

FIGURE 4.6
A body with translational and angular velocity.


t

h
O

FIGURE 4.7
Rotation about a fixed point.

this case, rG/D = 0 and the expression for the angular momentum about the center of mass
becomes

HG = IG ωk (4.25)

Another case of interest is when the point about which we take moments is a fixed center
of rotation, such as in Figure 4.7. Referring to the center of rotation as point O and using
a set of normal-tangential coordinates, we can write mrG/O = −mhen and vG = hωet , so
that

mrG/O × vG = mh2 ωk (4.26)

Applying the parallel axis theorem, the angular momentum about a fixed center of rotation
O can be expressed in terms of the mass moment of inertia about the fixed center of rotation
as

HO = mh2 ωk + IG ωk = IG + mh2 ωk = IO ωk

(4.27)
124 Applied Dynamics

Angular momentum of rigid bodies undergoing Rthree-dimensional motion is discussed in


Chapter 11, where it is shown that the expression body ρ × (ω
ω × ρ ) dm leads to the inertia
matrix.

4.4 Resultant Force and Moment


A useful method of analyzing forces and moments acting on a body is to look at their
combined effects. These combined effects are called resultant forces and resultant moments.
We first define what a force and a moment are. A force is generally defined as

Force: The action of one body on another.

It is interesting to note that no method exists to directly measure a force. Using laws gov-
erning deformations or accelerations caused by forces, the magnitude of a force is calculated
by measuring the effects of the force and by invoking the cause-effect relationships.
One of the most important steps in analyzing the forces and moments that act on a
system is to draw a free-body diagram. A free-body diagram leads to a better visualization
of the forces and moments, makes it easier to write force and moment balances, and prevents
mistakes. The following steps should be taken when drawing a free-body diagram:
• Isolate the bodies involved and determine the number of degrees of freedom.
• Select a coordinate system and positive directions, as well as motion variables.
• Draw the isolated bodies and mark their centers of mass.
• Draw all forces and moments that act on the isolated bodies and clearly label them.

P
F
rP/D
rP/G
G
D

FIGURE 4.8
Force acting through a point P .

A force acting on an unrestrained rigid body has two effects: it causes translational
motion and, unless the force acts through the center of mass, it causes a rotational motion.
Figure 4.8 shows a force F acting through point P . The moment of the force F about a
point D and about the center of mass G are

MD = rP/D × F MG = rP/G × F (4.28)

Consider next a rigid body that is acted upon by N external forces Fi (i = 1, 2, . . . , N )


and N 0 external moments Mj (j = 1, 2, . . . , N 0 ), as shown in Figure 4.9a. The forces go
Kinetics Fundamentals 125

through points Pi (i = 1, 2, . . . , N ) and the vector ri connects the center of mass and point
Pi . The locations at which the external moments are applied do not affect the results, as
the body is considered to be rigid for the purpose of this discussion.1

a) Fi b) c)
Mi
Pi

r1 G ri MG G
= = M D rG/D
rG/D F
F1 P1 r2 rN D G
P2 PN D F
MN F2 FN

FIGURE 4.9
Forces, moments, and resultants: a) forces and moments, b) resultant force and moment, c)
resultants about D.

The resultant force is defined as the sum of all external forces as


N
X
F = Fi (4.29)
i=1

The resultant force acts through the center of mass. The resultant moment about the center
of mass is
0
N
X N
X
MG = Mj + r i × Fi (4.30)
j=1 i=1

The resultant force and moment are shown in Figure 4.9b. Sometimes it is of interest
to express the net effect of all forces and moments as a resultant force applied through
a certain point other than the center of mass and a resultant moment about that point.
Consider point D in Figure 4.9c. The resultant moment about D is obtained by summing
moments of the resultant force and moment, which gives

MD = MG + rG/D × F (4.31)

Note that all three parts of Figure 4.9 are equivalent.

4.5 Laws of Motion


The translational and rotational laws of motion were stated by Isaac Newton in 1687 and
Leonhard Euler in 1750, respectively. The general nature of these laws was probably known
earlier. In Newton’s case there was a dispute between him and Robert Hooke regarding who
developed the laws of motion. Both worked on the subject separately, but Newton presented
and published the results sooner.
The laws of motion are stated here for translational as well as rotational motions.
1 For flexible bodies this is not the case; the location of the applied moments makes a difference.
126 Applied Dynamics

4.5.1 First Law


Translational Motion: The linear momentum of a body remains unchanged if the resul-
tant of all forces acting on the body is zero. A body that is at rest remains at rest and a
body in motion retains its linear momentum unless acted upon by forces whose resultant is
not zero. Also known as Newton’s First Law, this law is one form of the statement of the
principle of conservation of linear momentum.
Rotational Motion: The angular momentum of a body about its center of mass remains
unchanged if the resultant moment about the center of mass is zero. This law is one form
of the statement of the principle of conservation of angular momentum.

4.5.2 Second Law


Translational Motion: The rate of change of the linear momentum of a body is equal to
the resultant of all forces acting on it. This law is known as Newton’s Second Law, and can
be expressed mathematically as
d
p = F (4.32)
dt
in which F is the resultant force. For a rigid body, p = mvG and, when the mass remains
constant, Newton’s Second Law becomes

maG = F (4.33)

For a particle the subscript G is dropped and Newton’s Second Law becomes F = ma.
We can use any coordinate system, including those discussed in Chapter 2, when applying
Newton’s Second Law.
Rotational Motion: The rate of change of angular momentum about the center of mass
is equal to the resultant moment about the center of mass. Mathematically,

ḢG = MG (4.34)

As in the first law, the above expression is valid for both two- or three-dimensional motion.
For plane motion, the angular momentum is HG = IG ω and when the mass moment of
inertia is constant the rotational counterpart of the second law becomes

IG ω̇ = MG (4.35)

The relationship between acting forces and moments, their resultants, and the linear
and angular accelerations is shown in Figure 4.10.

MG HG G
= = M D rG/D
rG/D F rG/D ma G
D G D G D ma G

FIGURE 4.10
Forces, resultants, and accelerations.
Kinetics Fundamentals 127

Equation (4.34) is valid for the angular momentum balance about the center of mass.
We can extend it to the angular momentum balance about an arbitrary point, say, D. For
the case of plane motion, summing moments about D and considering Figure 4.9c gives

MD = rG/D × F + MG = IG ω̇k + rG/D × maG (4.36)

In terms of individual forces and moments, the moment balance about point D becomes
X X
MD = r∗i × Fi + Mi = IG ω̇k + rG/D × maG (4.37)

in which r∗i is the vector connecting point D with the point to which force Fi is applied.
A special case involves summing of moments about a fixed center of rotation when
such a point exists. Going back to Figure 4.7, where point O is the center of rotation,
mrG/O = −mhen , and aG = hω̇et + hω 2 en , the moment about point O becomes

MO = IG ω̇k + mh2 ω̇k = IG + mh2 ω̇k = IO ω̇k



(4.38)

In scalar form, MO = IO ω̇. For general three-dimensional motion about a fixed point O,
MO = ḢO . For rotation about a fixed point, Equation (4.38) is preferable as it eliminates
the need to calculate the reaction forces at O. Of course, in an actual design problem, we
need to calculate the reaction forces. It should be reiterated that that Equation (4.38) is
only valid when there is a fixed center of rotation.

!
#
m
"
D

FIGURE 4.11
Force acting on a particle.

For a particle, such as the one shown in Figure 4.11, the angular momentum about D
is HD = r × mv and when D is fixed, the angular impulse momentum relationship is

ḢD = MD (4.39)

in which MD = r × F is the resultant applied moment.


The second law relates translational and rotational accelerations to the acting forces
and moments. Once these relationships are obtained for a given problem, we can then carry
out a number of operations:
• We can conduct an instantaneous analysis and calculate accelerations from the acting
forces and vice versa.
• We can manipulate these relationships, if necessary, to obtain equations of motion, which
are in the form of ordinary differential equations and can be integrated analytically or
numerically to find the response. This subject is discussed in Sections 5.6–5.8.
• We can integrate these relationships over a time interval, which leads to the impulse-
momentum relationships, or over a displacement, which leads to work-energy relation-
ships. Sections 5.4 and 5.5 discuss this process.
128 Applied Dynamics

• We can conduct qualitative analysis to extract information from the relationships at


hand, by means of motion integrals. This process is discussed in more detail in Chapters
8 and 11.

4.5.3 Third Law


The translational part of the third law of motion basically states that for every action
there is an equal and opposite reaction. This law is known as Newton’s Third Law. When
two bodies exert forces on each other, these forces are equal in magnitude and opposite in
direction, and they act along a line going through the contact point between the two bodies.
If the bodies are not in contact, as in celestial mechanics problems, the forces go through
the line joining the centers of mass of the two bodies.
The rotational counterpart of the third law states that when two bodies exert moments
on each other, these moments are equal in magnitude, are opposite in direction, and act
about an axis that goes through the contact point between the bodies.

Example 4.3—Helicopter Blades

a) b)
Tail rotor Top view
of rotor
F
T
Rotor

FIGURE 4.12
a) Moment acting on helicopter due to rotor, b) top view of rotor.

A helicopter, such as the one shown in Figure 4.12, is an illustration of the application
of the third law. A helicopter gets its lift from the rotation of its blades. The engine of
the helicopter exerts a rotational moment on the blades, which makes the blades turn. The
third law indicates that the blades exert an equal and opposite moment on the body of the
helicopter, causing the helicopter to rotate in the opposite direction of the blades.
In the absence of other moments acting on the helicopter, the helicopter would rotate in
the opposite direction of the rotor blades. This rotation of the helicopter body is prevented
by the tail rotor. Shown in Figure 4.12a, the tail rotor generates an aerodynamic force
F that creates a moment M = F d that counterbalances the moment T generated by the
blades. As a result, the body of the helicopter does not rotate.
Kinetics Fundamentals 129

4.5.4 Inertia Forces and Inertia Moments


The translational and rotational laws of motion, as given in Equations (4.33) and (4.34),
can also be written by taking the acceleration terms to the same side as the force terms

F − maG = 0 MG − ḢG = 0 (4.40)

The above forms of the force and moment balances are known as dynamic equilibrium
equations, where the negatives of the accelerations are treated as inertia forces (−maG )
and inertia moments (−ḢG ). Treating the accelerations as inertia forces results in a system
that is in static equilibrium.
The above equations, especially the translational one, are sometimes referred to as
D’Alembert’s principle or D’Alembert’s equations, named after the French scientist and
mathematician Jean-Baptiste le Rond D’Alembert (1717–1783). There is debate in the lit-
erature as to what D’Alembert referred to as his principle. We will not refer to Equation
(4.40) as D’Alembert’s principle and define the D’Alembert’s principle in Chapter 8, when
discussing analytical mechanics.

Example 4.4
Consider the bicycle model of a vehicle from Figure 3.9. A tractive force FT acts on the
vehicle as the vehicle takes a turn with constant steer angle δ. Calculate the magnitudes of
the lateral forces acting on the vehicle.



 


 



 
 
  

 

FIGURE 4.13
Free-body diagram of bicycle model (top view).

The free-body diagram is shown in Figure 4.13. The three forces acting on the vehicle
are the tractive force FT and the lateral forces FLr and FLf . The xyz coordinate system
moves with the vehicle. The sum of forces gives
X
F = maG = FT i + FLr j + FLf j0 [a]

Summing moments about the center of mass leads to


X
MG = IG θ̈ = −FLr c + FLf b cos δ [b]
130 Applied Dynamics

Equations [a] and [b] represent three equations. The problem has five unknowns: aG
represents two unknowns, and the angular acceleration θ̈ and the two lateral forces are the
other three. We need two additional relations to solve this problem. These relations come
from the roll without slip constraints at the wheels.
Let us use vA and θ̇ as the velocity variables. The velocity of A is vA = vA i. The angular
velocity of the vehicle body is ω = θ̇k. It follows that the acceleration of point A becomes

aA = (v̇A )rel + ω × vA = v̇A i + θ̇k × vA i = v̇A i + θ̇vA j [c]

Using the relative acceleration equation and noting that α = θ̈k and rG/A = ci, we
obtain the acceleration of the center of mass as
   
aG = aA + α × rG/A + ω × ω × rG/A = v̇A − cθ̇2 i + vA θ̇ + cθ̈ j [d]

The second constraint arises from the restriction on point B, whose velocity lies along
the x0 axis. The velocity of B is

vB = vA + ω × rB/A = vA i + θ̇k × Li = vA i + Lθ̇j [e]

Noting that j0 = cos δj − sin δi, the velocity constraint leads to


vA tan δ
v B · j0 = 0 =⇒ θ̇ = [f ]
L
which provides a relationship between vA and θ̇. It follows that the acceleration of the center
of mass and the angular acceleration can be expressed in terms of a single velocity variable.
This variable and the two lateral forces are the three unknowns of the problem.
Solving the force and moment balances for general values of the steer angle is alge-
braically complex. Significant simplification can be realized by using a small angles assump-
tion for the steer angle δ: sin δ ≈ δ, cos δ ≈ 1. Also, higher-order terms in δ, such as δ 2 ,
can be neglected. Incorporation of these approximations leads to the following force and
moment balances
In x direction: mv̇A = FT − FLf δ

m 2 
In y direction: vA δ + cv̇A = FLf + FLr
L
v̇A δ
About z axis: IG θ̈ = IG = FLf b − FLr c [g]
L
which can be solved for the unknowns v̇A , FLf , and FLr . Note from the first equation that
the lateral force on the front axle has a component that acts against the direction of travel,
which slows the vehicle. This is part of the reason why a vehicle slows when taking a turn.
The other reason, which is due to tire slip, will be discussed in Chapter 14.
Consider the special case of when the vehicle is moving with constant velocity (v̇A = 0).
The tractive force FT acts to maintain speed. It follows that the angular acceleration of the
vehicle is also zero. We use the last two expressions in Equation [g] to solve for the lateral
forces in terms of the lateral acceleration as
2 2
vA δ c c vA δ b b
FLf = m = may FLr = m = may [h]
L L L L L L
2
where ay = mvA δ/L is the lateral acceleration of point A. We can show that the angular
acceleration of the center of mass G is very close to the lateral acceleration of point A. This
result will be used in Chapter 12 when calculating wheel loads and weight shift.
Kinetics Fundamentals 131

Next, let us consider some numbers: wheelbase 2.70 m, center of mass 1.50 m from rear
axle, mass 2,000 kg. The steer angle is constant at 1.5◦ , and the vehicle is traveling with
constant speed vA = 90 kph (25 m/s). The lateral acceleration becomes
2
vA δ 252 1.5
ay = = = 6.060 m/s2 [i]
L 2.7 57.296
which is around 0.6g. The lateral force on the front axle becomes
c 1.5
FLf = may = 2000 × 6.06 = 6734 N [j]
L 2.7
The component of the front lateral force along the line of motion is the tractive force
needed to maintain speed
1.5
FT = FLf δ = 6734 = 117.5 N [k]
57.296
Dividing this force by the mass gives a deceleration of d = 117.5/2000 = 0.0587 m/s2
(≈ 0.006g). While not a very large value, the effect accumulates if the vehicle takes a turn
for a long time, for example, in an exit ramp.

4.6 Forces and Moments Acting on Bodies


We next analyze the different types of forces that act on bodies. Broadly speaking, forces
can be categorized into two types: external or internal. External forces are generated from
outside the body or collection of bodies that make up the dynamical system. Internal forces
are generated among the components of a dynamical system.
Forces are transmitted from one part of the body to another through internal forces.
In a rigid body, magnitudes of the internal forces and moments are unknown,2 but the net
effect of the internal forces and moments can be calculated along a cross-section. In an
elastic body, the internal forces and moments result in stresses across the cross-section.
We can further classify forces into three categories:
1. Body Forces: These forces act over the entire body without necessarily making contact
with the body. Included in this category are gravitational and electromagnetic forces.
2. Contact Forces: The most frequent type of forces that act on a body are due to contact
with another body. Reaction forces, normal forces, resistive forces, impact forces, and
friction forces are contact forces. When the body is in contact with a fluid such as air
or water, the fluid exerts hydrodynamic or aerodynamic forces on the body.
3. Component Forces: These forces develop when a motion-transmitting or motion-
generating component is used to connect two bodies. The most common type of such
components are joints. Joints permit motion to be transmitted in a specific direction
and they create reaction forces and moments in the directions that they prevent motion.
Springs and dampers, which are discussed later in this chapter, generate forces and
moments between the components they connect. A motor is a power source and adds
energy to the components to which it is connected.
2 This mathematical oddity arises because of the idealization of a rigid body. In reality, all bodies have

some amount of compliance.


132 Applied Dynamics

4.7 Force of Gravity

m2 m2
r r F21

F12
m1 m1

FIGURE 4.14
Gravitational force between two masses.

Newton’s law of gravitation states that two bodies exert equal and opposite attractive
gravitational forces on each other. The magnitude of the forces is inversely proportional to
the square of the distance between the bodies. Considering the two point masses in Figure
4.14, Newton’s law of gravitation states
Gm1 m2
F12 = F21 = (4.41)
r2
in which G is the universal constant of gravitation, whose value is G = 6.673 × 10−11 m3 /
kg·s2 in SI units and G = 3.439 × 10−8 ft4 /lb·sec4 in U.S. Customary units.
Because G is such a small quantity, the gravitational force becomes significant only when
at least one of the bodies involved is very large, such as in the vicinity of a celestial body.
Equation (4.41) can be expressed in vector form as

Gm1 m2
F12 = −F21 = r (4.42)
r3
where r is the position vector between the centers of mass of the two bodies.

F = m2 g
z m
2

re

Earth
me

FIGURE 4.15
Gravitational force in the vicinity of the Earth.

For motion near the Earth, using the values for the mass of the Earth as m1 = me =
5.976 × 1024 kg and mean radius as re ≈ 6, 371 km and assuming that the distance of the
body from the surface of the Earth (z in Figure 4.15) is negligible compared to the radius
of the Earth, we can define the gravitational constant as g = Gme /re2 and obtain the force
Kinetics Fundamentals 133

of gravity as F = m2 g. The mean value of the gravitational constant is

6.673 × 10−11 × 5.976 × 1024 2


g = 2 = 9.824 m/s (4.43)
[6.371 × 106 ]
Using the above equation to calculate the gravitational constant g is not accurate. Equa-
tion (4.43) is based on treating the Earth as a particle (or rigid uniform sphere), and it
ignores the actual shape of the Earth,3 centrifugal effects due to the rotation of the Earth,
and gravitational effects of the sun and moon. Moreover, the density of the Earth is not
uniform. These effects result in different values for g at different points on the Earth.
A more accurate approximation for the gravitational constant g is the International
Gravity formula. This formula treats the Earth is a rigid ellipsoid and takes into consider-
ation the Earth’s rotation. The approximation for g near the surface of the Earth is given
as a function of the latitude and altitude as
2
g = 9.780327 1 + 0.005279 sin2 λ + 0.00023 sin4 λ − 3.086 × 10−6 h m/s
 
(4.44)

where λ is the latitude and h is the altitude in meters. The lowest value for g occurs at the
Equator because there the centripetal acceleration due to the rotation of the Earth is the
largest.4
For dynamics problems that do not require substantial precision, average values of g =
9.81 (or 9.807) m/s2 or g = 32.2 (or 32.17) ft/sec2 are used at sea level. These values
correspond the value of g from Equation (4.44) calculated at a latitude of 45◦ .

4.8 Contact and Reaction Forces


Reaction forces arise from enforcement of constraints on a body and from the contact of
one body with another. As we learned in Chapters 2 and 3, a constraint is a kinematical
relationship that restricts motion. What enforces a constraint is the constraint force. A
reaction force can be viewed as a constraint force. For example, a chair exerts a force on
the floor on which it is resting in the amount of its weight, and the floor exerts a reaction
force that is equal in value. The reaction force is usually called normal force.
A pin joint that permits rotation about an axis creates reaction forces in three directions
and reaction moments in the two directions orthogonal to the axis of rotation, as shown in
Figure 4.16. A pin joint in plane motion generates two constraint forces.
As discussed in Chapter 3, when two bodies come in contact with each other, the contact-
ing points define a common normal and a common tangent. For three-dimensional motion,
the common tangent becomes a surface that is tangent to both bodies. The associated
unit vectors are denoted by n and t. Note that, unlike normal and tangential coordinates,
positive directions of the unit vectors along the common normal and common tangent are
arbitrary. The reaction force is along the common normal, hence, the name normal force.
Figure 4.17 illustrates various forms of contact.
Another commonly encountered constraint force is encountered in bodies that are con-
nected by cables, such as in Figure 4.18a and the pendulum in Figure 5.18b. In Figure 4.18a,
the cable, when taut, exerts a tensile force T in the direction of the cable, as shown in the
3 The actual shape of the Earth is an oblate spheroid. The Earth looks more like an apple, with the

radius larger around the equator, smaller at the poles, and the poles slightly pressed in.
4 This is the reason why launch locations for spacecraft are selected as close to the Equator as possible,

in order to minimize the power and energy required for launch.


134 Applied Dynamics

a2


M2
F2 a1

M1
F3
a3 F1

FIGURE 4.16
A pin joint and its free-body diagram.

n n
n

n
F
n n
F t t t
t t
t F F

FIGURE 4.17
Various forms of contact.

free-body diagram in Figure 4.18b. The system in Figure 4.18 has one degree of freedom
when the cable is taut, in which case mass m1 moves the same amount as m2 .

4.9 Dry Friction Forces


Friction forces are developed when a body in contact with another moving or fixed body
slides, or has a tendency to slide, over the other body. Because the contact takes place along
the common normal, friction forces are generated along the common tangent or along the
tangent plane.
The amount of slippage between two bodies depends on the material properties of the
contacting bodies, characteristics of the surfaces (rough, smooth, etc.) of the contacting
bodies, amount and type of fluid (oil, water, etc.) between the contacting points, relative
speed of the contacting bodies, as well as temperature.
The study of friction is complex and an accurate representation of friction forces is diffi-
cult. A widely used model is known as dry friction or Coulomb friction. This model describes
the friction force by two coefficients of friction: µs , the coefficient of static friction, and µk ,
the coefficient of kinetic friction. Engineering handbooks contain tables of coefficients of
friction for a variety of materials and surfaces.
The friction force between two sliding bodies opposes the relative velocity of sliding,
regardless of what other forces are acting on the body. Figure 4.19 shows a few examples.
In the presence of sliding, the magnitude of the friction force is Ff = µk N . The friction
Kinetics Fundamentals 135

"# $#
k kx T
m! m!

N
x

T
m% m%

m% g

FIGURE 4.18
a) Two bodies connected by a cable, b) free-body diagrams.

g
v

Ff v
v
Ff
Ff

FIGURE 4.19
Friction forces for dynamic case (v > 0).

force can be written in vector form as


v
Ff = −µk N or Ff = −µk N et (4.45)
|v|
Recall that v = vet , where where v is the speed. For rectilinear motion Ff = −µk N sign (v),
where the signum function is defined as

sign (v) = 1 when v > 0 sign (v) = −1 when v < 0 (4.46)

For static problems, as shown in Figure 4.20, there is no sliding and the friction force
opposes impending velocity (motion that would result in the absence of the friction force).
In Figure 4.20a the block would slide down in the absence of friction, so the friction force
acts upwards. In Figure 4.20b the block would move to the right in the absence of friction
so the friction force is to the left.
For static problems, the magnitude of the friction force Ff varies between zero and µs N ,
where N is the normal force. The no-sliding condition is maintained while the friction force
is less than the maximum available friction force, that is, Ff < µs N . When the amount of
the friction force needed to prevent sliding exceeds the maximum available friction force, the
body begins to slide (Figure 4.19b) and the value of the friction force becomes Ff = µk N .
The kinetic coefficient of friction is smaller than or equal to its static counterpart, or µk ≤ µs .
In problems involving friction, it is necessary to determine whether there is sliding or
not. To determine whether there is sliding or not, we can begin by assuming that there is
no sliding and check the validity of this assumption by comparing the friction force required
to prevent sliding with the maximum available friction force, µs N . Alternatively, we can
136 Applied Dynamics

a) b)
g g
W W
F (external force)
m
N
Ff Ff N

FIGURE 4.20
Friction forces for static cases: a) block on an incline, b) block on level plane acted upon by
external force F .

assume sliding and calculate the acceleration. Validity of the sliding assumption can be
checked by relating the magnitude of the friction force to the acceleration.
Problems involving friction are an illustration of the cause-and-effect principle. We can-
not know, before solving the problem, the magnitudes of both the friction force and the accel-
eration it causes. If the body is sliding, magnitude of the friction force is known (Ff = µk N ),
but the acceleration is not known until the problem is solved. If the body is not sliding, the
acceleration is known (a = 0), but the magnitude of the friction force (0 ≤ Ff ≤ µs N ) is
not known until the problem is solved. Table 4.3 summarizes the cause-and-effect principle
for sliding friction problems.

TABLE 4.3
Known and unknown quantities for friction problems

Type Known Quantity Unknown Quantity


Sliding Friction force Ff = µk N Acceleration
No sliding Acceleration (= 0) Friction force Ff (0 ≤ Ff ≤ µs N )

This section is concerned with sliding friction. The effect of frictional forces during
rolling will be discussed in Section 5.2. Another form of friction that is encountered is
internal friction, which is due to hysteresis in the stress-strain curve, and will be studied in
Section 5.5.

Example 4.5
Consider the block on an incline in Figure 4.21. The block is initially at rest. Calculate the
acceleration of the block as a function of the propulsive force F .
Summing forces perpendicular to the incline, the normal force is N = mg cos θ. Consider
the case when the propulsive force is sufficient to push the vehicle upward. The free-body
diagram is given in Figure 4.21a, where the friction force acts down the incline. Summing
forces along the incline in Figure 4.21a gives
X
% F =⇒ F − mg sin θ − Ff = ma [a]

The limiting case is when a = 0 and Ff = µs mg cos θ. Introduction of this value into the
Kinetics Fundamentals 137
!" !t #"

v F
m m
F Ff v Ff
N N
mg  mg 
 
!t

FIGURE 4.21
Block on an incline: a) sliding up, b) sliding down.

above equation and solving for F gives

F = mg sin θ + µs mg cos θ [b]

If the propulsive force F > mg sin θ + µs mg cos θ, the block will move upward. Otherwise,
the block will not move upward.
Next, consider the downward motion of the block. The free-body diagram is given in
Figure 4.21b. For this to happen, friction, in the absence of external forces, is not sufficient
to prevent slip. The limiting case is when a = 0, F = 0, and Ff = µs mg cos θ. The force
balance along the incline gives

mg sin θ − µs mg cos θ = 0 [c]

We can solve for the critical value of the incline angle after which sliding down begins (or
the critical value of the friction coefficient to prevent sliding) as

µs = tan θ [d]

Next, consider the case when mg sin θ − µs mg cos θ > 0, in which case the block would
slide down on its own, and apply the external force F to prevent downward motion. Summing
of forces along the incline gives
X
. F =⇒ −F + mg sin θ − Ff = ma [e]

As mentioned previously, the limiting case is when a = 0 and Ff = µs mg cos θ. Intro-


ducing this condition into Equation [c] and solving for F , the following result is obtained:
the block will slide downwards when F < mg sin θ − µs mg cos θ. The motion of the block is
summarized in Table 4.4:

TABLE 4.4
Sliding or no sliding results

Condition Ensuing Motion


F < mg sin θ − µs mg cos θ Block will move down
mg sin θ − µs mg cos θ ≤ F ≤ mg sin θ + µs mg cos θ Block will not move
F > mg sin θ + µs mg cos θ Block will move up
138 Applied Dynamics

4.10 Aerodynamic Forces


Except when in a vacuum or in outer space, objects in motion continuously come into contact
with fluids, such as air and water. This section presents a primer on aerodynamic forces
and hydrodynamic effects. It should be noted that, considering the tremendous advances
in the field, we cannot do justice to the subject of aerodynamics in a few pages. The
interested reader is referred to several excellent texts on aerodynamics. Aerodynamic and
hydrodynamic forces affect the stability and control of vehicles.
Contact of gas particles with each other leads to two types of forces:
• Collision forces, which are normal forces, generate lift or downforce;
• Sliding forces, which are tangential and thus frictional, generate drag.
These forces are illustrated in Figure 4.22. The magnitudes of these forces are propor-
tional to the square of the relative air velocity, that is, the difference between the object’s
speed and the air speed.

&) *)

!"##"$ !"##"$
%&$'($% %&$'($%

FIGURE 4.22
a) Collision and b) sliding of air particles.

A vehicle moving with the speed of, say, vplane = 40 ft/sec, and subjected to a headwind
(wind coming at the vehicle) of vwind = 15 ft/sec can be treated as a vehicle moving with
a speed of vrel = vplane + vwind = 55 ft/sec in still air, as shown in Figure 4.23a, or as
a stationary vehicle subjected to an airflow of 55 ft/sec. By contrast, a tailwind (wind
coming from the rear of the vehicle) of 15 ft/sec decreases the relative speed of the vehicle
to vrel = vplane − vwind = 25 ft/sec, as shown in Figure 4.23b.
It is common to model aerodynamic forces by considering stationary objects immersed
in an airstream. The speed of the airstream is referred to as the free stream velocity and is
denoted by V∞ .
In aerodynamics (except hypersonic), we assume that the fluid involved obeys the ideal
gas law and it is incompressible. We describe the properties of the fluid by its density ρ
(or specific weight γ) and its viscosity µ. Viscosity is a measure of how sticky a fluid is,
or its resistance to an object moving through it. Viscous forces are related to the speed of
the moving object. Kinematic viscosity is defined by absolute viscosity divided by density.
Table 4.5 gives the properties of air at standard temperature and pressure. The density of
air changes proportionally with pressure and inversely proportionally with temperature.
The airfoil is the most commonly used shape for analyzing aerodynamic forces. The
wings of airplanes have the shape of airfoils. Figure 4.24 gives the nomenclature of an
airfoil. The flowing air meets the airfoil at the leading edge and loses contact with the
Kinetics Fundamentals 139

a) b)
vwind
vwind

vplane vplane

vrel = vplane + vwind vrel = vplane – vwind

FIGURE 4.23
Relative speed of an airplane: a) headwind, b) tailwind.

TABLE 4.5
Properties of dry air at standard atmosphere (p = 14.7 lb/in.2 ), at temperature of 59◦ F
(15◦ Celsius), and at sea level

Property Symbol U.S. Units SI Units


Specific weight γ 0.07651 lb/ft3 Not used
Density ρ 0.002378 slug/ft2 1.225 kg/m3
Absolute viscosity µ 3.74 × 10−7 slug/ft·sec 1.79 × 10−5 N·s/m2
Kinematic viscosity ν 1.57 × 10−4 ft2 /sec 1.46 × 10−5 m2 /s

airfoil at the trailing edge. The straight line joining the leading and trailing edges is called
the chord and its length is denoted by c. The angle that the chord makes with the direction
of the free stream velocity is the angle of attack α. The quarter chord point is at a distance
c/4 from the leading edge.

Angle of attack ( )
Thickness
V
Mean camber
8

Chord (c) line

c/4
Leading Camber Trailing
edge edge
Quarter chord point

FIGURE 4.24
Airfoil nomenclature.

A major characteristic of an airfoil is the difference in shape between its upper and
lower surfaces. The distance between the upper and lower surfaces, measured perpendicular
to the chord at any point on the chord, gives the thickness of the airfoil. The loci of the
mid-points of the thicknesses define the mean camber line. The distance between the mean
camber line and the chord is the camber. Camber is a very important property, as it affects
140 Applied Dynamics

the lift characteristics. Thickness and camber are usually defined as a percentage, obtained
by dividing the thickness or camber by the chord length.

4.10.1 Lift Force and Drag Force


The flow of air over a body results in two main effects that act on the surface of the body:
the pressure distribution, which acts perpendicular to the surface, and shear stress, which
acts tangent to the surface and generates drag. These forces are sketched in Figure 4.25
for an airfoil. The resultant aerodynamic force is obtained by integrating these forces over
the entire surface of the airfoil. Introducing the unit vectors n and t that are normal and
tangent to the surface, respectively, the resultant aerodynamic force can be written as the
surface integral
Z Z
FA = (−pn + τ t) ds (4.47)
s

in which p is the pressure, τ is the shear stress, and s is a surface integral.

n (n)
p(s) t (t)

V s
8

ds

FIGURE 4.25
Pressure and shear stress distribution.

The pressure distribution generates lift and drag forces. The shear stress generates a
drag force which is known as skin-friction drag. The resultant aerodynamic force depends
on the free stream velocity, the shape and size of the object immersed in the flow, the
characteristics of the surface of the object (e.g., smooth vs. rough), the angle of attack,
whether the body has any rotational motion,5 and the properties of the air, such as density
and viscosity. The component of the resultant aerodynamic force perpendicular to the free
stream velocity is called the lift force and is denoted by FL . The component along the free
stream velocity is denoted by FD and is called the drag force.
The net effect of forces acting on a body is characterized by the resultant force, which
goes through the center of mass, and a resultant moment about the center of mass. Equa-
tion (4.31) expresses the resultant force and moment about a point other than the center of
mass. There exists a point on the airfoil such that the resultant moment about this point
is zero and the net effect of the aerodynamic forces can be described as the lift and drag
forces going through this point. This point is referred to as the center of pressure.
While use of the center of pressure simplifies the force and moment balance equations,
the center of pressure is difficult to locate, as its position changes depending on the charac-
teristics of the flow. For airfoils it is customary to express the resultant of the aerodynamic
5 The rotational motion of the ball has a significant effect on the path followed by the ball.
Kinetics Fundamentals 141

FL

V Mc/4
FD
8

c/4

FIGURE 4.26
Resultant aerodynamic forces and moment about the quarter chord point.

effects as going through the quarter chord point, the point on the chord line that is a dis-
tance c/4 from the leading edge. The resultant moment acting through the quarter chord
point is denoted by Mc/4 . Figure 4.26 shows the resultant lift and drag forces and resultant
moment about the quarter chord point.

4.10.2 Aerodynamic Coefficients


We will make use of Bernoulli’s principle and dimensional analysis to express the lift and
drag forces. As discussed in Chapter 1, dimensional analysis, where a number of physical
parameters are bunched to create dimensionless parameters, is useful, particularly in fluid
mechanics. We can create scale models for testing and analyze an aerodynamic phenomenon
using equivalent mathematical models.
The dynamic pressure of the free stream (air coming at the body) is defined as
1 2
q∞ = ρ∞ V∞ (4.48)
2
in which ρ∞ is the density of the ambient air. The unit of dynamic pressure is energy/volume.
Bernoulli’s principle is an energy balance for incompressible flow that assumes no energy
loss and relates the static pressure and dynamic pressure of the free stream with that of a
point on the surface of the body
1 2 1
ρ∞ V ∞ + p∞ = ρV 2 + p (4.49)
2 2
The lift and drag forces and the aerodynamic moment can be expressed in terms of the
dynamic pressure and surface area A, also referred to as planform area, in contact with the
airflow, as

FL = CL q∞ A FD = CD q∞ A M = CM q∞ Ac∗ (4.50)

in which CL , CD , and CM are the lift coefficient, drag coefficient, and moment coefficient,
respectively. The area parameter A and chord parameter c∗ in the above equations can be
chosen by convenience. They basically are reference values. For an airfoil we usually select
c∗ as the chord length c.
The aerodynamic coefficients are functions of the shape of the body, the angle of attack
α, the Reynolds number Re, and Mach number M∞ associated with the free stream. The
Reynolds and Mach numbers are dimensionless quantities defined as
ρ∞ V∞ c V∞
Re = M∞ = (4.51)
µ∞ a∞
142 Applied Dynamics

in which µ∞ is the ambient viscosity of the free stream and a∞ is the speed of sound in the
free stream. The Reynolds number is the ratio of the inertial forces (ρV 2 /L) to the viscous
forces (µV /L2 ), where L is a length parameter, selected as the chord c for airfoils.
At low Reynolds numbers, viscous forces are dominant, resulting in smooth flow of the
fluid, known as laminar flow. For higher Reynolds numbers (> 5 × 105 for flow over a plate),
inertial forces dominate, resulting in vortices and eddies. This type of flow is referred to as
turbulent flow. The Mach number is the ratio of the fluid speed to the ambient speed of
sound. The dependence of the aerodynamic coefficients on the Reynolds and Mach numbers
becomes more significant at high speeds. At low speeds, we can treat the aerodynamic
coefficients (especially the lift and moment coefficients) as dependent primarily on the angle
of attack.
We can understand why airfoils are shaped the way they are by examining Figure 4.25
and Bernoulli’s equation. An airfoil is sloped more on the top than on the bottom, a property
we defined earlier as camber. The air on top of the airfoil travels longer to reach the trailing
edge. The airflow speeds up to accomplish this. Bernoulli’s equation indicates that the
higher airspeed over the airfoil results in a lower pressure distribution on the top of the
airfoil than on the bottom. This pressure difference, when integrated over the entire airfoil,
gives the lift force.
The pressure difference between the top and bottom of the wings of an airplane is
referred to as wing loading and is calculated as the airplane weight divided by the wing
area. For passenger airplanes, the wing loading is about 1 psi.
As an aircraft prepares to land (or as it takes off) the pilot extends the trailing edge flaps
and the leading edge slats, which makes the wings wider and more curved. The increase in
camber results in a higher lift force, which the aircraft needs during take-off as well as when
the pilot reduces speed during the landing approach. To maximize lift, airplanes taking off
and landing fly into the wind, increasing their relative speed.

CL max
CL
Slope

CD
CDmin

CM

Negative Zero
lift lift

FIGURE 4.27
Generic lift, drag, and moment coefficients as a function of the angle of attack α.

Figure 4.27 shows generic plots of the lift, drag, and moment coefficients for an airfoil
as functions of the angle of attack. Plots of the lift and moment coefficients usually have a
linear range. The maximum value of the lift coefficient is known as CLmax . At higher values
of the angle of attack, the lift coefficient becomes smaller due to flow separation. The drag
force usually varies nonlinearly. It is interesting to note that the minimum value of the drag
Kinetics Fundamentals 143

coefficient does not occur when the angle of attack is zero. An important design criterion
in aerodynamics is the maximization of the lift to drag ratio, FL /FD .
While increasing lift is a laudable goal for designers of flying vehicles, lift is undesired for
ground vehicles. Lift forces reduce the normal forces (wheel loads), thus reducing the friction
forces that can be generated for cornering, braking, acceleration, or avoiding slip. Also, the
possibility for the vehicle to lose contact with the road surface is increased. Sporty and race
cars are designed so that the shape of the vehicle minimizes the lift force. In sports cars a
frequently used component to reduce lift is a spoiler, which is shaped as an upside down
airfoil, that is, with negative camber. Ground vehicle aerodynamics is further discussed in
Chapter 13.

4.10.3 Flow Separation


At high values of the angle of attack, flow over the body begins to separate, usually at the
trailing edge. The net effect of this separation, which is a complex phenomenon to analyze
in itself, is a change in the pressure distribution, which results in reduced lift and increased
drag.
Consider an airfoil and the air flowing around it. At any point in the vicinity of the
airfoil the air will have a certain velocity. We can draw the tangents to the air velocity at
every point. The family of curves that are instantaneously tangent to the velocity vector of
the flow are known as streamlines. In general, we plot the streamlines of a body at different
distances from the body to get an idea of how the air is flowing around the body. Figure 4.28
depicts streamlines over an airfoil for different values of the angle of attack. What is known
as attached flow occurs at low angles of attack and for bodies whose shapes resemble airfoils.

Increasing

FIGURE 4.28
Streamlines for attached flow. Separation begins as the angle of attack α gets larger.

At higher angles of attack, the streamlines stop following the contour of the body, and
separated flow takes place. Depending on the bluntness of the body and flow conditions,
the flow separation can begin right after the leading edge, or near the trailing edge. The
amount of drag in a body is related to how early the flow separates from the body. Vortices
form past the separation point due to the pressure imbalance between the front and rear
of the separation point, as well as due to the viscosity of the fluid. All this has the effect
of reducing lift and increasing drag. Figure 4.29 depicts flow separation and resulting drag
coefficients for a variety of shapes.
Vortices that result from flow separation can be sources of instability. The phenomenon
of vortex shedding can cause unstable vibrations and resonance, as was observed with the
spectacular collapse of the Tacoma Narrows bridge in 1940.
Vortex shedding occurs in the range 60 < Re < 5000. Especially after Re > 1000,
vortices generated by the flow separation become regularly spaced. This regular spacing,
144 Applied Dynamics

Separation point
a)

Separation point
b)

c) Separation point

FIGURE 4.29
Vortex generation due to flow separation at Re = 105 . All bodies have the same length
parameter. a) Plate length d, CD = 2.0, b) cylinder diameter d, CD = 1.2, c) airfoil
thickness d, CD = 0.12.

observed first by Theodore von Karman in 1911, leads to a lift force that is harmonic. The
harmonic nature of the lift force is characterized by the Strouhal number (St), which is
defined as
fd
St = (4.52)
V
where f is the frequency of the vortex shedding and d is the body thickness. When the
Strouhal number has the value St ≈ 0.21, a harmonic lift force develops, which has the form
1
FL = CD ρAV 2 sin ωt (4.53)
2
where ω = 2πf . The lift force provides harmonic excitation, and when the excitation is
close to the natural frequency of the system, the lift force may cause resonance.6
Vortex shedding is commonly encountered in tall chimneys and transmission lines (for
electricity) that are covered with ice. In airfoils, vortex shedding results in flutter, which
may lead to unstable vibrations of the airfoil. Tall chimneys in windy areas have vibration
dampers or helical spoilers attached to them to disrupt the harmonic nature of the vortices
caused by wind. Vibrations caused by flow around a body, such as vortex shedding, or due
to fluid flow inside pipes, are referred to as self-induced vibrations.
The flow separation analysis above was for a single airfoil, or an idealized wingspan of
infinite length. A finite length wingspan leads to the formation of vortices at the edges of the
wing, as depicted in Figure 4.30. These vortices, which form in airplanes as well as ground
vehicles, lead to changes in the velocity and pressure fields. The resulting aerodynamic forces
have the effect of pushing the trailing edge down, a phenomenon known as downwash. In
addition, the change in the pressure distribution results in increased drag. This drag created
by downwash is known as induced drag.
Modern aircraft wings are fitted with winglets, also known as wing tip devices, which
reduce the vortices generated at the wing tips and thus reduce the drag forces. They also
have the added benefit of making the plane more stable.
6 Resonance is discussed in Chapter 6.
Kinetics Fundamentals 145

8
Wing

Vortices

FIGURE 4.30
Vortex formation at edges.

There are several other sources of drag in a body that moves in an airstream. In ground
vehicles, air flow into the engine, as well as flow under the vehicle and inside the wheel
wells, results in additional drag forces. Such forces are known as interference drag.
In general, it is harder to model drag than it is to model lift, especially for complex
shapes. Computational models are more accurate for predicting lift than they are for pre-
dicting drag. We note that there are several other forms of drag that are not discussed
here.

4.10.4 Drag Approximation for Very Low Reynolds Numbers


For bodies that move very slowly, as in a body moving in water, the Reynolds number is
very small
ρvR
Re = ≤ 1 (4.54)
µ
where R is a dimensional characteristic of a body, such as its length or radius. The drag
force is primarily due to friction (as there is little change in the pressure distribution) and
linearly proportional to velocity. For a sphere of radius R, the drag force can be modeled as

FD ≈ 6πµRV (4.55)

in which V is the speed of the body (or free stream velocity). The above equation is used
more frequently in hydrodynamics, rather than aerodynamics. Also, as will be discussed in
Section 4.12, the above equation provides a linear approximation of hydrodynamic drag and
viscous damping forces.

Example 4.6—Aerodynamics of Sport Balls


Shown in Figure 4.31a is a spherical ball launched with initial speed v0 and at an angle
ψ0 with the horizontal. The only forces acting on the ball are gravity and aerodynamic
resistance (drag). We will ignore rotational motion of the ball and use normal and tangential
coordinates to describe the motion.
We express the velocity as v = vet , where v is the speed of the ball. The angle the
velocity vector makes with the horizontal is denoted by ψ. The free-body diagram is shown
146 Applied Dynamics

!" t #" t

!  ! 
m
"D x


vo  mg
o

z n
n

FIGURE 4.31
a) Ball in air. b) Free-body diagram.

in Figure 4.31b. Summing forces and using Newton’s Second Law gives

v2
X  
F = ma =⇒ m v̇et + en = −mgk − FD et [a]
ρ
where FD is the drag force, expressed as
1
FD = ρ̃Av 2 CD [b]
2
in which ρ̃ is the density of air (note the change in notation: the radius of curvature ρ and
density ρ̃), A is the cross-sectional area A = πr2 , where r is the radius of the ball, and CD
is the drag coefficient.
Noting7 that v/ρ = −ψ̇, the force balances in the normal and tangential components
are
1
mv̇ = −mg sin ψ − ρ̃Av 2 CD − mv ψ̇ = mg cos ψ [c]
2
which can be rearranged as
g cos ψ
v̇ = −g sin ψ − kv 2 ψ̇ = − [d]
v
in which k = ρ̃AC2m . These forms of the equations are more useful than using rectilinear
D

coordinates because they provide more insight.


The next step is to nondimensionalize the equations. The angle ψ is already nondimen-
sional, so that scaling constants are needed for speed and time. For speed, we can select v0 ,
the usual launch speed of the ball. Also, v0 /g has units of time. As in Chapter 1, rewrite
Equation [d] with stars, to indicate that the quantities are dimensional
g cos ψ
v̇∗ = −g sin ψ − kv ∗2 ψ̇∗ = − [e]
v∗
and introduce the nondimensional variables and derivative operators
v∗ t∗ d g d
v = t = = [f ]
v0 v0 /g dt∗ v0 dt
7 To see this, think of the x and z axes as the tangential and normal directions and the center of

curvature as an instant center. The velocity can be written as v = ω × r, where ω = ψ̇j and r = −ρk, so
that ω × r = −ρψ̇i = vi.
Kinetics Fundamentals 147

It follows that v̇∗ can be nondimensionalized as


dv∗ v0 dv dv
v̇∗ = = = g [g]
dt∗ v0 /g dt dt

After dividing Equation [e] by g, the nondimensional equations become

cos ψ
v̇ = − sin ψ − v 2 ψ̇ = − [h]
v
where
kv02 ρ̃ACD v02
 = = [i]
g 2mg
is the initial drag force divided by the weight and is known as the drag-to-weight ratio. This
ratio is the most significant factor when analyzing the motion of a projectile in a resistive
medium. A larger drag to weight ratio implies that the drag force is more significant than
the weight and vice versa.
Table 4.6, taken from the book by De Mestre with some additions, gives a listing. It
turns out that the Reynolds number of the ball is in the range of Re = 1.0×105 to 2.0×105 ,
a transition region when the flow changes from laminar to turbulent. The drag coefficient
CD is 0.45 when the Reynolds number is less than 1.4 × 105 and 0.2 otherwise.

TABLE 4.6
Typical values of drag-to-weight ratio for spherical balls

ρ̃ACD v02
Ball Type Diameter (cm) Mass (gm) v0 (m/s) = 2mg

Baseball 7.27 142–150 40 0.53


Basketball 23.9–24.3 625 15 0.20
Golf 4.27 45.9 70 1.23
Ping Pong 4.00 2.7 25 8.80
Racquetball 5.70 39.8 40 2.75
Soccer 22.3 425 25 1.02
Squash 4.00 23–25 50 3.52
Tennis 6.54–6.86 60 40 1.00

4.11 Spring Forces


Many times, dynamical systems contain components that provide flexibility and/or dissipate
energy. For example, ground vehicles have suspension systems, which consist of springs and
dampers, that dissipate shocks and help the vehicle hold the road. Some of the components
that make up dynamical systems may have a certain amount of elasticity themselves. For
example, the wings of an aircraft have flexibility, as building a rigid wing would result in an
extremely heavy structure and also because a flexible wing can better distribute the effects
of the loads that act on the airplane.
This section studies springs, which store energy and add flexibility. The following section
discusses dampers, which dissipate energy.
148 Applied Dynamics

4.11.1 Modeling of Springs


Springs store potential energy by generating a restoring (resistive) force that always acts
opposite to the springs’ deformation. The spring force is always towards undeformed position
of the spring. Springs can be classified in two different ways. The first classification is
based on the type of deformation the springs oppose. Figure 4.32a shows an axial spring,
also known as an extensional spring, which opposes translational motion and generates a
restoring force Fs (x), for the two cases of extension and compression. Figure 4.32b shows
the corresponding free-body diagrams.
Shown in Figure 4.32c, a torsional spring opposes rotational motion, and it generates
a restoring moment. A spring can be manufactured from a coil or from elements with
compliance, such as beams. For example, the suspension systems of cars of the past and of
trucks use beam elements known as leaf springs.

a) b) c)
x
k F F
Fs = kx 
x Undeformed
x
k F F Ms= kT  position
Fs = kx
Undeformed
position kT

FIGURE 4.32
a) An axial (extensional) spring. b) Free-body diagram. c) A torsional spring.

The second classification of springs is based on the linearity of the restoring force and
moment. For many springs, the spring force and moment can be approximated as a linear
function of the spring deformation for a certain range of the deformation. The approximation
is known as the ideal spring assumption. A spring whose restoring force or moment is linearly
proportional to its deformation is known as a linear spring.8 We can write the magnitudes
of the spring forces in the linear range as

Fs (x) = kx Ms (θ) = kT θ (4.56)

in which k and kT are called axial spring constant or spring rate and torsional spring
constant, respectively. For an axial spring, the spring force can be expressed in vector form
as Fs = −kr, where the vector r denotes the deformation of the spring. Spring constants
describe the resistance of springs to deformation. The unit of an axial spring constant is
force/displacement and a torsional spring has units of moment/angle = moment.
A more accurate model is that of a nonlinear spring, where the restoring force or moment
is a nonlinear function. For axial springs, two types of nonlinear spring models are used.
One is a softening spring, where the resistive force of the spring gets smaller as the spring
deflection becomes larger. A commonly used model for a softening spring is

Fs (x) = kx 1 − x2

(4.57)
8 Some texts refer to an axial spring as linear.
Kinetics Fundamentals 149

where  denotes the softening property.


In some springs, the spring force increases more rapidly than the deformation. The spring
force is usually modeled as

Fs (x) = kx 1 + x2

(4.58)

Fs

Stiffening
spring
x

Linear
range

x
Softening
spring

FIGURE 4.33
Softening and stiffening springs.

Such a spring is known as a stiffening spring. Figure 4.33 plots the force-deformation curves.
The softening spring model is encountered far more frequently than the stiffening spring
model. The stiffening spring model is usually encountered under compressive loads, as the
compression level approaches solid height, that is, when the spring is fully compressed.
Springs are usually modeled as massless, as the mass of the spring is much less than the
mass of the body to which it is attached. Including the mass of the spring in a mathematical
model is discussed in vibrations texts.

4.11.2 Equivalent Spring Constants


Certain springs are designed by connecting more than one spring together or are built using
elastic members (such as beams or torsion bars) to generate restoring forces. The cumulative
effect of the combined springs and other elastic components is described by an equivalent
spring constant, or keq .
Two applications involve use of combinations of springs: connecting the springs in series
and in parallel. Consider first springs in parallel, as shown in Figure 4.34a. Upon loading
by force F , the springs deflect by x. When two springs are parallel, their deflections are the
same, so the total resistive force Fs shown in Figure 4.34b has the form

Fs = F = k1 x + k2 x = (k1 + k2 ) x = keq x (4.59)

The equivalent spring constant is keq = k1 + k2 .


Shown in Figure 4.35a are springs that are connected in series. The unstretched lengths
of the springs are L1 and L2 . Here, the resistive force generated by each spring is the same
and each spring deflects by a different amount. As shown in the free-body diagrams in Figure
150 Applied Dynamics

a) Undeformed Deformed b)

k1 k1 x

k2 F F

k2 x
x

FIGURE 4.34
a) Springs in parallel, b) free-body diagram.

a)

k1 k2 F

L1 L2 x

b)

k1 k2 Fs Fs F
Fs Fs

L 2 + x2
L1 + x1

FIGURE 4.35
a) Springs in series, b) free-body diagrams.

4.35b, the deflections are x1 and x2 . The total deflection of the springs is x = x1 + x2 . The
spring force Fs can be expressed as

Fs = k1 x1 = k2 x2 = keq x = keq (x1 + x2 ) (4.60)

The total spring deflection is


Fs Fs Fs
x = x1 + x2 = + = (4.61)
k1 k2 keq

which leads to the equivalent spring constant as


1 1 1 k1 k2
= + or keq = (4.62)
keq k1 k2 k1 + k2

When springs are connected in parallel, the equivalent spring is stronger than the indi-
vidual springs. When springs are connected in series, the equivalent spring is weaker than
Kinetics Fundamentals 151

both springs. This brings the question as to why anyone would design springs in series if
the resulting spring is weaker. Springs in series are used in some special applications, such
as in a box spring and mattress, or in cases where we initially want a small amount of
resistance to the applied force. Subsequently, after a certain level of deformation, a larger
amount of resistance is desired. The inner sole and outer sole of a shoe, for instance, serve
that purpose.

4.11.3 Stiffness Generating Components


The analysis of springs here has so far considered ideal springs. We next consider elastic
components that are used as springs, as stiffness-producing devices or as components that
generate restoring forces.

a) F b) F

d _
T=F D
2
D
F

F
D
FIGURE 4.36 d
A coil spring: a) geometry, b) free-body diagram.

A coil spring (Figure 4.36a) is the most commonly used type of spring and is manufac-
tured by curving a circular rod into a coil. Coil springs are torsion devices, as seen from the
free-body diagram in Figure 4.36b. Derivation of the spring constant of a coil spring can be
found in texts on mechanical design. The equivalent spring constant has the form
Gd4
keq = (4.63)
8N D3
in which N is the number of coils, D is the mean coil diameter, d is the diameter of the
rod (or wire) formed into the coil, and G is the shear modulus, or modulus of rigidity. The
torque generated by the applied force F is T = F D/2.
Having a spring with N coils is the equivalent of N single coil springs connected in series.
While the N coils reduce the equivalent spring constant, the energy-absorbing capacity of
the combined spring is increased. For a spring with one coil, denoting by h the maximum
compression of the single coil and by k the spring constant of the coil, the maximum energy
the one coil can absorb in compression is
1 2
E1 = kh (4.64)
2
By contrast, when there are N coils, the equivalent spring constant is keq = k/N ,
while the total deformation is N h. Hence, the maximum energy the spring can absorb in
compression is
1 k 2 1
EN = (N h) = N kh2 = N E1 (4.65)
2N 2
152 Applied Dynamics

so a spring with N coils can absorb N times more energy as a spring with one coil. In
addition, the spring with N coils can compress N times more than that of a single coil.

F F

FIGURE 4.37
A tension spring.

Springs are usually designed to operate only in tension or in compression. A tension


spring (Figure 4.37) has ends that are looped, to make it easier for the operator of the
spring to pull the spring out. An exercise chest pull is in this category. By contrast, as
shown in Figure 4.36a, a compression spring has ends that are squared and ground, in order
to better deal with compressive forces and to reduce the possibility of buckling.
Springs that are designed to operate in both tension and compression are designed as
compression springs, as in vehicle suspensions and ball point pens. In such systems, the
springs are installed with an initial compressional deflection, so that when the spring oper-
ates, it compresses or deflects around the initial deflection, without reaching solid height,
when the spring is fully compressed, and also without reaching the undeformed position.

a) b)
EI
L P L/2 2P
3
PL
_
3EI

_PL3
Deformed shape 96EI

FIGURE 4.38
Beams as springs (deflections not drawn to scale): a) fixed-free beam, b) fixed-fixed beam.

Elastic beams are also used (or modeled) as springs, such as in suspension systems of
trucks or in columns of buildings. For example, consider a fixed-free beam of length L,
shown in Figure 4.38a. The deflection at the tip of the beam for a load P applied at the tip
can be shown to be
P L3 P
δ = = (4.66)
3EI keq

where I is the area moment of inertia of a cross section and E is the modulus of elasticity,
so the equivalent spring constant for the tip deflection of a fixed-free beam is
P 3EI
keq = = (4.67)
δ L3
If the beam is made longer, the spring constant becomes smaller, and the deflection
becomes larger. Also, when both ends of the beam are fixed, the resulting equivalent spring
Kinetics Fundamentals 153

is much stronger, as evidenced by the much smaller deflection in Figure 4.38b. Leaf springs
used in trucks and older cars as suspension systems are a collection of thin beams, as shown
in Figure 4.39.

P P

2P

FIGURE 4.39
A leaf spring.

Figure 4.38b is the simplified schematic of the leaf spring in Figure 4.39 (the stiffness
EI is not uniform in a leaf spring). Using beams as equivalent springs assumes only static
deformations and that the weight of the beam is small compared to the load it supports.
The validity of both these assumptions needs to be checked before the expression for the
equivalent spring is used.
Another common use of beams (or columns) as spring elements is in the modeling of
buildings, as in Figure 4.40a. Here, the boundary conditions on the individual beams are
zero deflection and slope at one end and zero slope at the other, which is depicted in Figure
4.40b. Applying these boundary conditions, the equivalent spring constant for each column
can be shown to be
P 12EI
keq = = (4.68)
δ L3
so that the building in Figure 4.40 can be modeled as a two mass-spring system shown in
Figure 4.41, with
12E1 I1 12E2 I2
k1 = 4 k2 = 4 (4.69)
L31 L32

Rods to which twisting moments are applied can also be used as springs. Given a rod
of length L and radius r to which a twisting moment T is applied, the twist of the rod is
φ = T L/GJ, where G is the modulus of rigidity and J = πr4 /2 is the area polar moment
of inertia. It follows that the equivalent torsional spring constant is keq = GJ/L. Anti-roll
bars, which are widely used in vehicles to increase roll stiffness, as discussed in Section
15.20, are torsional springs.
Tension in wires is also used to generate restoring forces. Shown in Figure 4.42a is a wire
of length 2L and tension T to which a force F is applied in the middle. Summing forces in
the vertical direction in the free-body diagram in Figure 4.42b, we obtain

F = 2T sin θ (4.70)

Assuming for small angles sin θ ≈ θ and denoting the deflection by x = L sin θ ≈ Lθ, we
can write the above equation as
2T
F = 2T sin θ ≈ x = keq x (4.71)
L
from which we conclude that the equivalent spring constant is keq = 2T /L.
154 Applied Dynamics

a) b)
m2 
m
L2, E2, I2
L
EI
m1
L1, E1, I1

FIGURE 4.40
Modeling a building: a) the building as an assembly of beams, b) zero slope end conditions.

k1 k2
m1 m2

x1 x2

FIGURE 4.41
Equivalent model as a two mass-spring system.

a) Undeformed position b)

L L
x x
T T
Tension T
F F

FIGURE 4.42
a) A wire in tension, b) free-body diagram.
Kinetics Fundamentals 155

4.12 Dampers
Dampers (or dashpots) are components that dissipate energy. Every moving system loses
energy through one means or another. Earlier, we saw friction forces and aerodynamic
drag as resistive forces that dissipate energy. Another way to describe energy dissipation is
through a viscous damping model, by means of shock absorbers (dampers). The damping
force is linearly proportional to the velocity, making it possible to use linear theory to find
the response.

FIGURE 4.43
A damper (shock absorber).

A viscous damper can be used to model internal energy dissipation as well as to model
an actual shock absorber. A shock absorber consists of a cylinder with a viscous fluid inside
and a piston, which has holes in its bottom, as shown in Figure 4.43. As the piston moves,
the viscous fluid exerts a resistive force proportional to the velocity. In the simplest model,
the force generated by the damper is modeled as linearly proportional to the velocity of the
body, in the form
Fd = cẋ (4.72)
where Fd is the force of the damper and c is the damping coefficient, or coefficient of viscous
damping, damping constant. Figure 4.44 gives the schematic of a damper. The first use of
dampers in automotive applications was in 1902.
The general equation used for calculating the damping coefficient can be shown to be
3πD3 L
 
2d
c = µ 1 + (4.73)
4d3 D
in which µ is the viscosity of the fluid, D is the piston diameter, d is the distance between
the outer edge of the piston and the fluid container (or the diameters of the holes in the
piston), and L is the piston thickness. Equations (4.72) and (4.73) agree reasonably well
with shock absorber data. Shock absorbers used in the automotive industry actually have a
slightly more complex model, as they have different damping coefficients depending on the
direction of the motion of the piston velocity, as discussed in Section 15.8. An ideal damper
is modeled as massless.
For bodies moving slowly through a fluid, such as a boat or submarine, the drag force can
be modeled as being linearly proportional to the speed. For a sphere of radius R moving
with speed v, the drag force is linearly proportional to the speed, FD = cv, where from
Equation (4.55) the hydrodynamic damping constant c is
c = 6πµR (4.74)
156 Applied Dynamics

Cylinder

Piston L

Viscous fluid
D d
d

FIGURE 4.44
Schematic of a damper.

4.13 Bibliography
Anderson, J.D., Aircraft Performance and Design, McGraw-Hill, 1999.
Anderson, J.D., Fundamentals of Aerodynamics, 4th Edition, McGraw-Hill, 2005.
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Benaroya, H., and Nagurka, M., Mechanical Vibration, 3rd Edition, CRC Press, 2009.
Bottega, W.J., Engineering Vibrations, CRC Press, 2006.
De Mestre, N., The Mathematics of Projectiles in Sport, Cambridge University Press, 1990.
Greenwood, D., Principles of Dynamics, 2nd Edition, Prentice-Hall, 1988.
Mehta, R.D., “Aerodynamics of Sports Balls,” Annual Review of Fluid Mechanics, Vol. 17,
1985, pp. 151–189.
Stengel, R.F., Flight Dynamics, Princeton University Press, 2004.
Wilson, D.G. (with Papadopoulos, J.), Bicycling Science, 3rd Edition, MIT Press, 2004.

4.14 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 4.2—Rigid Body Geometry


4.1 (E) Consider the L-shaped body in Figure 4.45a of total mass M , shaped from uniform
wire of length 3L/2. Calculate the mass moment of inertia about an axis perpendicular to
the body and going through point O.
4.2 (M) Consider the T-shaped body in Figure 4.45b of total mass M and uniform density.
Calculate the location of the center of mass and mass moment of inertia about the center
of mass.
4.3 (M) Consider the disk with two triangular gaps in it, shown in Figure 4.46a, where
Kinetics Fundamentals 157

a) b) L/5
M
L

M
L/2

L/5
O L L

FIGURE 4.45
Figures for a) Problem 4.1, b) Problem 4.2.

a = 2R/9. Calculate the location of the center of mass and mass moment of inertia about
the center of mass.

a) b)
2a

2a
Ro
Ri
a O
2a O
a
R 2a

FIGURE 4.46
Figures for a) Problem 4.3, b) Problem 4.4.

4.4 (M) Calculate the center of mass and mass moment of inertia about O of the arc in
Figure 4.46b with inner radius Ri and outer radius Ro as a function of the angle θ. The
mass entity is ρ.
4.5 (M) Calculate the mass moment of inertia about O of the beam in Figure 8.33 to which
concentrated masses are added at the ends.
4.6 (M) The compound pendulum in Figure 4.47 consists of a large disk of mass m and
radius R that rotates about its center O, a rod of mass m and length 3R, and a smaller
disk of mass m/2 and radius R/2. Calculate the mass moment of inertia about O.

Section 4.3—Linear Momentum and Angular Momentum


4.7 (E) The body of mass m = 2 kg and mass moment of inertia IG = 3 kg·m2 in Figure
4.48 is moving so that the center of mass velocity is vG = 4i + 0.8j m/s and its angular
velocity is ω = −0.9k rad/s. Calculate the angular momentum of the body about point D.
4.8 (M) Consider the body in Examples 4.1 and 4.2. You are given that the angular velocity
158 Applied Dynamics

R 3R
O R/2
m
m/2
m

FIGURE 4.47
Figure for Problem 4.6.


vG x
G
0.4 m 1.8 m D

FIGURE 4.48
Figure for Problem 4.7.

of the body is 0.3 rad/s ccw and that point O is the instant center. Calculate the angular
momentum about points O and G.

Section 4.4—Resultant Force and Moment


4.9 (E) Consider the vehicle in Figure 12.21. Calculate the resultant force and resultant
moment about the center of mass G.
4.10 (M) Consider the vehicle in Figure 12.4. Calculate the resultant force and the resultant
moment about points A and B.

Section 4.5—Laws of Motion

v L tail
z
L wing

x G B
W 100"

Rx
Rz

100" 50" 200" 150"

FIGURE 4.49
Figure for Problem 4.11.
Kinetics Fundamentals 159

4.11 (M) The airplane in Figure 4.49 has a weight of W = 800, 000 lb and a centroidal mass
moment of inertia of IG = 4 × 107 lb·in·sec2 . As the aircraft lands, the rear wheels make
contact with the ground and the wing and tail lift forces are Lwing = 0.8W, Ltail = 0.04W .
The accelerations of the airplane are ax = −0.2g (to the right) and az = 0.4g (upwards).
Find the reactions at the rear wheel, as well as the pitch acceleration of the airplane. Then,
determine the forces acting on the support braces of a package weighing 50 lb, located at
point B. The pitch angular velocity of the airplane is zero.
4.12 (M) A plate of mass m, length L and width 6L/13 is hanging from a cable of length
L/2, as shown in Figure 4.50a. The rod is subjected to a force F at point D. Calculate the
accelerations of points D and E on the rod. Note: This is a two-degrees-of-freedom problem.

a) b)
g
L/2
O
E z z

6L/13 x d x
L
G L
F
L/2
F D

FIGURE 4.50
Figures for a) Problem 4.12, b) Problem 4.14.

4.13 (M) A vehicle of mass 1,500 kg is traveling at a speed of 90 kph. The effect of wind
resistance on the vehicle is F = −kv 2 . Find the value of k if the wind slows the vehicle to
a speed of 45 kph over a distance of 400 m. How much time did it take for the speed to be
reduced to half its original value?
4.14 (M) The thin rod in Figure 4.50b of mass m and length L is hinged at point O. A force
is applied to the rod. Calculate the point (known as the center of percussion, also known
as the sweet spot) through which the force needs to be applied, so that the reactions at O
are zero the instant the force is applied.9 Ignore gravity (or think of the rod as horizontal).
4.15 (M) A body of mass m = 2 kg is moving with speed v = 6i − 3j m/s in the xy plane.
It is acted by a force of F = −4i + 3k N. Using normal-tangential coordinates, calculate the
change in speed and the radius of curvature.
4.16 (D) Pin P in Figure 4.51 moves without friction inside guide DA as arm OB rotates
with constant angular velocity θ̇ =0.3 rad/sec. Using polar coordinates, calculate the forces
exerted on the pin when θ = 45◦ . Hint: Each guide exerts a force on the pin that is
perpendicular to it.
4.17 (M) Two masses are connected to each other by a rope, as shown in Figure 4.52a. The
incline angle is θ = 30◦ and there is no friction at the ramp. The masses are released from
rest. For what ratio of m1 /m2 will the mass m2 begin to slide down?
4.18 (M) Two rods, each of mass m and length L, are released from rest in the position
shown in Figure 4.52b with θ = 30◦ , where the tip P makes frictionless contact with the
9 Note: The definition of sweet spot as the center of percussion is but one definition of sweet spot. Several

other definitions exist.


160 Applied Dynamics

O g
y

r
30 cm

A
D P
B
x

FIGURE 4.51
Figure for Problem 4.16.

a) b)

g g
m, L
m2 z
 m, L
s
 x

P
m1

FIGURE 4.52
Figures for a) Problems 4.17 and 4.22, b) Problem 4.18.

ground. Calculate the acceleration of the tip P and the angular accelerations of the rods at
the instant the rods are released.
4.19 (M) A rod of mass m and length L is released in the position shown in Figure 4.53a.
Calculate the acceleration of the center of mass and the angular acceleration if the contact
at point A is smooth (no friction).

Section 4.9—Dry Friction Forces


4.20 (E) A vehicle is entering a turn with radius of curvature 150 m. Its speed is 70 km/h.
If the maximum acceleration that the wheels can tolerate is 0.85g, how much (in g can the
vehicle decelerate or accelerate during the turn?
4.21 (M) Ring B of mass m in Figure 4.53b can slide over a circular rod OA, which is
rotating in the horizontal plane. The coefficient of static friction between the ring and the
arm is µs = 0.3. The arm starts to rotate with r = 0.15 m and constant angular acceleration
of θ̈ = 0.05 rad/s2 . Find the time it takes for the ring to begin to slide over the arm.
4.22 (M) Two masses are connected to each other by a rope, as shown in Figure 4.52a. The
incline angle is θ = 30◦ and coefficient of friction is µs = 0.2. The masses are released from
rest. For what ratio of m2 /m1 will the mass m1 begin to move down?
4.23 (M) A rod of mass m and length L is released in the position shown in Figure 4.53a.
Kinetics Fundamentals 161
x
a) b) Top view
B A
g
m, L y
B
x y r
G
o 
 A 75 O

FIGURE 4.53
Figures for a) Problems 4.19, 4.23, and 4.24, b) Problem 4.21.

Calculate the minimum value of the coefficient of friction necessary to prevent sliding of the
contact at point A.
4.24 (M) A rod of mass m and length L is released in the position shown in Figure 4.53a.
Friction is not sufficient to prevent sliding of the contact at point A. Calculate the angular
acceleration.

a) b)
g g

s
m

o F
 o 30 M
20

FIGURE 4.54
Figures for a) Problem 4.25, b) 4.26.

4.25 (M) A vehicle (Figure 4.54a) is traveling on a curved road with radius of curvature
ρ = 100 ft. The road is also banked by 20◦ . The coefficient of static friction between the
vehicle and the road is µs = 0.3. Using a particle model of the vehicle calculate a) the
minimum speed the vehicle needs to have before the vehicle slides down, b) the speed after
which the vehicle begins to slide up.
4.26 (M) The block of mass m = 0.5 kg in Figure 4.54b rests on a wedge of mass M = 3
kg, which makes an angle of 30◦ with the horizontal. A force F = 70 N is applied to the
wedge. Find the minimum coefficient of friction µ between the block and the wedge so that
the block will not move with respect to the wedge. Hint: Define the kinematic condition for
which the block will not slide over the wedge.

Section 4.10—Aerodynamic Forces


4.27 (C) Consider Example 4.6 and numerically integrate the motion equations for the
values of the drag-to-weight ratio and launch speed in Table 4.6 using an initial launch
angle of ψ0 = 35◦ . Then, increase the initial speed of each ball by 20%, which increases the
162 Applied Dynamics

drag-to-weight ratio by 1.22 = 1.44, and plot the response again. Compare the results and
draw conclusions.
4.28 (C) Consider the baseball in Example 4.6. Calculate and plot the distance traveled by
the ball as a function of the launch angle. Hint: You first need to develop the kinematic
differential equations associated with the displacements in the x and z directions. Consult
Section 3.6 for this.
4.29 (M) In bicycle aerodynamics, the expression 0.5ρCD A is referred to as drag factor KD
and the drag force is calculated as FD = KD v 2 . For most cyclists, the drag factor is in the
range 0.2 to 0.8 kg/m, with KD = 0.4 describing the average rider. Calculate the force such
a cyclist would have to exert to go at a speed of 9 km/h in the presence of a headwind of 2
km/h, given that the bicycle is going up a hill with a 4% grade. The combined mass of the
cyclist and bicycle is 92 kg.
4.30 (E) Consider Example 4.6 and calculate for each ball the speed at which a Reynolds
number of Re = 3 × 105 is reached.
4.31 (E) A boat of mass 2,000 kg is traveling at a speed of 12 m/s. At this speed, the
hydrodynamic drag force can be approximated by F = −kv. If the engines of the boat are
turned off, the boat’s speed reduces to 4 m/s in 10 seconds. Calculate the value of k and
the distance travelled by the boat during the 10 seconds.
4.32 (C) Terminal velocity of an object is reached when the aerodynamic drag acting on a
falling body becomes equal to the force of gravity. For an object of weight 2 lb, CD = 0.4
and planform area of 1.8 ft2 , calculate the value of the terminal velocity and the minimum
height this object needs to be dropped from so that it will reach terminal velocity. Assume
that density of air remains constant throughout the fall.

Section 4.11—Spring Forces

a) b)
2k k 2k
k
F 3k

2k F
k 2k

FIGURE 4.55
Figures for Problem 4.33.

4.33 (E) Calculate the equivalent spring constants in Figures 4.55a–b.

!k !k k
k F

k "k

FIGURE 4.56
Figure for Problem 4.34.
Kinetics Fundamentals 163

4.34 (E) Calculate the equivalent spring constant in Figure 4.56.

a) b)
F L
L
O G
P
L/2
k
x EI k B
2k 2k

FIGURE 4.57
Figures for a) Problem 4.35, b) Problem 4.36.

4.35 (E) Calculate the equivalent spring constant for a load applied at x = L in Figure 4.57a.
The beam has a stiffness of EI. The extensional spring has a constant k = 0.5EI/L3 .
4.36 (M) Calculate the equivalent spring constant for a load applied at x = L in Figure
4.57b. The force-rotation equation is P = keq Lθ. The hinged rod is of length L and is rigid
and assumed to be massless. Assume small rotations θ for the rod. What are the units of
keq ?
4.37 (E) A leaf spring is constructed out of a single steel beam of length 90 cm, width 10
cm, and thickness 1.2 cm. Assuming the spring can be modeled as hinged at both ends,
what is the equivalent spring constant for a load applied at the middle of the beam?

a) b)
c
k

k y
x

m
d
x
e

FIGURE 4.58
Figures for a) Problem 4.38, b) Problem 4.39.

4.38 (M) The mass in Figure 4.58a is connected to spring k at an angle β. Assume that β
remains almost constant as the mass vibrates. Express the component of the spring force
along the direction of motion as Fx = keq x and show that keq ≈ k cos2 β.
4.39 (D) The MacPherson strut in Figure 4.58b has a spring of constant k. Assume that
the tire is rigid, so that the vertical motion of the tire is y. Also assume that the angle β
remains almost constant as the wheel moves up and down. Express the component of the
spring force along the direction of motion as Fy = keq y and show that keq ≈ kd2 cos2 β/e2 .
164 Applied Dynamics

What is the equivalent damping constant? Note: The equivalent spring constant is called
the ride rate of the suspension.

a) b) z
k R
g F
A L
A y
2L B
3L Diameter
d
k
O 
x Fixed

FIGURE 4.59
Figures for a) Problem 4.40 and b) Problem 4.41.

4.40 (M) The L-shaped bracket AOB in Figure 4.59a has mass density (mass/unit length)
ρ and is connected to two springs. The system is released from rest when θ = 15◦ . Calculate
the angular acceleration of the bracket. The springs are not deformed when θ = 0.
4.41 The torsion bar in Figure 4.59b is a type of suspension in cars. The bar consists of
a rod of length L and diameter d, and an arm of length R. A vertical force applied to
the arm generates a moment on the rod, which deforms and creates a resistive moment.
Calculate the equivalent spring constant for the torsion bar (ratio of the applied force F
and deformation of point A in the −z direction.
5
Kinetics Applications

5.1 Introduction
This chapter discusses applications of the kinetics principles developed in the previous
chapter. We consider rolling motion, mechanical trail, impulsive loads, and impact, as well
as impulse-momentum and work-energy principles. The concepts of equations of motion
and equilibrium are introduced, together with equilibrium. Linearization of the equations
of motion about equilibrium is discussed. Motion in the vicinity of the Earth is studied. An
expanded analysis of impact is presented, together with applications from sports.
The examples considered in this chapter are primarily for two-dimensional motion of
rigid bodies and three-dimensional motion of particles. Three-dimensional motion of rigid
bodies will be considered in Chapter 11.

5.2 Rolling
This section considers the dynamics of rolling. We assume that the rolling body is a rigid disk
(or sphere) and that it makes point contact with the surface over which rolls. The main
difference between rolling and sliding is in the way the friction force affects the motion.
When moving an object by sliding it, the friction force works against the applied propulsive
force, resisting the motion. For wheels and rollers, the friction force is beneficial, generating
the propulsive forces along the direction of the desired motion.
When analyzing rolling of wheels on a vehicle, it is important to distinguish between
the powered and the nonpowered wheels. For example, in a bicycle or a rear wheel drive
vehicle, the rear wheels are powered. The power generated is transmitted to the wheel via
a torque. In a bicycle, the torque is generated by the pedals and chain; in a car, the torque
is transmitted by the drive shaft and differential.

v
G x
c b
h
A B z

C1 C2

FIGURE 5.1
Side view of moving vehicle.

165
166 Applied Dynamics

Figure 5.1 shows the side view of a rear wheel drive vehicle that is moving forward (in x
direction) and is not braking. Separating the tires from the vehicle, the free-body diagram
of the vehicle body is shown in shown in Figure 5.2. The rear (powered) wheel and the front
(nonpowered) wheel are shown in Figure 5.3.
When determining the direction of the friction force, we need to look at the forces and
moment acting on the wheels. For the powered rear wheel in Figure 5.2a, summing moments
about the axle A, in the absence of friction the wheel will rotate clockwise and the contact
point C1 will move to the left. The friction force Ff that is generated acts to the right. This
is intuitively obvious, as it is this friction force that propels the vehicle.

v, a
G
Mg x
Az Bz
Ax Bx z
T

FIGURE 5.2
Free-body diagram of vehicle body.

a) b)
T
Az Bz

A B
Ax v Bx v
mg mg
C1 C2

Ff N1 F2 N2

FIGURE 5.3
Free-body diagrams of a) rear wheel (powered), b) front wheel (nonpowered).

Consider now the nonpowered front wheel in Figure 5.3. The front wheel moves because
the body of the vehicle exerts a force Bx on it. Summing forces along the horizontal, in
the absence of friction the contact point C2 will slide forward. The friction force F2 at C2
opposes that motion, and hence, it is directed backwards. When there is roll without slip,
the friction force acting on the nonpowered wheel is a very small quantity; for an ideal (rigid,
point contact only) wheel, the friction force is zero when the wheel moves with constant
speed.
Consider the free-body diagram of the vehicle body as the vehicle accelerates forward,
shown in Figure 5.2. By virtue of Newton’s Third Law, a counterclockwise torque is exerted
on the vehicle body by the rear wheel as the vehicle accelerates forward. This torque has
the effect of rotating the vehicle body counterclockwise, lifting the front of the body, an
effect that we can feel when accelerating.
Kinetics Applications 167

When we apply the brakes, the braking mechanism creates a counterclockwise torque. In
the absence of friction, the torque has the tendency to decelerate the wheel counterclockwise,
so that the contact point decelerates to the right. The friction force that is associated with
braking is opposite the impending acceleration; the friction force points to the left, as seen
in Figure 5.4. This force slows the vehicle. As the wheel experiences a counterclockwise
torque, the body of the vehicle experiences a clockwise torque and the body dives forward
before it stops.

z y

T
Az x
v
A
Ax
a
mg

Ff N

FIGURE 5.4
Free-body diagram of braking rear wheel.

We determine whether there is roll with slip or roll without slip the same way as when
we analyze the sliding of blocks. Consider Figure 5.3a, which shows a powered rolling wheel,
under the action of a torque. When there is roll without slip, the friction force is less than
its maximum possible value, or Ff < µs N , where µs is the static coefficient of friction and
the velocity of the contact point is zero. When there is roll with slip, the friction force is
known and it is Ff = µk N , where µk is the kinetic coefficient of friction, but the velocity
of the contact point is not known. Table 5.1 summarizes the conditions of rolling.

TABLE 5.1
Summary of slip and no slip conditions for rolling

Roll without Slip Roll with Slip


Friction force Ff Ff < µs N Ff = µk N
Velocity of contact point C 0 Unknown
Speed of geometric center vG RΩ 6= RΩ

In general, it is not known whether the roll is with slip or without slip until the problem
is solved. As in the case of sliding, we begin with an assumption, solve the problem, and
then check the validity of the assumption. Two types of assumptions are possible:
• We first assume that there is no slip. The accelerations and magnitude of the friction
force are calculated. The value of the calculated friction force Ff is compared with the
maximum available friction force, µs N . If Ff < µs N , the no-slip assumption is correct.
168 Applied Dynamics

If not, friction is not sufficient to prevent slip, there is slipping, and Ff = µk N . The
problem is solved again considering that there is slip and that Ff = µk N .
• We assume that there is slip. The translational accelerations of contact point and of
the center of rotation, as well as angular accelerations, are calculated using Ff = µk N .
These accelerations are then compared against their expected values by means of the
relative acceleration equation. This procedure is best explained by an example.

Example 5.1

a) b)
T T

z y
mg
G G
vG ,aG x
R

Ff
N

FIGURE 5.5
a) Accelerating disk. b) Free-body diagram.

A torque T is applied to the disk in Figure 5.5a, which is at rest. What is the smallest
value of the torque that will make the wheel roll with slip? The coefficients of friction are
µs and µk .
The free-body diagram is shown in Figure 5.5b. We can solve this problem by beginning
with either the no-slip assumption or the slip assumption. Consider first the slip assumption,
so that we assume the wheel to slip. The associated friction force is Ff = µk N . The force
balance in the vertical direction gives N = mg. The force balance along the horizontal gives
+
X
→ F = maG =⇒ maG = Ff = µk N = µk mg [a]

so that aG = µk g. The moment balance about the center of mass gives


X
 MG = IG Ω̇ =⇒ IG Ω̇ = T − Ff R [b]

Solving for the angular acceleration,


T − µk mgR
Ω̇ = [c]
IG

where for a disk IG = mR2 /2. For slip to occur, the acceleration of the contact point must
be to the left, so that the relation

aG − RΩ̇ < 0 [d]


Kinetics Applications 169

needs to be satisfied. Substitution for the values for the acceleration and angular acceleration
from Equations [a] and [c] into Equation [d] gives

T R − µk mgR2 2T
µk g − 1 2
= 3µk g − < 0 [e]
2 mR mR

which, when solved for the torque T , gives


3
T > µk mgR [f ]
2
We have calculated the minimum value of the torque to provide slip. However, this
answer is not completely correct. Because the static coefficient of friction is larger than the
kinetic coefficient, the torque has to exceed the static coefficient threshold, which can be
obtained by replacing µk in the above equation with µs . The minimum torque that will
result in the wheels slipping is then
3
T > µs mgR [g]
2

Next, consider the no-slip assumption. Here, aG = RΩ̇ and the magnitude of the friction
force is unknown. The horizontal force balance gives
+
X
→ F = maG =⇒ maG = Ff [h]

and the moment balance about the center of mass is Equation [b]. Substituting the values
of aG = Ff /m from Equation [h] into Equation [b] gives

IG + mR2 Ω̇ = T

[i]

Solving for Ω̇ yields


2 T
Ω̇ = [j]
3 mR2
The next step is to calculate the magnitude of the friction force, which from Equation
[h] is
2T
Ff = maG = mRΩ̇ = [k]
3R
The largest value of the friction force is Ff = µs mg. Substituting this value in the above
equation and solving for the minimum torque gives the same result as Equation [g].

5.3 Mechanical Trail


In a car the wheels that perform the steering are in the front. Steering the car forward is a
much more stable operation than driving it backwards. The same is true with a supermarket
cart, a stroller, or any other vehicle with a caster, such as the one shown in Figure 5.6a. A
caster is a wheel whose pivot axis connecting it to the vehicle is in front of the center of the
wheel.
The presence of a caster creates a mechanical trail, which provides the cart with direc-
tional stability. The concept of mechanical trail has been known for centuries. Historical
170 Applied Dynamics

#3 43 53
ȍ
! 1$/02$/
" -+#.
ȍ #/0%+ ȍ
&'"(
'))*+,
#

!"#$% !"#$%

FIGURE 5.6
a) Caster, b) front tire of a car, c) front wheel of a bicycle.

pictures of bicycles and of armchairs with wheeled legs show the use of casters. Figures 5.6b
and 5.6c depict the mechanical trail that arises in car tires and bicycle wheels, respectively.
To understand mechanical trail, consider the top view of a caster assembly, say a front
wheel of a supermarket cart, as the cart is making a left turn. The top view and free-body
diagrams are shown in Figure 5.7. The assembly pivots about point B, and the distance
between the pivot and center of the front wheel is e, to which we will refer as the trail.
The shopper seeking to turn left applies a force that is transmitted to the wheel assem-
bly at an angle θ. The front wheels immediately align themselves in the direction of the
force. Consider the top view of the free-body diagram. The propelling force F creates a
counterclockwise moment of magnitude M = F e sin θ on the wheel, forcing the wheel as-
sembly to turn counterclockwise, as shown in Figure 5.7b. The wheel eventually assumes an
orientation parallel to the applied force. The turning motion ends when the caster assembly
is aligned with the direction of the applied force and there no longer is a moment M acting
on the assembly.

a) b)

C M
A F
a B v
A x
r  Ff
t
F Fslide
y
e

FIGURE 5.7
a) Top view of front wheel of supermarket cart, b) top view of free-body diagram of wheel.

What regulates the turning motion of the wheel are the friction forces and moments
generated at the area of contact between the wheel and the floor. Figure 5.8 shows the
free-body diagram of the wheel. The friction force Ff along the line of motion (x) generates
the rolling motion and Fslide counters the component of F in the y direction, F sin θ. The
friction moment Mf is generated by the distribution of Fslide along the contact area between
the wheel and the floor.
Shoppers pushing supermarket carts that are warped or out of alignment observe that the
caster assembly often rattles, rather than providing a smooth turn. This happens because
Kinetics Applications 171

–z –y

Fsin
A
Fcos x
v

Ff Mf
Fslide
y y
z

FIGURE 5.8
Side view of free-body diagram of wheel.

there is not continuous contact between the caster wheel and the floor, thus preventing
continuous presence of the friction forces and moment. As a result, the turn of the caster
does not come to a smooth stop. A simple way of dealing with an annoying cart like that is
to place the heavier products in the cart over the caster that is uneven. This action pushes
the cart down and increases the amount of contact between the caster wheel and the floor
(or better, get another cart!).
Next, consider what happens if the shopper pushes a supermarket cart backwards (from
the front end of the cart, opposite the handle). Here the force F is behind the pivot, as
shown in Figure 5.9a. The net effect is a moment in the clockwise direction (when viewed
from the top), which makes the wheel assembly turn away from the direction of the applied
force. The entire assembly then rotates in the opposite direction by over 90◦ so that the
pivot point B is now in front of the wheel assembly. You can observe this effect by pushing
the cart backward and looking at the motion of the pivoting wheels. Also, it is harder to
navigate a cart that is pushed backwards.

a) b)
Fslide
C A M
B Ff
a v
A
r

t
F
F

FIGURE 5.9
a) Supermarket cart pushed backwards, b) top view of free-body diagram of wheel.

A caster essentially creates a frictionless point of contact. The text by Karnopp provides
a mathematical model of a caster. If the rear wheels are also replaced with casters, as we
sometimes find in small supermarkets or carry-on luggage, then the cart can move in every
172 Applied Dynamics

direction (forward, sideways, rotational) as if it is resting on a frictionless surface. However,


generating three types of motion with two inputs (two arms pushing the cart) is not easy,
as anyone who has navigated such a cart knows.

5.4 Impulse and Momentum


This section examines the effects of forces and moments on bodies over a period of time.
Consider Newton’s Second Law for a rigid body, F (t) = dp (t) /dt. Multiplying this equation
by dt and integrating over time, from an initial time t1 to final time t2 , gives
Z t2 Z p(t2 )
F (t) dt = dp = p (t2 ) − p (t1 ) = mvG (t2 ) − mvG (t1 ) (5.1)
t1 p(t1 )

Equation (5.1) is the impulse-momentum theorem for a rigid body. The term on the
Rt
left, t12 F (t) dt, is called the impulse, which is the cumulative effect of the force over a time
period, and is equal to the change over time of the linear momentum. This definition is
different from the colloquial use of the word impulse. The units of impulse are force × time.
When the left side of the above equation is zero, that is, the integral of the applied
force over the time interval of interest is zero, then the initial and final linear momenta
become the same. This is known as the principle of conservation of linear momentum. The
corresponding relationship for a particle is obtained by replacing vG with v.
We obtain the angular impulse-momentum theorem in a similar way. Integrating the
d
general form of the moment balance equation, MG = dt HG , over time results in
Z t2 Z HG (t2 )
MG (t) dt = dHG = HG (t2 ) − HG (t1 ) (5.2)
t1 HG (t1 )

When the integral over time of the applied moment is zero, there is conservation of
angular momentum over that range of time. For the special case of plane motion, the
angular impulse-momentum theorem becomes
Z t2
MG (t) dt = IG ω (t2 ) − IG ω (t1 ) (5.3)
t1

For a particle, measuring the angular momentum about a fixed point D, as shown in
Figure 4.11, the angular impulse-momentum theorem, when integrated over time becomes
the same as Equation (5.2), with G replaced by D.

5.4.1 Impulsive Forces


A very commonly encountered type of an external force is that of a large force applied over
a very short time period. Forces involving contact, such as a batter hitting a baseball or a
collision between two vehicles, result in the colliding bodies exerting very large forces on
each other during the brief moment of contact. Such forces are known as impulsive forces.
The simplest approximation of an impulsive force assumes that the magnitude of the
force remains constant during the application of the impulse, as shown in Figure 5.10a.
Denoting the magnitude of this average force by Fave and the duration by ∆, the impulse
can be expressed as
Z t0 +∆
F̂ = F (t) dt = Fave ∆ (5.4)
t0
Kinetics Applications 173

More accurate models assume that the magnitude of the impulsive force varies during
the impulse, beginning and ending with zero but increasing (and then decreasing) during
the application of the impulse. A commonly used approximation is the half-sine function
(Figure 5.10b)

a) b) c) d)
F F F F

B=2Fave C=2Fave

A= Fave
2
Fave

t0 t t0 t t0 t t0 t

FIGURE 5.10
Approximations of impulsive forces: a) Rectangular, b) half sine, c) sine squared, d) trian-
gular.

π 
F (t) = A sin (t − t0 ) t0 ≤ t ≤ t0 + ∆ (5.5)

Introducing the variable τ = t − t0 and noting that
Z t0 +∆ Z ∆
f (t − t0 ) dt = f (τ ) dτ (5.6)
t0 0

Integration of Equation (5.5) over the duration of the impulse gives


Z ∆ π   π  ∆
∆ ∆
F̂ = A sin t dt = −A cos t = 2A (5.7)

0 ∆ π ∆ 0 π
from which the value of A is obtained as
π F̂ π
A = = Fave (5.8)
2∆ 2
Comparing with Equation (5.4), the amplitude A for the half sine is π2 times higher than
the average value Fave .
Another approximation is the sine squared curve (Figure 5.10c) in the form
π 
F (t) = B sin2 (t − t0 ) t0 ≤ t ≤ t0 + ∆ (5.9)

Integrating the above equation over the duration of the impulse gives
Z ∆ π  B∆
F̂ = B sin2 t dt = (5.10)
0 ∆ 2
from which the value of the amplitude B can be calculated as

2F̂
B = = 2Fave (5.11)

174 Applied Dynamics

According to experimental studies, the sine squared curve approximates actual collisions,
such as a baseball being hit by a bat, more closely than the rectangular pulse or half sine
curves.
Yet another way of modeling impulsive forces is by treating them as a triangular pulse
(Figure 5.10d). Denoting the amplitude of the impulsive force by C and its duration by ∆,
the impulse becomes F̂ = C∆ 2 , which is the same as Equation (5.10), so that the triangular
pulse and the sine squared profile have the same maximum amplitude.
We assume that an impulsive force causes a sudden change in velocity with no change
in position. For the translational motion of a particle, we can write
F̂ (t0 )
vG (t0 + ∆) = vG (t0 ) + (5.12)
m
with rG (t0 + ∆) = rG (t0 ).

5.4.2 Idealized Model of an Impulsive Force

į!!"#"!"#

!" !

FIGURE 5.11
Dirac delta function.

The idealized model of an impulsive force is described by means of the Dirac delta
function. Visualized (because its actual shape is a spike) in Figure 5.11 and denoted by
δ (t − t0 ), the Dirac delta function is defined as

Z ∞
δ (t − t0 ) = 0 when t 6= t0 δ (t − t0 ) dt = 1 (5.13)
0

The unit of the Dirac delta function is 1/time. In the idealized model, the amplitude
goes to infinity and the duration of the impulse approaches zero. It is of interest to integrate
the product of a function, say, f (t), with the Dirac delta function, which yields
Z ∞ Z ∞ Z ∞
f (t) δ (t − t0 ) dt = f (t0 ) δ (t − t0 ) dt = f (t0 ) δ (t − t0 ) dt = f (t0 ) (5.14)
0 0 0

The above relationship is analogous to taking a snapshot of the function f (t) at time t = t0 .
In essence, the Dirac delta function is defined by what it does to a function.
The forcing function associated with an impulsive force F̂ applied at time t0 is described
as
f (t) = F̂ δ (t − t0 ) (5.15)
Response to impulsive excitation will be studied in the next chapter.
Kinetics Applications 175

Example 5.2
In a baseball game the batter hits a baseball that is traveling toward it at a speed of 90
mph. As a result of the hit, the ball acquires a speed of 112 mph. Considering that the
contact between the bat and the ball takes place for 0.7 milliseconds, what is the average
force that is applied by the bat to the ball? The baseball weighs 5 18 oz.
Let us approximate the impulsive force as an average force Fave applied over a time
period ∆ = 7 × 10−4 seconds, so that
Z 7×10−4
m (v2 − v1 ) = F dt ≈ Fave ∆ [a]
0

which can be solved for the average force as (note m = W/g and 60 mph = 88 ft/sec)

v2 − v1 5.125 (112 + 90) × (88/60)


Fave = m = = 4206 lb [b]
∆ 16 × 32.17 7 × 10−4
which is quite a large force. If the sine squared model is used to describe the impulsive force,
as discussed in Section 5.4.1, the maximum value of the force is twice that, or 8412 lb.
The above result leads to interesting observations. First, the assumption of ignoring all
nonimpulsive forces during the application of the impulse is a good one. The weight of the
ball is about 1/3 lb, and the force of gravity is about 10,000 times smaller than the average
value of the impulsive force. Second, the importance of what every coach says when hitting a
ball in any sport becomes obvious: follow through. Following through increases the duration
of impact, resulting in higher impact forces, and also provides better directional stability
to the ball.
It is interesting to calculate the average acceleration of the ball during impact. Approx-
imating Newton’s Second Law during impact as

Fave = maave [c]

we find the average acceleration (in g) as

aave Fave 16
= = 4206 × = 13, 131g [d]
g mg 5.125
which is very large. Of course, this acceleration takes place over a very short time period.

5.5 Work, Energy, and Power


The previous section considered the effects of forces and moments on a body over time.
The next step is to analyze these effects over a displacement of the body from one point
to another. This is accomplished by integrating the translational and rotational equations
over displacement. In the process, we arrive at useful quantities: work, kinetic energy, and
potential energy.

5.5.1 Kinetic Energy and Power


Consider Figure 5.12 and a particle of mass m, whose position is denoted by the vector r.
A force F acts on the particle. The incremental work is denoted by dW and defined as the
176 Applied Dynamics

2
F
m
1

r1 r r+dr t
r2 dr

Reference point

FIGURE 5.12
Force acting on particle moving along a path.

work that the force does on the particle as the particle moves by an incremental distance
dr

dW = F · dr = |F| |dr| cos ψ = ma · dr (5.16)

where ψ is the angle between F and dr. Recalling that v = dr/dt and a = dv/dt, and
multiplying and dividing the above equation by dt gives
dv dr 1
dW = F · dr = m · dt = mv · dv = md (v · v) (5.17)
dt dt 2
The kinetic energy T of the particle is defined as
1 1
T = mv · v = mv 2 (5.18)
2 2

where v = v · v is the speed. The incremental change in the kinetic energy is dT =
md (v · v) /2 = m v · dv. Kinetic energy is an absolute quantity, and hence, dT is a perfect
differential. Introducing the above equation to Equation (5.17) we can write

dW = dT (5.19)

The work done on the system by the force is denoted by W1→2 , which is obtained by
integrating the motion from an initial point 1 to point 2; hence,
Z r2 Z T2
W1→2 = F · dr = dT = T2 − T1 (5.20)
r1 T1

which gives the work-energy theorem as

T1 + W1→2 = T2 (5.21)

Power is defined as the rate at which work is done, or


dW
P = (5.22)
dt
where P is power and W is work. A more powerful motor does more work during a given
period of time than a less powerful motor.
Kinetics Applications 177

For translational motion, the incremental work done by a force on a body is dW = F·dr,
so that the expression for power becomes
dW dr
P = = F· = F·v (5.23)
dt dt
where v is the velocity of the point to which the force is applied. In scalar form, P = F v. For
rotational motion in a plane, following a similar approach to translational motion, we can
write P = M ω, where M is the applied torque and ω is the angular velocity. From Equation
(5.23) we can write P dt = F·dr, which makes it possible to express the work-energy theorem
in terms of power as
Z t2
W1→2 = P dt (5.24)
t1

Some texts define work as the integral of power over time.


In U.S. customary units force is expressed in pounds (lb) and velocity in ft/sec, so one
way of expressing the units of power is ft·lb/sec. It has become customary to express power
in terms of the unit horsepower (hp). One horsepower is defined as 1 hp = 550 ft·lb/sec.1
When dealing with rotational motion, the standard unit for describing angular displace-
ment is the radian, so when we describe power by P = T ω, the angular velocity is in terms
of radians/second, and the definition of horsepower becomes 1 hp = 550 ft·lb/sec. Many
times in engineering applications, it is preferable to describe rates of rotation by revolutions
60
per minute, or rpm. Noting that 1 rad/sec = 2π = 9.549 rpm, horsepower can be defined
as 1 hp = 550 × 9.549 ft·lb rpm = 5252.1 ft·lb rpm.
A commonly used unit for power in the SI system is Watt (W) or Kilowatt (kW). One
Watt is defined as 1W = 1 N·m/s. We can show that 1 hp = 0.7457 kW.
The kinetic energy of a rigid body can be obtained by a differential element approach
and by summing the kinetic energies of each individual element. The procedure can be
found in standard texts on dynamics. We can show that the kinetic energy of a rigid body
can be expressed as
T = Ttr + Trot (5.25)
in which Ttr is the translational kinetic energy and has the form
1
Ttr = mvG · vG (5.26)
2
and Trot is the rotational kinetic energy and has the form
1
Trot = ω · HG (5.27)
2
For the special case of rotation about a fixed point O, we can show that the kinetic
energy can be expressed as
1
T = ω · HO (5.28)
2
The expressions above for the kinetic energy are valid for any type of motion. For plane
motion, the rotational kinetic energy can be expressed as
1
Trot = IG ω 2 (5.29)
2
1 There are other definitions of horsepower that are used less frequently. It is important to clarify which

definition of horsepower is being used.


178 Applied Dynamics

and for rotation about a fixed point O the kinetic energy is T = 21 IO ω 2 .


For plane motion, the kinetic energy can also be written about the instantaneous center
of zero velocity. Denoting the instant center by C, the kinetic energy becomes T = 21 IC ω 2 .

Example 5.3
A cyclist is traveling with speed v into a headwind of va . Assuming that the only force the
cyclist fights is drag, calculate i) the power the cyclist has to generate to maintain speed
with a headwind of va = v/4, and ii) the power that needs to be generated if the cyclist
wishes to increase her speed by 25% in the absence of a headwind.
The most important force that a cyclist has to fight is aerodynamic drag. The drag force
is expressed as
2
FD = 0.5ρCD A (v + va ) [a]
where CD is the drag coefficient and A is the planform area of the bicycle-rider system.
Cyclists lean over and lower their bodies when riding to reduce this planform area, as well
as to lower the center of mass of the bicycle-rider system. The power that the cyclist needs
to generate is
2
P = FD v = 0.5ρCD Av (v + va ) [b]
so that for a headwind of va = v/4, the power requirement becomes
2
P1 = FD v = 0.5ρCD Av (1.25v) = 1.56P [c]

The power output has to go up by 56% to maintain speed while fighting the headwind.
For the second case, the power needed when the speed is increased to 1.25v becomes
2
P2 = 0.5ρCD A1.25v (1.25v + va ) [d]

In the absence of a headwind, va = 0, the power requirement for increasing speed by


25% becomes P2 /P = 1.253 = 1.953, so that a near doubling of the power generated by the
cyclist is required to increase speed by 25%.

5.5.2 Potential Energy


The expression for incremental work dW was defined earlier. It is useful to distinguish
between cases where dW is a perfect differential and where it is not. A perfect differential
can be expressed as the derivative of a function. It also leads to the development of the
potential energy.
For dW to be a perfect differential, the force F must be dependent on the position r alone,
F = F (r). Such a force is referred to as a conservative force. Examples of conservative forces
include gravity and spring forces, as well as internal forces of elastic bodies, such as beams
and taut strings. The incremental work in these cases can be expressed as the derivative of
a potential function

dW = F (r) · dr = −dV (r) (5.30)

When dW is a perfect differential, its integral is independent of the path followed and
its value is dependent only on the end points of the integration. Over a closed path the
value of the integral is zero, or
I
F (r) · dr = 0 (5.31)
Kinetics Applications 179

We can evaluate the potential function from a reference position rR (or datum) to yield
Z r
V (r) = − F (r) · dr (5.32)
rR

The potential function V (r) is known as the potential energy. Potential energy can be
explained as the potential of a body or a component to do work, or the stored energy in a
system.
Note that while kinetic energy is an absolute quantity, potential energy is relative: its
value depends on the reference position from which it is measured. Because the interest is
in change of potential energy, selection of the reference point does not make any difference.
We select the reference point, or datum, to simplify calculations or to give the problem at
hand a better physical interpretation. The units of potential energy are the same as the
units of kinetic energy.
Because Equation (5.31) involves a line integral, we can invoke Stokes’ theorem and
write for a conservative force

∇×F = 0 (5.33)

where ∇ is the del operator. The above expression becomes zero only if the force F can be
expressed as the gradient of a function. It is easy to show that this function is the negative
of the potential energy. A conservative force can thus be expressed in terms of the potential
energy associated with it as

F (r) = −∇V (r) (5.34)

In Cartesian coordinates the del operator has the form


∂ ∂ ∂
∇ = i+ j+ k (5.35)
∂x ∂y ∂z
so that, expressing the force vector as F = Fx i + Fy j + Fz k, we can relate the components
of the force vector to the potential energy as
∂V ∂V ∂V
Fx = − Fy = − Fz = − (5.36)
∂x ∂y ∂z

5.5.3 Gravitational Potential Energy


The gravitational attraction between two bodies is given in Section 4.7 as F = Gm1 m2 /r2 ,
where G is the universal gravitational constant, m1 and m2 are the masses of the two bodies,
and r is the distance between the centers of mass of the two bodies, as shown in Figure
5.13a. Consider m1 to be a celestial body. The incremental work done by m1 on m2 is a
perfect differential, so we can write the potential energy as
Z r Z r r
Gm1 m2 Gm1 m2 R
V (r) = − F (r) dr = − dr = (5.37)
rR rR r2 r
r

The datum position is commonly selected as the distant stars, so that rR = ∞, which
leads to the expression for the gravitational potential energy:
Gm1 m2
V (r) = − (5.38)
r
In the vicinity of the Earth (Figure 5.13b), m1 = me is the mass of the Earth and
180 Applied Dynamics

a) b)
m
F
r
m2
h
F(r)
re
F(r )
m1 me

FIGURE 5.13
Gravitational force: a) two general bodies, b) in vicinity of the Earth.

m2 = m is the mass of the body of interest, r is the distance between the center of the
Earth and the body, or r = re + h, where h is the altitude. Using the relationship for small
values of , 1/(1 + ) ≈ 1 − , the potential energy due to a gravitational force becomes
 
Gme m Gme m Gme m h
V = − = −   ≈ − 1−
r re 1 + rhe re re

Gme m Gme m
= − + h (5.39)
re re2
The first term on the right becomes zero if the datum is selected as the surface of the
Earth. Recalling from Chapter 1 the definition of the gravitational constant near Earth as
g = Gm
r 2 , the gravitational potential energy is written as
e
e

V = mgh (5.40)

We can also obtain the gravitational potential energy via Equation (5.32). Noting that
the force of gravity is F = −mgk and the position vector is r = zk so that dr = dzk,
Equation (5.32) gives
Z r Z h Z h
V = − F (r) · dr = − −mgk · dzk = mgdz = mgh (5.41)
rR 0 0

Example 5.4
Consider the pendulum in Figure 5.14 and write its potential energy for different choices of
the datum.
As the pendulum swings with angle θ, the vertical distance between the pivot and
pendulum becomes L cos θ. There are two obvious choices for the datum. One is to select
the datum point line to go through the pivot. In this case, the pendulum is below the pivot
and its potential energy becomes

V1 = −mgL cos θ [a]

A second suitable choice for the datum is the position of the pendulum when θ = 0.
In this case, the pendulum is always at or above the datum and the potential energy is
positive. The vertical distance of the pendulum from the datum is L (1 − cos θ), so that the
potential energy is
V2 = mgL (1 − cos θ) [b]
Kinetics Applications 181

Datum 1

Lcos L

m L(1– cos)

Datum 2

FIGURE 5.14
Different choices for datum.

As expected, the two potential energies are separated by a constant

V2 − V1 = mgL [c]

5.5.4 Potential Energy of Springs


Consider an axial spring and deformation only in the linear range, as shown in Figure
5.15a. In Figure 5.15b, the spring is stretched in the positive x direction and the spring
force becomes Fs (x) = −kx, where the minus sign indicates the direction of the spring
force.

a) b) c)
Deformed
position
k
F
FS F
x
x
Undeformed
position Potential
energy

FIGURE 5.15
a) Deformed spring, b) free-body diagram and spring force, c) Potential energy.

The potential energy can be calculated as


Z x Z x
1 2
V (x) = − Fs (x) dx = − −kxdx = kx (5.42)
0 0 2

which is recognized as the area of the shaded triangle in Figure 5.15c. Similarly, for a
torsional spring of constant kT , the potential energy becomes V (θ) = 12 kT θ2 . It is preferable
when obtaining the governing equations of a system acted upon by springs to select the
datum as the undeformed position of the spring.
182 Applied Dynamics

5.5.5 Work-Energy Relations


Some of the forces acting on a body may be conservative and some may not. Forces that are
not conservative are referred to as nonconservative. The resultant force vector in terms of
the conservative and nonconservative forces is F = Fc + Fnc , where the notation is obvious.
It follows that the incremental work done by the forces can also be divided into two parts;
that is,
dW = dWc + dWnc = −dV + dWnc (5.43)
where dWnc is the incremental work done by the nonconservative forces. The total work is
obtained by integrating the incremental work, with the result
Z r2 Z r2 Z r2
W1→2 = dW = (Fc + Fnc ) · dr = V1 − V2 + Fnc · dr (5.44)
r1 r1 r1

where the last term on the right is the work done by the nonconservative forces
Z r2
Wnc1→2 = Fnc · dr (5.45)
r1

Substituting Equation (5.44) and Equation (5.45) into Equation (5.20) the work-energy
theorem becomes
T1 + V1 + Wnc1→2 = T2 + V2 (5.46)
The total energy of the system is defined as the sum of the kinetic and potential energies,
E = T + V . Using this notation, the energy balance becomes
E1 + Wnc1→2 = E2 (5.47)
The total energy of a system changes when nonconservative forces act on it. When all
the forces acting on the body that do work are conservative, the total energy of the body
remains the same. This is known as the principle of conservation of energy and it explains
the terms conservative force and nonconservative force. This principle can be written as
E1 = E2 (5.48)
In reality, there always are nonconservative forces that act on a body that dissipate energy, so
that there is energy loss. The assumption of energy conservation is a simplifying assumption.

5.5.6 Forces That Do No Work


A special category of forces consists of forces that do no work. Recalling the definitions of
incremental work and power as dW = F · dr and P = F · v, respectively, then for a nonzero
force to not do any work either dr = 0 (or v = 0) or the force F is perpendicular to dr (or
to v). Included in the category of forces that do not do work are
• Normal forces and reaction forces perpendicular to the direction of motion (perpendic-
ular to the tangential direction); F is perpendicular to dr.
• Forces applied to points that have zero velocity or to a stationary point. A force in
this category is the friction force for rolling problems where there is no slip. Because
in rolling without slip over a fixed surface the velocity of the contact point is zero, the
force is always applied to a point that has an instantaneous zero velocity.
When a pneumatic tire rolls, an area of contact develops between the tire and the road
called the contact patch, and the friction force is distributed along the contact area.
Because not every point on the contact patch has zero velocity, the friction force in
pneumatic tires ends up doing negative work (dissipating energy).
Kinetics Applications 183

5.5.7 Hysteresis and Energy Loss


As discussed earlier, the load versus deformation curves of most materials and devices
resemble that of a softening spring. This means that once the linear range is exceeded the
increase in the resistive force becomes less than the increase in deformation. The stress-
strain curve of materials, the cornering stiffness of tires, tire deformation due to contact
with the ground, and impact forces developed during collisions all fall into this category.

a) b) c)

Load Load Load


Loading
Unloading
Energy
released
while
unloading
Linear Deformation Energy Deformation Lost Deformation
range energy

FIGURE 5.16
Loading and unloading curves. a) Linear range, b) loading and unloading in linear range,
c) loading and unloading beyond linear range.

Consider an elastic member or component whose load deformation curve behaves like a
softening spring, such as in Figure 5.16a. When loading and unloading this material with
forces that do not exceed the linear range, we go back and forth along the same line. Hence,
the work done by loading and unloading is the same, and it is represented by the area of
the triangle in Figure 5.16b.
Next, consider loading this material with a force that exceeds the linear range. When
removing the load, the material follows a different path in the load-deformation curve. This
path is usually parallel to the linear part of the loading curve, as shown in Figure 5.16c. It
follows that the energy used for loading is higher than the energy released when unloading.
The difference, that is, the lost energy, is the area enclosed by the loading and unloading
curves. The phenomenon of energy loss due to difference in force magnitudes during loading
and unloading is known as hysteresis, and materials that exhibit this type of behavior are
called hysteretic.
The resistive force generated by a hysteretic material, spring, or device is not a conser-
vative force. Even though the loading and unloading curves may individually be expressed
as a derivatives of functions, we cannot obtain a single expression for potential energy be-
cause we cannot specify whether the hysteretic material is expanding or compressing by
just looking at its position.
The deformation and expansion of the contact area of a rolling tire with the road is
affected by hysteresis, which leads to the rolling resistance force, which resists the motion
of the tire. Rolling resistance is described in detail in Section 13.7.
Hysteresis forces can be included in the description of motion by a mathematical de-
velopment called complex damping. The interested reader is referred to texts on vibration
modeling. It should be noted that, for some materials, even though the loading exceeds the
linear range, there is no yielding or permanent deformation. Pneumatic tires, for example,
are in this category.
184 Applied Dynamics

In addition to hysteresis, this and the previous chapter discussed three types of forces
that dissipate energy: aerodynamic (or hydrodynamic) drag, dry friction, and viscous damp-
ing. Energy loss is usually due to a combination of these effects. We usually model energy
loss by assuming that one of the energy dissipating forces dominates over the others.

5.6 Equations of Motion


Most analysis of dynamical systems discussed earlier involved an instantaneous analysis.
For example, given the force on a body or systems of bodies, we are asked to find the
accelerations, or vice versa. The impulse-momentum and work-energy relationships, which
relate positions and velocities at two different instances in time or at two different locations,
are obtained by integrating the force and moment balances.
Of interest is the capability to describe the system behavior continuously in time. This
is made possible by casting the describing equations as equations of motion (e.o.m.), which
are sets of differential equations, and subsequently integrating the equations of motion.
This section discusses the process of describing the system behavior in terms of equations
of motion.
Equations of motion have the following properties:
• There are as many equations of motion as there are degrees of freedom. For nonholo-
nomic systems, that is, systems subjected to nonholonomic constraints, the number of
e.o.m. is the same as the number of independent velocity variables.
• Equations of motion are in the form of second-order ordinary differential equations (par-
tial differential equations for elastic systems, such as beams) in terms of displacement
variables (or first-order differential equations in terms of velocity variables).
• Equations of motion consist of the motion variables and their time derivatives, multiplied
by known coefficients, as well as expressions describing the forcing. They do not contain
any reaction forces or constraint forces, which are revealed when the components of the
dynamical system are separated and analyzed individually.
The following procedure is used to obtain the equations of motion of a dynamical system:
1. Identify the number of degrees of freedom and select a set of unambiguous and indepen-
dent set of motion variables. Designate the positive directions of the motion variables
and the reference points from where these variables are measured.
2. Separate (if necessary) the system into its components, displace all components from
their initial positions in a consistent way (in their positive directions) and draw free-body
diagrams for each component.
3. Write the force and moment balance equations. The number of force and moment balance
equations are, in general, larger than the number of degrees of freedom. This is because
after separating the different components, we need to consider the reaction (constraint)
forces that act between the components.
4. Manipulate the force and moment balances to eliminate all the constraint forces and
moments. To this end, we make use of the kinematics of the system, as well as constraint
equations, which are geometric descriptions of the constraints. Doing so, we obtain a set
of differential equations, which are in terms of the motion variables and external forces.
Kinetics Applications 185

A set of force and moment balances that are free of internal forces (constraint forces)
and other unknowns, and where the only unknowns are the motion variables, are the
equations of motion.
It is customary to write the differential equations of motion so that the motion variables
and their derivatives are on the left, ordered from the highest derivative term, and all terms
not involving the motion variables are on the right. Given the external excitations acting
on a system and the initial conditions, we can solve the differential equations of motion.
The solution to the differential equations describes the response of the system.
Summing forces and moments is not the only way we can obtain equations of motion.
There are analytical techniques, such as Lagrange’s equations. These approaches are dis-
cussed in Chapter 8. Also, there are cases when we deliberately leave constraint forces in the
describing equations. This procedure is followed in special cases where it is advantageous
to do so, as will be discussed in Chapter 8. Additional examples of obtaining the equations
of motion can be found in Chapter 6 for single-degree-of-freedom systems and in Chapter
7 for multi-d.o.f. systems.

Example 5.5
Consider the rod of length L and mass m connected to a fixed support by means of a pin
joint, shown in Figure 5.17a. Two axial springs oppose the motion of the rod. Derive the
equation of motion. Neglect deformation of the springs in the vertical direction.

A A
g k1 k2
k1 k2

G
mg L/2

O Oh O
Ov

FIGURE 5.17
Rod supported by two springs. a) Displaced position, b) free-body diagram.

The system has one degree of freedom. A suitable motion variable is θ measured from the
upright position of the rod. The free-body diagram is given in Figure 5.17b. The deflection
of both springs is ∆ = L sin θ. Summing moments about the pin joint gives
X L
MO = IO θ̈ = −k1 L sin θ (L cos θ) − k2 L sin θ (L cos θ) + mg sin θ [a]
2
where IO = 13 mL2 is the mass moment of inertia about the pin joint. Rearranging terms,
the equation of motion can be written as
L
IO θ̈ + kL2 sin θ cos θ − mg sin θ = 0 [b]
2
in which k = k1 + k2 is the equivalent spring constant. Note that in this problem the springs
are not in series.
186 Applied Dynamics

Example 5.6
Gantry cranes, such as the one in Figure 5.18, are used worldwide to transport heavy loads
from one vehicle to another. One of their most important uses is in the loading and unloading
of container ships. The speed with which a container can be moved from ship to truck is of
utmost interest as faster transfer times translate to higher efficiency and lower cost.

a) x b)
Treated as known
F x
O O
x

 z
Lcos
L
T
 
m

Lsin mg

FIGURE 5.18
Gantry crane. a) Model, b) free-body diagram.

A simple model of a gantry crane consists of a pendulum attached to a moving trolley


(cart). The crane operator moves the trolley and while doing so inadvertently excites the
pendulum, which begins to sway. This sway motion of the pendulum is undesirable, as it
gives the payload unneeded motions, reducing clearances and creating safety hazards. It is
desirable for the operator to move the trolley as rapidly as possible while minimizing the
sway of the pendulum.
A number of control strategies have been proposed for moving gantry cranes. To analyze
how a control law for the crane works, we need to first derive the equation of motion of
the crane. The motion of the trolley, which is denoted by x, is treated as a known quantity,
so that the system has one d.o.f. A suitable motion variable is θ. The free-body diagram
is shown in Figure 5.18b. Summing forces for the pendulum in the horizontal and vertical
directions gives
+
X X
→ F = max +↓ F = maz [a]
where the forces are
+
X X
→ F = −T sin θ +↓ F = mg − T cos θ [b]

and the accelerations are


d2 d  
ax = (x + L sin θ) = ẍ + L θ̇ cos θ = ẍ + Lθ̈ cos θ − Lθ̇2 sin θ [c]
dt2 dt
d2 d  
az = L cos θ = − L θ̇ sin θ = −Lθ̈ sin θ − Lθ̇2 cos θ [d]
dt2 dt
The two force balance equations become
 
m ẍ + Lθ̈ cos θ − Lθ̇2 sin θ = −T sin θ [e]
Kinetics Applications 187
 
m −Lθ̈ sin θ − Lθ̇2 cos θ = mg − T cos θ [f ]

The equation of motion is obtained by multiplying Equation [e] by cos θ and subtracting
from it Equation [f] multiplied by sin θ, which eliminates the tension T and gives

mẍ cos θ + mLθ̈ + mg sin θ = 0 [g]

The equation of motion does not have the tension T in it, as T is an unknown quantity
until the problem is solved. Dividing the above equation by mL and taking the ẍ term to
the right (as the motion of the trolley x is treated as a known quantity and it is not a
variable) puts the equation of motion in standard form as

g ẍ
θ̈ + sin θ = cos θ [h]
L L
Next, consider small motions and linearize the equation of motion about the operating
point θ = 0. To this end, use of the small angle approximations sin θ ≈ θ and cos θ ≈ 1 gives
g ẍ
θ̈ + θ = [i]
L L
The above equation indicates that the acceleration of the trolley affects the motion of
the crane. For crane operators, the goal is to achieve rapid motion of the trolley while not
exciting the sway of the crane as little as possible. One approach is to accelerate the trolley
until a desired speed is reached, move the crane at this desired speed and then decelerate
using the inverse acceleration profile. This way, ẍ will be zero most of the time and there
will be less excitation of the crane. Such an acceleration profile is shown in Figure 5.19.

x(t)
a
tf – t0 tf
t0 t
–a

FIGURE 5.19
Acceleration profile for trolley.

It should be reiterated that the motion of the trolley is treated as a known quantity here
and as a result there is one degree of freedom and one e.o.m. Problem 5.27 considers the
motion of the trolley as a motion variable, which results in a two d.o.f. system.

Example 5.7
Consider the slider mechanism in Figure 5.20a. The system has one degree of freedom. Even
though two angles, θ1 and θ2 , are used to describe the position of the end point P , the two
angles are related to each other because the height of point B can be expressed as

hB = L1 sin θ1 = L2 sin θ2 [a]

which acts as a constraint relationship. Next, separate the two rods and draw their free-body
diagrams, as shown in Figure 5.20b. We can write six force and moment balances, three
188 Applied Dynamics
a) b)
By

B
y Bx Bx
L1 L2 L2
x L1
hB
O P By
1 2
Ff
Ox
Oy
FP

FIGURE 5.20
a) Slider mechanism and b) free-body diagram.

for each body. There are five constraint forces, Ox , Oy , Bx , By , and FP . These reaction
forces are five of the unknowns. The friction force Ff is not treated as an unknown, as its
magnitude depends on the normal force FP and the friction coefficient.
The motion variables θ1 and θ2 are the other two unknowns, for a total of seven un-
knowns. To solve for these unknowns, another equation is needed in addition to the six
force and moment balances. That equation is the constraint equation in Equation [a]. As a
result, there are seven equations for seven unknowns. We can combine the seven equations
to eliminate the five constraint forces and one of the motion variables, which results in a
single equation of motion, verifying once again that there are as many e.o.m. as d.o.f.

5.6.1 Deriving the Equation of Motion of Conservative One-Degree-of-


Freedom Systems Using Energy
We have described the procedure for obtaining the equations of motion of a system in the
context of Newtonian mechanics, by means of force and moment balance relationships. We
can also derive the equation of motion of one-degree-of-freedom dynamical systems using
energy principles.
The incremental work-energy expression or the derivative with respect to time of the
work-energy relationship in Equation (5.47) can be written as

dE dWnc dr
= = Fnc · (5.49)
dt dt dt
For a conservative system, Fnc = 0 and we can write

dE d (T + V )
= = 0 (5.50)
dt dt
By manipulation of this expression, we can obtain the equation of motion.

Example 5.8
Consider the rod pendulum in Figure 5.21 of mass m and length L and obtain the equation
of motion by differentiating the system energy.
We select the motion variable of this conservative system as θ and the datum position
for the pendulum as point O. Because the pendulum is rotating about the fixed point O,
Kinetics Applications 189

Datum
O
g L/2

L/2

FIGURE 5.21
Rod pendulum.

the kinetic energy of the pendulum is


1
T = IO θ̇2 [a]
2
where IO is the mass moment of inertia about O. Noting that the mass moment of inertia
of a rod about its center of mass is IG = mL2 /12, we find IO using the parallel axis theorem
as  2
L 1
IO = IG + m = mL2 [b]
2 3
The potential energy is
L
V = −mg cos θ [c]
2
and the total energy becomes
1 L
E = T +V = mL2 θ̇2 − mg cos θ = const. [d]
6 2
Differentiation of the energy gives
dE 1 L
= mL2 θ̇θ̈ + mg sin θθ̇ = 0 [e]
dt 3 2
Division of Equation [e] by θ̇ results in the equation of motion
1 L
mL2 θ̈ + mg sin θ = 0 [f ]
3 2

5.7 Solution of the Equations of Motion


Once the equations of motion are obtained, the next steps involve extraction of information
from them. We can extract information in a different number of ways, as discussed in Section
4.5. Here, we discuss integrating the equations of motion and obtaining the response by
solving the differential equations that the e.o.m. define.
Solution methods for differential equations vary greatly depending on whether the equa-
tions of motion are linear or nonlinear. There are a multitude of solution methods for linear
190 Applied Dynamics

differential equations, such as Laplace transform solutions and homogeneous and particular
solution approach. These methods are discussed in Chapters 6 and 7.
When the equations of motion are nonlinear, we can deal with them directly, or we
can linearize the equations of motion about an operating point. Section 5.8 discusses lin-
earization of equations of motion. Because the solution methods for nonlinear differential
equations are much more limited and also more complex than solution methods for linear
systems, linearizing equations of motion makes sense in a large number of applications.
Figure 5.22 describes the different ways we can obtain the response.

FIGURE 5.22
Solution procedures for equations of motion.

5.8 Linearization, Equilibrium, and Stability


Chapter 1 discussed linearization of an expression or of a differential equation and the
advantages of dealing with linear equations. Such advantages include ease of solution and the
ability to use superposition and deal with large numbers of degrees of freedom. This section
analyzes points about which it makes sense to linearize equations of motion. Dynamical
systems are usually linearized about an operating position that is significant. Two such
positions are
• An equilibrium position;
• The position where all elastic members, such as springs, are not deformed.
Static equilibrium is defined as the state when all of the components making up the
system are at rest. The velocity, acceleration, angular velocity, and angular acceleration of
each component are zero at equilibrium.
When discussing the equilibrium of a system, two important questions come to mind:
• How do we calculate the equilibrium position(s)?
• What happens to a system at equilibrium after it is displaced from equilibrium?
Kinetics Applications 191

5.8.1 Calculating the Equilibrium Position(s)


To calculate the static equilibrium position(s), we take the describing equations and set all
the time derivative terms, as well as time dependent terms in the forcing, equal to zero.
Alternatively, we can use energy principles and set the derivative of the potential energy
with respect to the motion variables equal to zero.
Consider, for example, the pendulum in Figure 5.21. The static equilibrium position is
obtained by setting the time derivatives equal to zero, resulting in the relationship sin θ = 0.
This equation has two solutions, θe = 0 and θe = π. The two equilibrium positions are shown
in Figure 5.23b.

a) b)
O


e = 0 e = 

FIGURE 5.23
Equilibrium positions of pendulum: a) θe = 0, b) θe = π.

Calculation of the equilibrium position(s) of a one-degree-of-freedom system requires


the solution of one equation. It follows that calculation of the equilibrium positions of an
n d.o.f. system requires the solution of n simultaneous equations, which for many cases
cannot be decoupled from each other and solved individually. The pitch and bounce model
of a vehicle in Section 15.6 is a typical example of finding the equilibrium positions of a
multi-degrees-of-freedom system.

5.8.2 Motion in the Vicinity of Equilibrium


Now, consider the second question. What happens to the system when it is displaced from
equilibrium? Three scenarios are possible:
1. The system returns to the equilibrium position and comes to rest. The equilibrium
position is called stable or asymptotically stable.
2. The system hovers around equilibrium without coming to a rest at one point, but it
does not get away from or return to the equilibrium position. This equilibrium position
is referred to as critically stable, merely stable, or neutrally stable.
3. The system moves away from the equilibrium position and it never returns. The equi-
librium position is called unstable. Once the system leaves the unstable equilibrium
position, it moves towards one of the stable equilibrium positions, if a stable equilib-
rium position exists. Otherwise, motion amplitudes grow indefinitely.
There are several approaches that enable us to examine behavior in the vicinity of
an equilibrium position. The simplest is visual.2 The next approach is to linearize the
2 Please be careful when ascertaining stability visually. Looks may be deceiving.
192 Applied Dynamics

equations of motion about equilibrium. As discussed in the previous section, solving linear
differential equations is much simpler than solving nonlinear differential equations. There
also are substantially more methods of solution for linear equations than for nonlinear
equations.
Linearized equations lead to additional stability results. A stability theorem states
If the linearized equations of motion in the neighborhood of equilibrium exhibit sig-
nificant behavior (continuously decaying or growing in amplitude), then the general
motion is governed by this significant behavior.
Consider a single-degree-of-freedom system and write the equation of motion as

ẍ (t) + f (x (t) , ẋ (t)) = 0 (5.51)

in which all of the position and velocity dependent terms are collected in the f (x (t) , ẋ (t))
term. At equilibrium, all the velocity and acceleration terms are zero, and the equilibrium
condition is obtained from

f (xe , 0) = 0 (5.52)

where xe denotes the equilibrium position. Recall that at equilibrium ẋe = 0, ẍe = 0.
Depending on the nature of the function f , there can be more than one equilibrium
position. As discussed in Chapter 1, to find the linearized equations of motion, we define
a local variable  = x − xe and expand each term in the equation of motion about the
equilibrium position. The term ẍ becomes ¨, as ẍe = 0. Similarly, ẋ = , ˙ as ẋe = 0. The
next step is to expand f (x (t) , ẋ (t)) in a Taylor series and to retain the linear terms:

∂f ∂f
f (x, ẋ) ≈ f (xe , 0) +  + ˙ (5.53)
∂x xe ∂ ẋ xe

The first term on the right, f (xe , 0), is zero, by virtue of the definition of the equilibrium
relationship in Equation (5.52). Defining by γ1 and γ2 the partial derivatives evaluated at
the equilibrium position

∂f ∂f
γ1 = γ2 = (5.54)
∂x xe ∂ ẋ xe

f (x, ẋ) can be linearized as

f (x, ẋ) ≈ γ1  + γ2 ˙ (5.55)

Substituting the above relationship into the equation of motion in Equation (5.51) yields
the linearized equation of motion about equilibrium, also known as local behavior

¨ + γ2 ˙ + γ1  = 0 (5.56)

All the coefficients in the above equation are constant. Considering the itemization of
the steps 1–4 in the previous section for obtaining the equations of motion, the following
fifth step is added in order to arrive at a set of linearized equations of motion:
5. Calculate the equilibrium positions and linearize the equations of motion about the
equilibrium positions.
We reiterate here that linearity is not a property of a system but of its range of operation,
so that any linearity assumption should always be monitored and verified.
Kinetics Applications 193

5.8.3 Nature of the Response of the Linearized Equations


To analyze the response of the linearized system, consider a solution in the form  (t) = Eeλt ,
where E is the amplitude and λ denotes the time dependency. Introducing this into the
linearized equation of motion in Equation (5.56) and collecting terms gives

λ2 + γ2 λ + γ1 Eeλt = 0

(5.57)

For this equation to have a nontrivial solution Eeλt cannot be zero, so the relationship
that must be equal to zero is identified as

λ2 + γ2 λ + γ1 = 0 (5.58)

The above equation is known as the characteristic equation. It is a polynomial of order


two in λ. The roots of the characteristic equation are
p
−γ2 ± γ22 − 4γ1
λ = (5.59)
2
The response of the linearized system, or behavior in the neighborhood of equilibrium,
is dictated by the roots of the characteristic equation. The following cases are possible:

1. When γ2 > 0 and γ1 ≥ 0, the roots are real and negative or they are complex with
negative real parts. The response of the linearized system is asymptotically decaying
oscillation or aperiodic decay. The equilibrium position is stable.
2. When γ2 = 0 and γ1 > 0, the roots of the characteristic equation are pure imaginary.
The response is oscillation with constant amplitude and does not exhibit significant
behavior. No conclusions can be drawn and further analysis is needed to determine the
nature of the motion. Such analysis includes qualitative approaches such as the energy
theorems, Lyapunov method, or quantitative techniques such as numerical integration
of the nonlinear equations.
3. For any other combination of γ1 and γ2 , the response is either asymptotic growth or
oscillatory growth. Hence, the equilibrium position is unstable.

Advanced concepts from stability theory are beyond the scope of this text. We will
discuss one important theorem, involving the potential energy. The theorem states
For conservative systems (total energy = kinetic energy + potential energy = con-
stant), if the potential energy has a minimum at the equilibrium position, then the
equilibrium position is critically stable. Otherwise it is unstable.
Figure 5.24 depicts the concept of stability of the local equilibrium points for a point mass
subjected to gravitational forces. Figure 5.25 summarizes the stability results. Chapters 6
and 7 deal with the response of single and multi-degrees-of-freedom systems.

5.8.4 Equations of Motion of Linear Systems about Equilibrium


It turns out that it makes sense to write the equations of motion of linear systems about
equilibrium, as well. Such an approach simplifies the equations of motion by eliminating
forces of constant magnitude, such as weight, making it easier to obtain the response. The
procedure is best explained by an example, which will soon follow.
194 Applied Dynamics

g
Unstable

Unstable

Unstable

Stable Stable

FIGURE 5.24
Stability of equilibrium points for gravitational forces.

FIGURE 5.25
Summary of stability properties of equilibrium positions.

Example 5.9
Consider Example 5.5 and the rod of length L and mass m in Figure 5.17a. Two axial springs
oppose the motion of the rod. Find the equilibrium positions and analyze the stability of
the equilibrium positions.
From Example 5.5, the equation of motion has the form
L
IO θ̈ + kL2 sin θ cos θ − mg sin θ = 0 [a]
2
in which k = k1 + k2 is the equivalent spring constant. Setting θ̈ = 0 gives the equilibrium
equation for the angle θe
 
2 L 2 L
kL sin θe cos θe − mg sin θe = sin θe kL cos θe − mg = 0 [b]
2 2
The equilibrium equation has the following solutions:
mg  mg 
sin θe = 0 =⇒ θe = 0, π cos θe = =⇒ θe = ± cos−1 [c]
2kL 2kL
mg mg
The second equilibrium position, θe = ± cos−1

2kL exists when 2kL ≤ 1. Next, let us
Kinetics Applications 195

linearize the equations of motion. There are no velocity dependent terms, so γ2 = 0. The
stability of the system is dictated by γ1 . When γ1 < 0, the equilibrium position is unstable
and when γ1 > 0, we cannot draw any conclusions, which necessitates looking at the value
of the potential energy at equilibrium, as the system at hand is conservative (all forces that
do work, the spring forces and gravity, are conservative forces). The potential energy is

1 2 L
V = k (L sin θ) + mg cos θ [d]
2 2
whose derivative with respect to θ is
dV L
= kL2 sin θ cos θ − mg sin θ [e]
dθ 2
The terms in Equation [e] are recognized as the last two terms in Equation [a]. The
relationship dV /dθ = 0 gives the equilibrium equation for one-degree-of-freedom systems.
The second derivative of the potential energy becomes

d2 V L
= kL2 cos2 θ − sin2 θ − mg cos θ

[f ]
dθ2 2
For motion in the vicinity of the equilibrium position θe = 0, indicating that the rod
is upright, using the small angles assumption of sin θ ≈ θ, cos θ ≈ 1 leads to the linearized
equation of motion  mg 
IO θ̈ + kL2 1 − θ = 0 [g]
2kL
mg
so that γ1 > 0 when 2kL < 1.
Evaluating the second derivative of the potential energy at θe = 0, we obtain

d2 V

mg  mg 
2
= kL2 − = kL2 1 − [h]
dθ θe =0
L 2kL

which, of course, is the coefficient of θ in Equation [g]. Hence, the equilibrium position
mg mg
θe = 0 is stable when 2kL < 1 and unstable when 2kL > 1.
At the equilibrium position θe = π, the rod is hanging below the pivot, so we expect it
to be stable. Indeed, the second derivative of the potential energy is

d2 V

mgL
= kL2 + > 0 [i]
dθ2 θe =π 2

so that this equilibrium position is stable for all values of the spring constant. It should be
noted, however, that in this position, the springs are stretched quite a bit. This equilibrium
position is mathematically feasible, but physically not realistic.
Let us evaluate the sign of
 the second derivative of the potential energy at the equilibrium
mg mg
position θe = ± cos−1 2kL , which exists only when 2kL < 1. Consider the positive value
mg
first, introduce the notation a = cos θe = 2kL , and write

1 d2 V

mg
2 2
= cos2 θe − sin2 θe − cos θe
kL dθ θe =cos−1 ( mg )
2kL
2kL

= a2 − 1 − a2 − a2 = − 1 − a2 < 0
 
[j]
so that this equilibrium position is unstable. It turns out that the negative value for cos θe
gives the same result. Table 5.2 summarizes the stability results.
196 Applied Dynamics

TABLE 5.2
Stability results

mg mg
Equilibrium Position 2kL <1 2kL >1
θe = 0 Stable Unstable
θe = π Stable Stable
mg
θe = ± cos−1

2kL Unstable Not applicable

Example 5.10—Static Tire Balancing


An interesting use of the stability theorem associated with potential energy is static balanc-
ing of tires. Static balancing ensures that the center of mass of the tire lies on the rotation
axis of the tire. Consider a disk with an imbalance, say, with a concentrated mass at point
B, as shown in Figure 5.26a. If the disk is mounted vertically and given a spin, as shown
in Figure 5.26b, the disk will eventually come to rest in the position where the potential
energy has its lowest value, at which point the mass imbalance will be directly below the
axle.

a) b) g
A

C M C
M

B
m m B

z (Vertical)

FIGURE 5.26
a) Disk with a mass imbalance, b) disk hanging from an axle.

If the disk does not have an imbalance, the orientation at which it will come to a rest
will be random, as in a roulette wheel. The balancer spins the wheel a few times. When the
wheel comes to rest, the mechanic marks the bottom, as in Figure 5.26b, which identifies
the location of the imbalance. The disk is balanced by placing a counterweight along the
line from C to A. In a tire, the counterweight is attached to the rim.
This balancing method is the simplest and least accurate way of wheel balancing. The
method determines that there is an imbalance and locates the axis of the imbalance (line
BC in Figure 5.26b). But it doesn’t identify the magnitude of the imbalance. The counter-
balancing is done by trial and error. This type of balancing is usually done for smaller wheels
that are symmetric about the center plane (Figure 5.26a), such as a bicycle or motorcycle
wheel. For car tires, a more accurate method is needed, such as bubble balancing.
In bubble balancing, the tire is placed on a platform that is supported by springs. The
springs compress and an equilibrium position is reached. When the tire is not properly
balanced, the equilibrium position of the tire is at an angle. This angle is measured (usually
Kinetics Applications 197

visually by means of the location of the bubble on a level) and a counterweight is added to
bring the tire to a level position.

a) b)
M m P
D G E Undeformed

a zD D
 P E zE
k k x
G
L/2 L/2
Mg mg
kzD kzE z
a
L/2

FIGURE 5.27
a) Beam with an imbalance, b) free-body diagram.

Bubble balancing can be illustrated in two dimensions by considering the beam of mass
M and length L in Figure 5.27a. The beam has an imbalance of mass m at point P . Select the
motion variables as the spring deflections zD and zE (positive downwards). The free-body
diagram is shown in Figure 5.27b. Assume that the imbalance weight is small with respect
to the beam, which usually is the case, so that the angle θ is small and sin θ ≈ tan θ = θ,
cos θ ≈ 1. Summing moments about D and E at equilibrium gives
 
L 1 L
MD = −M g − mga + kLzE = 0 =⇒ zE = M g + mga [a]
2 kL 2
 
L 1 L
ME = M g + mg (L − a) − kLzD = 0 =⇒ zD = M g + mg (L − a) [b]
2 kL 2
The angle with which the beam rests at equilibrium is calculated from the geometry.
For small angles, the equilibrium position, denoted by θe , becomes
 
zE − zD mg (L − 2a) mg 2a
θe ≈ tan θe = = = 1 − [c]
L kL2 kL L
The next step is to design the counterbalance. The angle θe is measured by the balancer.
The mass and location of the imbalance are not known. The balancer needs to select the
mass of the counterweight and decide where to place it. Considering a counterbalance of
mass mc at a distance b from point D, as shown in Figure 5.28, we can use Equation [c] to
determine the new equilibrium position in the presence of the counterweight as
     
0 mg 2a mc g 2b mc g 2b
θ ≈ 1− + 1− = θe + 1− [d]
kL L kL L kL L

The desired value for the new equilibrium angle is θ0 = 0. Substituting this value into
Equation [d], we can solve for mc and b. There is an infinite number of solutions, as two
parameters need to be selected, the counterweight mass mc and location of counterweight
b, to satisfy one relationship. For a tire, the counterbalance is usually placed at the rim, so
b is specified. Solving for the mass of the counterweight gives
kL 1
mc = −θe [e]
g 1 − 2b
L
198 Applied Dynamics

D B
mc  P
m
b E
G
kzD
Rim
location Mg
mg
kzE

FIGURE 5.28
Beam with counterbalance added.

Static balancing does not take into consideration that the wheel is a three-dimensional
object and that the imbalance and the counterweight may end up being located on different
planes perpendicular to the axis of rotation. This affects the inertia matrix of the tire-rim
assembly, as we will see in Chapters 10 and 11. Balancers use time-proven techniques, such
as placing counterweights on both sides of the rim, in order to minimize dynamic imbalances.

Example 5.11—Bicycle Balancing

a) b)
G G
Turn axis
mg
L L y
h
z

FL
N

FIGURE 5.29
a) Rear view of bicycle taking a turn, b) free-body diagram.

Bicycle riders lean into the turn when taking a turn. Leaning into the turn shifts the
center of mass and stabilizes the bicycle. Consider the rear view of a cyclist (more like a
unicycle) in Figure 5.29a taking a turn of radius ρ at constant speed v. Determine the value
of the lean angle that will keep the bicycle stable.
The free-body diagram is shown in Figure 5.29b. The xyz coordinates rotate with the
bicycle. The only motion of the bicycle with respect to the xyz coordinates is the lean, θL .
We consider the case where the speed v is constant, so the acceleration of the bicycle is the
normal acceleration an = mv 2 /ρ, acting towards the turn center. It follows that the lean
angle θL is also constant. This situation is a special form of equilibrium.
Three forces act on the bicycle: its weight, the normal force N , and the lateral force
Kinetics Applications 199

generated by the tires. Summing forces for the bicycle in the yz plane gives
X v2
F = −man j = −m j = mgk − N k − FL j [a]
ρ
where FL is the lateral force at the tires. Equating components in the y and z directions
gives
v2
FL = man = m N = mg [b]
ρ
Summing moments (about x axis) about the center of mass, we obtain
MG = 0 = N h sin θL − FL h cos θL [c]
Solution of the above equation gives the lean angle for equilibrium as
     2
FL an v
θL = tan−1 = tan−1 = tan−1 [d]
N g ρg
For small lean angles, the above equation can be approximated as θL ≈ an /g. As can
be expected, a faster (or sharper) turn by the cyclist requires a larger lean angle. It is
interesting to note that the lean angle is independent of the center of mass height. Chapter
11 discusses this type of motion in more detail.

Example 5.12
The mass-spring-damper system in Figure 5.30a is moving without friction on an incline.
Obtain the equation of motion using the variable x, which is measured from the undeformed
position of the spring. Next, obtain the equation of motion using the variable , which is
measured from static equilibrium.

a) b)
c
g x cx
kx mg
k m

FIGURE 5.30
a) Mass on an incline. b) Free-body diagram.

The free-body diagram is given in Figure 5.30b. Summing forces along the incline gives
X
& F = ma =⇒ mg sin β − kx − cẋ = mẍ [a]

and the resulting equation of motion is


mẍ + cẋ + kx = mg sin β [b]
We obtain the equilibrium position by setting ẋ = 0, ẍ = 0, with the result
kxe = mg sin β [c]
200 Applied Dynamics

so that the equilibrium position is xe = mg sin β/k and ẋe = 0, ẍe = 0.


Next, introduce the variable  = x − xe , or x = xe + . It follows that ẋ = ,
˙ ẍ = ¨.
Introducing these values to the equation of motion in Equation [b], we obtain
 
mg sin β
¨ + c˙ + k +  = mg sin β [d]
k
We see that there is an mg sin β term on both sides of the equation above. These two
terms cancel and the equation of motion, in terms of  as the motion variable, becomes

¨ + c˙ + k = 0 [e]

This equation is simpler than Equation [b], the equation of motion in terms of x. Further,
its solution is also simpler.

5.9 Motion in the Vicinity of the Earth


In Chapters 2 and 3 we developed relative velocity and acceleration expressions and we
analyzed the kinematics of motion with respect to the Earth. The analysis is extended here
to the kinetics and the effects of the rotation of the Earth on the equations of motion. While
the developments in this section involve rotation of the Earth, the analysis is applicable to
any type of motion with respect to a rotating reference frame.

y "
 P m ! z
"g'
B

FIGURE 5.31
Particle in vicinity of earth.

Figure 5.31 describes the geometry (not to scale) and free-body diagram. Point B is on
the surface of the Earth and the position vector of a point mass P with respect to B is
r = rP/B = xi + yj + zk. The relative velocity and acceleration terms become

vrel = ẋi + ẏj + żk arel = ẍi + ÿj + z̈k (5.60)

and, from Equation (3.4), the acceleration of point P is

aP = a = aB + ω × (ω
ω × r) + 2ω
ω × vrel + arel (5.61)

where aB = ω × (ω
ω × rB ).
The force balance for the mass becomes

ma = F + F0g (5.62)
Kinetics Applications 201

where F0g is the force of gravity and F denotes the sum of all other external forces. In-
troducing Equation (5.61) to Equation (5.62), the force balance in terms of the relative
coordinates x, y, z (or r) is obtained as

ω × vrel + mω
marel + 2mω ω × r) = F + F0g − mω
ω × (ω ω × (ω
ω × rB ) (5.63)

As discussed in Sections 3.2 and 4.7, the centrifugal force due to the acceleration of point
B, −maB = −mω ω × (ω
ω × rB ) is included in the gravitational force. The term Fg is defined
as the augmented gravitational force and approximated as

Fg = F0g − mω
ω × (ω
ω × rB ) ≈ −mgk (5.64)

This simplification is possible because the radius3 of the Earth is nearly constant, so we
can write rB = re k, where re is the radius. Recalling that the angular velocity vector is
ω = Ω (sin λk + cos λi), where λ is the latitude, the centripetal acceleration of B becomes

ω × rB ) = Ω2 re − cos2 λk + sin λ cos λi



ω × (ω (5.65)

The component in the z direction changes the value of the gravitational constant and
the component in the x direction is negligible compared to external forces that can act in
that direction. Hence, the equations of motion become

ω × vrel + mω
marel + 2mω ω × (ω
ω × r) = F − mgk (5.66)

Writing the external force as F = Fx i + Fy j + Fz k and dividing the above equation by


the mass m, the equations of motion can be expressed as
Fx
ẍ − 2Ωẏ sin λ − Ω2 x sin2 λ − z sin λ cos λ

=
m

Fy
ÿ + 2Ω (ẋ sin λ − ż cos λ) − Ω2 y =
m

Fz
z̈ + 2Ωẏ cos λ + Ω2 −z cos2 λ + x sin λ cos λ

= −g (5.67)
m
The equations of motion can also be written as

marel = Feff − mgk (5.68)

where the effective force Feff is the force felt by a body due to the rotation of the Earth

Feff = F + FCoriolis + Fcentrifugal (5.69)

and where the Coriolis and centrifugal forces are

FCoriolis = −2mω
ω × vrel Fcentrifugal = −mω
ω × (ω
ω × r) (5.70)

The centrifugal force −mωω × (ω ω × r) is very small and is usually neglected for bodies
close to the Earth. The Coriolis force has a distinctively more pronounced effect. While the
magnitude of the Coriolis force is also small, its direction is always perpendicular to the
3 The radius of the Earth varies between 6371 and 6378 km.
202 Applied Dynamics

a) z (Height) g b)

Air flow
Plane of Low pressure
vrel , ( .k) k

sin k

Component of y (West)
–2 x vrel Northern hemisphere
in xy plane vrel

x (North)

FIGURE 5.32
a) Coriolis force in northern hemisphere (vrel is in xy plane). b) Hurricane formation.

velocity; thus, this force causes a change in direction. In the absence of significant forces
perpendicular to the velocity, the effect of the Coriolis force builds over time.
The component of the Coriolis force due to the z component of the angular velocity in
the Earth’s northern hemisphere is depicted in Figure 5.32a. The Coriolis force causes a
particle to always veer to the right in the northern hemisphere. So does the component of
the angular velocity in the x direction. The Coriolis effect in the vertical direction is usually
ignored, as it is dwarfed by the force of gravity.
The Coriolis effect is used to account for several kinds of phenomena. Pertinent to
weather analysis, the motion of air masses is affected by the Coriolis force. For example, in
the northern hemisphere the spin of air masses in hurricanes is counterclockwise, as shown
in Figure 5.32b for when vrel is in the xy plane. A hurricane occurs when a low pressure
center attracts air particles inward with large speeds. The motion of projectiles, such as
missiles, is also affected by Coriolis forces, as is the whirl of water going down the sink.

Example 5.13
Consider a 20 m long bowling alley in Rio de Janeiro. A ball is launched with a speed of 7
m/s. Calculate the Coriolis deflection and the direction of the deflection.
The latitude of Rio is −23.5◦ . Without loss of generality, assume that the alley is along
the x axis. Neglecting any motion in the vertical (z direction) and centrifugal effects, from
Equation (5.67) the two equations of motion become

ẍ − 2Ωẏ sin λ = 0 ÿ + 2Ωẋ sin λ = 0 [a]

Assuming that the ball travels along the alley with constant speed, we can deal only with
the equation of motion in the y direction, which becomes a constant acceleration problem
2
ÿ = a = −2Ωẋ sin λ = −2 × (7.292 × 10−5 ) × 20 sin(−23.5◦ ) = 4.071 × 10−4 m/s [b]

The acceleration is in the positive y direction. In a right-handed coordinate system, if


x is forward and z is up, then y is to the left of x. As expected, the ball veers to the left.
Kinetics Applications 203

Recall that the motion is taking place in the southern hemisphere. The time of travel is
7/3 = 2.857 s, so the Coriolis deflection when the ball reaches the end of the alley is
1 2
y = at = 0.5 × 4.071 × 10−4 × 2.8572 = 1.660 × 10−3 m = 1.660 mm [c]
2
Air resistance and the lack of a perfectly smooth surface cause deflections that are much
larger than the Coriolis deflection. It takes time for the Coriolis effect to become significant.

5.10 Collisions
This section considers collision, which is a special case of contact between two bodies,
involving large velocities of one body relative to the other before contact. Collisions give rise
to impulsive forces and their study involves impulse-momentum and work-energy principles.
Of interest are the velocities and angular velocities of the colliding bodies immediately after
impact and the energy lost during the collision.
We need to be aware of the assumptions made for analyzing collisions:
• The collision takes place over a very short period of time, making it possible to treat
the forces generated as impulsive.
• There is no (or very little) damage to the colliding bodies.

These two assumptions are contradictory. There have to be high velocities involved for
rapid collisions and there have to be low velocities for no damage. We must always evaluate
the impact conditions to ensure that the assumptions are applicable.
Recall that, as we saw in Chapter 3, contact between two bodies involves a common
normal and a common tangent. In collisions, the common normal is also referred to as the
line of impact. Our study of impact here will first consider particles and then extend the
concept to rigid bodies.

5.10.1 Collisions of Particles


Consider two particles, of mass m1 and m2 , that are approaching each other with velocities
of v1 and v2 , respectively, as shown in Figure 5.33. The common normal is along the line
that joins the two particles and the common tangent is perpendicular to that line. The unit
vector along the line of impact (common normal) is denoted by n. The components of the
velocities along the line of impact are

v1 = v 1 · n v2 = v2 · n (5.71)

For collisions of particles (or small smooth rigid bodies) we commonly assume that the
impact forces act only along the line of impact (common normal) and no forces act along the
common tangent. The impulsive forces are generated based on Poisson’s hypothesis, which
assumes that there is a small amount of compliance in the colliding bodies and that impact
takes place in two stages. In the first stage, called the period of compression (Figure 5.34),
the bodies compress each other until the relative velocity between the colliding particles
becomes zero along the line of impact so both masses have the same velocity vc . In the next
stage, called the period of restitution, the bodies begin to separate from each other and
they regain their original shapes.
204 Applied Dynamics

!" !- %&'()& #" n $" *'+&)

F^ m-
!, m,
m- F^ m,
m- "-
m,
",
n n

FIGURE 5.33
Approaching and colliding particles: a) approach, b) free-body diagram during impact, c)
separation.

1 m1 m2 3
Approach Restitution
v1 v2 F^r F^r

2 4
Compression Separation
F^c F^c u1 u2

FIGURE 5.34
The four stages of impact.

The ratio of the strength of the two impulses is denoted by the coefficient of restitution
e. Denoting the impulsive forces associated with the stages of impact as F̂c and F̂r , where
c and r stand for compression and restitution, respectively, the coefficient of restitution is
commonly defined as

F̂r
e = (5.72)
F̂c
The restitution forces are smaller than the compression forces because of hysteresis,
which was discussed in Section 5.5.7. The collision leads to compression of the masses
beyond the linear range, resulting in different loading and unloading curves. The energy
loss during restitution translates to a smaller impact force during that period.
Denoting the total strength of the impact by F̂ = F̂c + F̂r , the impulse-momentum
relationships become

Mass 1: Compression m1 v1 − F̂c = m1 vc Restitution m1 vc − F̂r = m1 u1


Mass 2: Compression m2 v2 + F̂c = m2 vc Restitution m2 vc + F̂r = m2 u2 (5.73)

Adding the four expressions in the above equation gives the impulse-momentum rela-
tionship

m1 v1 + m2 v2 = m1 u1 + m2 u2 (5.74)

The above equation states that the combined linear momentum of the colliding particles
along the line of impact is conserved. Considering the two masses as a system, there are no
Kinetics Applications 205

forces external to the system. Linear momentum is also conserved perpendicular to the line
of impact, as we assume that no forces act in that direction during impact.
By using the above equations, we can also relate the velocity components (along the line
of impact) before impact to those after impact. The results can be shown as

u2 − u1 = −e (v2 − v1 ) (5.75)

so that the relative velocity of separation is related to the relative velocity of approach by
the coefficient of restitution.
The coefficient of restitution e is a quantity in the range 0 ≤ e ≤ 1 that is dependent
on the material properties of the colliding bodies, the hysteresis effect discussed in Sec-
tion 5.5.7, and internal damping, as well as on the relative speed of collision. The coefficient
of restitution becomes smaller as the relative speed before collision becomes larger.
The special case of e = 1 is known as perfectly elastic impact. In this case, the strength
of the impact is the same in the compression and restitution stages, and there is no energy
loss. The case of e = 0 is known as perfectly plastic impact. In a perfectly plastic impact, the
colliding bodies do not separate from each other immediately after impact and energy loss
is the largest. All of the kinetic energy associated with the relative motion of the colliding
bodies is dissipated by the impact forces.
Equations (5.74) and (5.75) can be combined to obtain the velocities immediately after
impact as
1 1
u1 = [(m1 − em2 ) v1 + am2 v2 ] u2 = [(m2 − em1 ) v2 + am1 v1 ] (5.76)
m m
where m = m1 + m2 is the total mass of the colliding bodies and a = 1 + e. From the above
equation, when e = 0, the velocities after impact are the same, u1 = u2 , confirming that
there is no relative velocity between the impacting bodies after plastic impact.
The change in energy before and after impact is
1  1
m1 v12 − u21 + m2 v22 − u22

∆T = (5.77)
2 2
Substituting Equation (5.76) in the above equation, we can describe the change in energy
in terms of the coefficient of restitution.

Example 5.14—The Immaculate Reception


It’s 1972. Twenty-two seconds remain in an NFL playoff game between the Pittsburgh
Steelers and the Oakland Raiders. Oakland is leading by a point and Pittsburgh has the
ball on their own 40 yard line with 4th down and 10 and no timeouts. Pittsburgh quarterback
Terry Bradshaw scrambles, evades a few defenders, and throws a pass. The Oakland safety
(defender) rushes towards the Pittsburgh receiver and makes contact just as the ball arrives.
The football hits the safety and bounces into the arms of Pittsburgh running back Franco
Harris, who runs the ball into the end zone, winning the game for Pittsburgh.4 Figure 5.35
depicts the immaculate reception.
The play in itself was remarkable, thus earning its famous name. But there was a twist.
According to the NFL rules at the time, had the football bounced off the Pittsburgh receiver
the ball could be caught but it could not be advanced. Oakland vigorously protested the
call, claiming that the football hit the Pittsburgh player and not the Oakland defender.
The referees did not change the call. Replays of the play were not conclusive.5 However,
4 You can watch actual footage of this play by going to a public domain video site.
5 In 1979, a video surfaced showing clearly that the ball bounced off the Oakland defender.
206 Applied Dynamics

&'()' %$

*(++
#
$,-./0
"
#$
!"
#-.

FIGURE 5.35
The immaculate reception (QB: quarterback, R: receiver, D: defender, RB: running back,
TD: touchdown).

an analysis of the collision indicated that the ball most likely bounced off the Oakland
defender.
Let us assume that the defender and receiver both have the same weight, say 179 lb,
and the football weighs 1 lb. According to researchers at Carnegie-Mellon University, the
ball was thrown with a speed of 55 ft/sec. Let us assume a coefficient of restitution between
the players and the ball of e = 0.3. Also, assume that the receiver was running forward with
a speed of 7.5 mph (11 ft/sec) and the defender was running toward the receiver with the
same speed. Using Equation (5.76) and denoting by m1 the receiver and m2 the ball, the
speed of the ball after impact with the receiver (note we substitute weight instead of mass
in the calculations) becomes
1
u2receiver = [(1 − 0.3 × 179) 55 + 1.3 × 179 × 11] = −1.88 ft/sec [a]
180
and, for the case of impact with the defender, the post-impact speed of the ball is
1
u2defender = [(1 − 0.3 × 179) 55 − 1.3 × 179 × 11] = −30.3 ft/sec [b]
180
There is quite a difference in the velocities. From the replay video, the speed of the
football after impact was estimated as 30 ft/sec. So, the call on the field was correct and
the Steelers rightfully won the game.

Example 5.15—Accident Reconstruction


An important application of energy and momentum principles is accident reconstruction
and determination of pre-accident positions and velocities, as well as forces of impact. One
way of modeling crash resistance is by considering the vehicle as mass-spring systems, with
the crush stiffness 6 as the spring constant. Using crush stiffness as a parameter eliminates
the need to know the duration of impact for the purpose of calculating forces experienced
during the collision.
Two vehicles collide. Vehicle A has a weight of WA = 2600 lb and its crush stiffness
is kA = 8500 lb/in. Vehicle B has a weight of WB = 5000 lb and has a stiffness of kB =
6 Assuming a linear spring model for the crash resistance of a vehicle turns out to be remarkably accurate.
Kinetics Applications 207

5400 lb/in. Vehicle A is stationary when it is rear-ended by vehicle B. The vehicles are
mangled and they skid together to a stop. Measurements after the accident show that after
the crash the two mangled vehicles skidded 20 ft before coming to a stop. The sliding
friction coefficient between the two crushed vehicles and the road is estimated as µk = 0.75.
Calculate the impact forces and accelerations experienced by the vehicles, as well as the
speed of vehicle B immediately before the crash.

!" #"
kB kA
mB mA mB mA

vA $%& v
vB

FIGURE 5.36
Colliding vehicles: a) immediately before impact, b) immediately after impact.

Figure 5.36 shows the schematic of the impact. The two masses come together, they
crush each other (compress the springs), and after the maximum crush occurs (max. com-
pression of the springs), they slide together. Denoting the velocity of the mangled vehicles
immediately after impact by v, the linear momentum relationship during impact is

mA vA + mB vB = (mA + mB ) v [a]

It is known that before impact vA = 0, but the velocity of the moving vehicle before
impact, vB , is not known. After impact, the kinetic energy of the mangled cars is dissipated
by friction as the vehicles slide together, so the energy balance during sliding becomes

T = 0.5 (mA + mB ) v 2 = Ff d = µk (mA + mB ) gd [b]

where T is the kinetic energy immediately after impact, Ff is the friction force, and d = 20
ft is the sliding distance. Solving for the speed v immediately after impact gives
p √
v = 2µk gd = 2 × 0.75 × 32.17 × 20 = 31.07 ft/sec [c]

Introducing this result for the post-impact speed from the above equation into Equation
[a] yields the pre-impact velocity of vehicle B as
mA + mB 5000 + 2600
vB = v = 31.07 = 47.23 ft/sec [d]
mB 5000

mB F F mA
vB

FIGURE 5.37
Free-body diagrams of the colliding vehicles.

We next consider the energy balance during impact. The two springs absorb a portion
208 Applied Dynamics

of the energy. From the study of springs in series, each spring deflects by a different amount
but transmits the same force, as shown in the free-body diagram in Figure 5.37. Denoting
the spring force (impact force) by F and deformations of the springs by xA and xB , the
spring force becomes
F = kA xA = kB xB [e]
so that the energy absorbed by each spring is

1 1 F2 1 1 F2
EA = kA x2A = EB = kB x2B = [f ]
2 2 kA 2 2 kB
The energy balance during the collision can now be written. The initial kinetic energy
of vehicle B is dispersed into the potential energies of the springs and to the kinetic energy
of the mangled cars. We express the energy balance as
1 2 1
mB vB = EA + EB + (mA + mB ) v 2 [g]
2 2
or
2 F2 F2
mB vB = + + (mA + mB ) v 2 [h]
kA kB
The only unknown above is the spring force during impact. Solving for it gives
r
kA kB 2 − (m + m ) v 2 ) = 68, 533 lb
F = (mB vB A B [i]
kA + kB
The average accelerations experienced by each vehicle (in g) are
aA F 65, 533 aB F 65, 533
= = = 25.6 = = = 13.3 [j]
g WA 2, 600 g WB 5000
and the amount each vehicle crushes is
F 65, 533 F 65, 533
xA = = = 8.06 in. xB = = = 12.7 in. [k]
kA 8, 500 kB 5, 400
The smaller vehicle experiences less deformation as it is stiffer, but the collision forces
felt by it are much larger. The weight of the larger car is the most significant factor in
determining which vehicle feels the impact more.
The analysis above is based on certain assumptions: the collision takes place along a
straight line, there are no velocity components perpendicular to the line of impact before or
after, the crush resistance of both vehicles behaves like a linear spring, and the two vehicles
slide together after the collision.

5.11 Impact of Rigid Bodies: Simple Solution


The developments of the previous subsection, where it was assumed that nothing happens
perpendicular to the line of impact, are valid when the rotational motion of the colliding
bodies is negligible, components of the velocities perpendicular to the line of impact are
small, and friction between colliding particles is negligible. This assumption loses its validity
for rigid bodies, where we need to consider forces perpendicular to the line of impact, such
as frictional impact forces generated along the common tangent.
Kinetics Applications 209

Impact of rigid bodies is a less mature topic than impact of particles, and there is ongoing
research to model this complex phenomenon. This subsection presents a simple model that
is valid primarily for bodies with smooth contours, such as balls. The next section will
discuss a more advanced model. Because it is easier to visualize, the derivations here are
for a body that makes impact with a fixed surface.
Take a ball and throw it forward with no angular velocity (most passes in basketball,
but not shots at the basket, have this characteristic). As soon as the ball hits the ground,
it acquires an angular velocity. This is because of the impulsive frictional force that is
generated between the ball and the floor, as depicted in Figure 5.38, in addition to the
impulsive normal force.

Before z y After
During impact
=0 R
G vG G
x G vG
mg
C C
C
^ ^
Ff N

FIGURE 5.38
Bouncing basketball acquiring angular velocity.

The impulsive friction force is treated here the same way friction forces are modeled
in Section 4.9. Denoted by F̂f , the friction force opposes velocity or impending velocity.
Since point C was moving to the right before impact, the friction force acts to the left. The
magnitude of the friction force lies in the range

0 ≤ F̂f ≤ µs N̂ (5.78)

where the impulsive normal force is denoted by N̂ and µs is the static coefficient of friction
between the ball and the impact surface. The next step is to write the linear and angular
momentum balances for immediately before and after the impact. The linear momentum
balance along the common tangent (x axis) is
+
X
→ F̂ =⇒ mvGx − F̂f = muGx (5.79)

and the linear momentum balance along the line of impact (the vertical, z axis) is
X
+↑ F̂ =⇒ mvGz + N̂ = muGz (5.80)

The angular momentum balance about the center of mass is


X
 M̂G =⇒ IG Ω + F̂f R = IG Ω0 (5.81)

where Ω is the angular velocity before impact and Ω0 is the angular velocity after impact.
According to the coordinate system selected, a clockwise rotation is positive.
The force and moment balances yield three equations. The unknowns consist of the
two components of the velocity of the ball after impact, uGx and uGz ; the angular velocity
210 Applied Dynamics

Ω0 ; the impulsive normal force N̂ ; and the impulsive friction force F̂f , for a total of five
unknowns. It follows that two more equations are needed to equate the number of unknowns
and equations.
One of these equations comes from the strength of the impact, and it makes use of the
coefficient of restitution. This equation, similar to Equation (5.75), relates the velocities at
the point of contact immediately before and immediately after impact as

uCz = −evCz (5.82)

in which vCz = vC · k is the vertical component of the velocity of the point of contact imme-
diately before impact and uCz is its counterpart immediately after impact. The impulsive
normal force is quantified by the coefficient of restitution e. Noting that the position vector
from G to C is rC/G = −Rk and ω = Ωj, the velocity of C is related to the velocity of the
center of mass by

vC = vG + ω × rC/G = (vGx − RΩ) i + vGz k (5.83)

The second equation comes from the slide or no-slide condition at the contact point.
Similar to the analysis in Section 4.9, there are two possibilities:

1. The contact point C slides during impact. Here, the horizontal velocity of the impact
point C is not known immediately after impact, but the magnitude of the impulsive
friction force is known, and it is related to the normal force by F̂f = µk N̂ , which
becomes the fifth equation.
2. Sliding ends during impact. The horizontal velocity of the impact point immediately
after impact is zero, and the fifth equation becomes uCx = 0. Here, the magnitude of
the friction force is not known and needs to be calculated.
The two scenarios are summarized in Table 5.3.

TABLE 5.3
Conditions for sliding and no sliding at point of impact

Condition Friction Force Ff Horizontal Vel. of C


No sliding at contact point Unknown (0 ≤ F̂f ≤ µs N̂ ) 0
Sliding at contact point µk N̂ Unknown

As discussed earlier, friction problems where we do not know whether there is sliding or
not are solved by assuming that there is sliding (or no sliding) and then checking from the
solution if the assumption makes sense.
Consider a solution based on the no sliding assumption. The assumption is valid if the
impulsive friction force, F̂f , is sufficient to prevent sliding. If the calculated value for the
friction force F̂f is less than its maximum possible value, that is, if F̂f ≤ µs N̂ , then the no
sliding assumption is correct. If it is not, then there is sliding and the problem has to be
solved again, this time by setting F̂f = µk N̂ .
Alternatively, we can assume that there is sliding and calculate the velocities and angular
velocity after impact. The validity of this assumption is checked by comparing the velocity
of the contact point after impact. The horizontal velocity of the contact point, uCx =
uGx − RΩ0 , has to be larger than zero. Otherwise, the assumption is not correct.
Kinetics Applications 211

The five equations that need to be solved are linear equations. The derivation above
assumes that the colliding bodies are smooth, circular in shape, or the line of impact goes
through the centers of mass of the impacting bodies. If not, do not use Equation (5.82), but
rather go back to Poisson’s hypothesis and analyze the compression and restitution stages
separately. We will do that in the next section.

Example 5.16—Bouncing Ball


Consider a bouncing ball. Nondimensionalize and solve the impact equations. Obtain a
condition for sliding of the ball during impact.
From Figure 5.38, the vertical velocity before impact is negative and the positive direc-
tion for the angular velocity is clockwise. Also, the impulsive normal and friction forces are
shown in their correct directions. The impulse momentum relationships for translation are
+
X
→ F̂ =⇒ mvGx − F̂f = muGx [a]

X
+↑ F̂ =⇒ mvGz + N̂ = muGz [b]
The angular impulse-momentum relationship, obtained by the angular momentum bal-
ance about the center of mass, is
X
 ĤG =⇒ IG Ω + F̂f R = IG Ω0 [c]

where Ω is the angular velocity before impact and Ω0 is the angular velocity after impact.
Modeling the ball as a spherical shell, the mass moment of inertia is IG = 2mR2 /3. The
impact condition at point C is
uCz = −evCz [d]
in which e is the coefficient of restitution. The relative velocity expression between the
impact point C and center of mass G (which is valid at all times, for v and for u) is

vC = vG + ω × rC/G = vGx i + vGz k + Ωj × −Rk = (vGx − RΩ) i + vGz k [e]

It follows for a spherical object that the counterpart of Equation [d] in terms of the center
of mass G is
uGz = −evGz [f ]
Equations [a], [b], [c], and [f] need to be solved for the unknowns uGx , uGz , Ω0 , N̂ , and
F̂f . There are four equations and five unknowns. A fifth equation comes from the sliding
condition during impact. Two scenarios are possible:
1. Sliding at point C comes to an end during impact. In this case, uCx = 0 and from
Equation [e]
uGx = RΩ0 [g]
which provides the fifth relationship. Note that in this case the friction force has to be
less than its maximum value, F̂f ≤ µs N̂ .

2. Point C slides during impact. In this case, the horizontal velocity at C immediately after
impact is not known, but we can relate the horizontal impulsive force to the impulsive
normal force and the fifth equation becomes

F̂f = µk N̂ [h]
212 Applied Dynamics

Let us the assume that sliding ends during impact. The five equations to be solved are
Equations [a], [b], [c], [f], and [g]. It is convenient to nondimensionalize the impact equations.
Noting that vGz < 0, we introduce the nondimensional quantities
uGx vG x uGz RΩ
ux = vx = uz = ω =
−vGz −vGz −vGz −vGz

F̂f N̂
F = N = [i]
−mvGz −mvGz
Dividing all impact equations by −mvGz gives the nondimensional impact equations as
2 2
vx − F = u x − 1 + N = uz ω + F = ω0 uz = e ux = ω 0 [j]
3 3
The first, third, and fifth equations above are independent of the second and fourth. In-
troduction of the fifth of Equation [j] into the third gives F = 2 (ux − ω) /3, which when
introduced to the first expression in Equation [j] gives the results
 
3 2 2
ux = ω 0 = vx + ω F = (vx − ω) [k]
5 3 5
Introducing the fourth expression in Equation [j] into the second gives the normal force
as
N = 1+e [l]
To check the validity of the no-sliding assumption, we need to calculate the minimum
amount of friction required to stop the slide of the ball during impact. This analysis is more
descriptive in terms of dimensional variables. The impulsive normal and friction forces
become
2
N̂ = −mvGz N = − (1 + e) mvGz F̂f = −mvGz F = m (vGx − RΩ) [m]
5
Note that vGz , the initial speed in the vertical direction, is negative. For there to be no
slip, the friction force F̂f must be less than the maximum available normal force, µs N̂ , or
F̂f ≤ µs N̂ , so the available friction must be

F̂f 2 vGx − RΩ
µs ≥ = [n]
N̂ 5 (1 + e) vGz

The above result leads to interesting interpretations. If the initial horizontal speed be-
comes higher, more friction is necessary to prevent sliding during impact. Similarly, if the
vertical speed before impact is increased, less friction is needed to prevent sliding. A higher
vertical speed results in a higher normal force and thus a higher friction force. Similarly, if
the impact is more elastic, that is, the coefficient of restitution is higher, this results in a
higher denominator and hence a lower amount of friction needed to prevent slip.
A ball thrown with a topspin (Ω > 0) has a higher horizontal speed after impact than a
ball thrown with a backspin. Tennis courts are rated based on the coefficient of friction of
the court and also on the coefficient of restitution of the balls. A clay court is slow (more
friction) and a grass court is fast (less friction).
Another interesting result from Equation [k] is that when friction is sufficient to prevent
slip during impact, and the ball is thrown with no initial angular velocity, the horizontal
speed of the ball after impact becomes
3
uGx = ux vGz = vG [o]
5 x
Kinetics Applications 213

so that after impact the horizontal speed reduces to 60% of its value immediately before
impact. However, when watching a tennis ball or a basketball as it bounces, we get the
impression that the ball is gaining speed. The bouncing of the ball creates a freeze frame,
so when we observe the ball after impact, we think it is moving faster.
The observations above also provide an explanation as to why basketball players shoot
the ball with a slight backspin. The ball with a backspin will have a smaller forward velocity
than a ball with a forward spin (topspin) after collision with the basket. It is preferable for
the ball to have a smaller velocity after hitting the rim, as this way the ball will have a higher
chance of hanging around the rim and falling into the basket. In addition, a basketball shot
with a backspin provides enhanced stability as well as a smoother trajectory.

Example 5.17—Bouncing Basketball

20 ft/sec

4'

FIGURE 5.39
Bouncing basketball.

A basketball is released with a horizontal speed of 20 ft/sec and no angular velocity, as


shown in Figure 5.39. The basketball is at a height of 4 ft when it is released. Neglecting
the aerodynamics, calculate the friction necessary to prevent slip at impact and the angular
velocity of the basketball immediately after hitting the ground.
Table 5.4 gives regulation basketball dimensions. Organizations also have different ma-
terial and color requirements, as well as restrictions on which brands of balls to use.

TABLE 5.4
Official basketball specifications

Organization Size Circumference Weight Bounce or Pressure


NCAA Men 7 29.5–30 in. 20–22 oz 49–54 in. from 6 ft
NCAAW, WNBA 6 28.5–29 in. 18–20 oz 51–56 in. from 6 ft
NBA 7 29.5 in. 22 oz 7.5–8.5 psi
FIBA 7 749–780 mm 567–650 gm 1.3 m from 1.8 m

Let us take a 21 oz NCAA men’s basketball with a 29.75 in. circumference. For a spherical
shell, the mass moment is IG = 23 mR2 . Consider that the ball bounces up to 52 inches after
it is dropped from a height of 6 ft.
We can calculate the coefficient of restitution from the bounce data. Noting that the
velocities before and after impact are related by Equation (5.75) and, considering impact
with the ground, the vertical velocities before and after impact are related by

uGz = −evGz [a]

The maximum height can be obtained by an energy balance. Equating the kinetic and
214 Applied Dynamics

potential energies gives


1 2
mvG = mgh [b]
2 z

from which the height that the ball reaches is obtained as


2
vG z
h = [c]
2g

Equating Equations [a] and [c] gives a relationship for the height reached after impact
as
u2Gz e2 vG
2
z
hafter = = [d]
2g 2g
The ratio of the height after and before the bounce can then be expressed as
hafter
= e2 [e]
hbefore
The coefficient of restitution e then becomes
r r
hafter 52
e = = = 0.8498 [f ]
hbefore 72
Let us return to the basketball launched horizontally from a height of hbefore = 4 ft. The
vertical velocity of the ball immediately before impact is
p √
vGz = − 2ghbefore = − 2 × 32.17 × 4 = −16.04 ft/sec [g]

so that the friction parameter derived in Equation [n] in the previous example (no initial
angular velocity) becomes
2 vG x 2 20
µs = − = = 0.2696 [h]
5 (1 + e) vGz 5 × 1.8498 16.04

and, from Equation [o] of the previous example, uGx = 20 × 3/5 = 12 ft/sec.
A coefficient of friction of 0.2696 or higher is required for the ball to not slip during
impact. Assuming the court has this much or more friction (which it does), the angular
velocity, in dimensionless form, is ω = 3/5, so the dimensional angular velocity after impact
becomes
vG 0.6 × 20
Ω0 = ω x = = 4.840 rad/s [i]
R 29.75/12

5.12 A More Accurate Model of Rigid Body Impact


The previous section considered impact of particles and of smooth spherical rigid bodies
and related the velocities of the colliding bodies along the line of impact by means of the
coefficient of restitution. This model begins to lose its validity when the impacting object
loses its symmetry and the line of impact no longer passes through the center of mass.
Under these circumstances, it is preferable to treat the compression and restitution stages
separately, and to keep in mind the possibility that the direction of the friction force may
change during impact.
Consider the impact of a rectangular object with a flat surface, as shown in Figure 5.40.
Kinetics Applications 215

vG
z
G
vG
x

C x
F^ ^
N

FIGURE 5.40
Free-body diagram of a falling box.

The line connecting the impact point and the center of mass makes an angle γ with the
vertical. Immediately before impact the center of mass has a velocity of vG = vGx i + vGz k
and the angular velocity is Ωj. vGx > 0 and vGz < 0. The line connecting the impact point
C and center of mass G is described as

rC/G = −R sin γi − R cos γk (5.84)

so that the velocities of C and G are related by

vC = vG + Ωj × rC/G = (vGx − RΩ cos γ) i + (vGz + RΩ sin γ) k (5.85)

This kinematic relationship is valid throughout impact. The forces generated during impact,
the impulsive normal and friction forces, are described as

F̂ = N̂ k + AF̂f i (5.86)

in which N̂ and F̂f are positive. The parameter A denotes the direction of the impulsive
friction force and its value is 1 or −1. The value of A is governed by the following rules:
• When vCx is not zero immediately before impact, A = −sign (vCx ). This rule states that
the impulsive friction force opposes the horizontal velocity of the impact point.
• When vCx = 0 immediately before impact, the direction of the friction force is dictated
by the horizontal acceleration of the impact point in the absence of a friction force.
The horizontal acceleration depends on the value of γ which dictates the sense of the
rotational moment generated by the normal force.
When γ > 0, which is the case shown in Figure 5.40, the moment that the normal force
generates is clockwise. Hence, C will have an acceleration to the left. The friction force
is to the right and A = 1. When γ < 0, the normal force generates a counterclockwise
moment, pushing point C to the right. Hence, the friction force is to the left and A = −1.
The two cases can be combined by defining A as A = sign (sin γ).
It follows that the moment generated by the impact forces can be expressed as
 
MG j = rC/G × F̂ = (−R sin γi − R cos γk) × N̂ k + AF̂f i
216 Applied Dynamics
 
= RN̂ sin γ − ARF̂f cos γ j (5.87)

Let us consider the case where γ > 0 and the initial translational and angular velocities
are such that the velocity of point C in the horizontal direction is greater than zero, or
vCx = vGx − RΩ cos γ > 0. Hence A = −1.

5.12.1 Compression Stage

vG
z
G
vG
x

C x
F^c ^
Nc

FIGURE 5.41
Free-body diagram during compression for vCx > 0 and γ > 0.

The free-body diagram for γ > 0 is shown in Figure 5.41. The horizontal velocity of
the contact point is to the right, so the friction force is to the left. During the compression
stage, the force balances have the form
+
X X
→ F =⇒ mvGx − F̂c = mwGx +↑ F =⇒ mvGz − N̂c = mwGz (5.88)

with w denoting the velocities at the end of the compression stage and subscript c denoting
compression. Using Equation (5.87), the moment balance is obtained by summing moments
about the y direction and going through the center of mass G as
X
 MG =⇒ IG Ω + RN̂c sin γ − ARF̂c cos γ = IG Ωc (5.89)

The three equations above hold regardless of whether sliding continues throughout com-
pression or not. They have to be supplemented with two additional equations that deal with
the strength of the impact and kinematic condition involving impact. Following are the two
possible scenarios:
1. Sliding Continues throughout Compression Stage: In this scenario, sliding con-
tinues throughout the compression stage. The corresponding fourth and fifth equations
are
F̂c = µk N̂c wCz = wGz + RΩc sin γ = 0 (5.90)
where the first expression indicates that sliding occurs throughout compression and the
second equation indicates that the vertical speed of contact point is zero at the end of
the compression stage.
Kinetics Applications 217

2. Sliding Ends during Compression Stage: In this scenario, sliding stops before the
compression stage ends. Such a scenario occurs when initially vCx > 0 and γ > 0, so
that the clockwise moment generated by the normal force further reduces the sliding
speed. In such a situation, the compression stage needs to be divided into two parts and
each part has to be analyzed separately.
Denote the velocities at the end of the sliding section by primes: w0 and Ω0c . In this
regime, the contact forces are N̂c0 and F̂c0 . The force and moment balances are
+
X X
→ F =⇒ mvGx − F̂c0 = mwG 0
x
+↑ F =⇒ mvGz − N̂c0 = mwG 0
z
(5.91)

X
 MG =⇒ IG Ω + N̂c0 R sin γ + F̂c0 R cos γ = IG Ω0c (5.92)

The fourth and fifth relationships become

F̂c0 = µk N̂c0 0
wC x
0
= wG x
− RΩ0c cos γ = 0 (5.93)

where the first expression indicates that the contact point slides and the second expres-
sion indicates that the horizontal velocity of the contact point C is zero at the end of
sliding.
Once sliding ends during compression, then during the second part of the compression
stage, the contact forces change. Let us denote the impulsive normal and friction forces
by N̂c00 and F̂c00 , respectively. The sign of the friction force depends on the angle γ. When
γ > 0, in the absence of friction, the impulsive normal force creates a clockwise moment,
which would cause point C to move to the left. Hence, the friction force is to the right,
as shown in Figure 5.42.
Two scenarios are possible for the second part of the compression stage: i) the impact
point C does slide to the left, and ii) the impact point does not slide. The determination
of whether sliding occurs or not is carried out the same way as in the first part of the
compression stage.

For negative values of γ (object leaning to the left), the normal force will create a coun-
terclockwise moment, giving point C an impending acceleration to the right. The friction
force will be to the left. But the friction force has already stopped a velocity to the right,
so that no further sliding of point C is possible.
Since γ varies between −90◦ and 90◦ , we can write the friction force in terms of the
signum function as F̂c00 sign(sin γ). Let us continue with the case of γ > 0. When no sliding
of the contact point occurs during the second part of compression, the force and moment
balances and the fourth and fifth relationships become
+
X X
0 00 0
→ F =⇒ mwG x
+ F̂ c = mw G x
+↑ F =⇒ mwG z
− N̂c00 = mwGz (5.94)

X
 MG =⇒ IG Ω0c + N̂c00 R sin γ − F̂c00 R cos γ = IG Ωc (5.95)

wCz = wGz + RΩc sin γ = 0 wCx = wGx − RΩc cos γ = 0 (5.96)

In the event of sliding during the second part of compression, the fourth and fifth equa-
tions, as represented by Equation (5.96) for no sliding, are replaced by

wCz = wGz + RΩc sin γ = 0 F̂c00 = µk N̂c00 (5.97)


218 Applied Dynamics
z

w'G
z
G
w'G 'c
x

R
w'C
x
C
^
F"
c ^
N"
c

FIGURE 5.42
Free-body diagram during second part of the compression stage.

At the end of the compression stage, the vertical velocity of the contact point is zero.
The total impulsive normal force during compression is the sum of the normal forces during
the sliding and no-sliding regimes, so that

N̂c = N̂c0 + N̂c00 (5.98)

The total impulsive friction force is the sum of the friction forces in the two parts,
F̂c = F̂c0 + F̂c00 . Note that the friction forces may have opposite directions.

5.12.2 Restitution Stage


The restitution stage depends on what happens during the compression stage. A number of
different scenarios are possible, which are summarized in Table 5.5. Note that it is possible
for sliding to come to an end during restitution and for the friction force to change direction.

TABLE 5.5
Possible sliding scenarios during impact

During Compression Stage During Restitution Stage


Slides throughout Slides throughout
Slides throughout Sliding stops
Slides throughout Sliding stops and reverses direction
Sliding stops No sliding
Sliding stops and reverses direction Sliding throughout in reverse direction
Sliding stops and reverses direction Sliding stops

Example 5.18
A pendulum of mass m and length L is released and makes impact with a horizontal surface.
Immediately before impact, the pendulum makes an angle of θ with the plane of impact, as
Kinetics Applications 219

shown in Figure 5.43, and it has angular velocity ω. Calculate the angular velocity of the
pendulum as it bounces back.

y
O L
C x
vC
F^cx F^cy

FIGURE 5.43
Free-body diagram during compression stage.

Separating the impact into the compression and restitution stages, during compression
the velocity of the impact point is towards the impact surface. The impulsive forces, shown in
Figure 5.43, are the impact force F̂cy and the friction force F̂cx , with subscript c denoting the
compression stage. The compression stage ends when the velocity along the impact direction
becomes zero. Because this problem involves a single degree of freedom and rotation about
a fixed point, the end of the compression stage is marked by the angular velocity of the
pendulum becoming zero. Summing impulsive moments gives
X
M̂O =⇒ −IO ω + M̂c = 0 [a]

where positive ω is counterclockwise. The impulsive moment during compression is

M̂c = F̂cy L cos θ + F̂cx L sin θ [b]

The impulsive friction force is assumed to be related to the impulsive normal force by
F̂cx = µs F̂cy . Introducing this to Equations [a] and [b] and solving for the impulsive force
gives
IO ω
F̂cy = [c]
L (cos θ + µs sin θ)

v'C y
O L
C
 x
F^rx
F^ry

FIGURE 5.44
Free-body diagram during restitution stage.

Next, consider the restitution stage, where the associated free-body diagram is given
in Figure 5.44, with F̂ry denoting the impact force, F̂rx the impulsive friction force, and
subscript r denoting restitution. Note that the direction of the friction force has changed
220 Applied Dynamics

from the compression stage. The impending velocity is to the right (positive x direction)
so that the impulsive friction force is to the left. Denoting the angular velocity at the end
of the restitution stage by Ω (positive counterclockwise) and summing impulsive moments
give X
M̂O =⇒ 0 + M̂r = IO Ω [d]
The impulsive moment during restitution is
M̂r = F̂ry L cos θ − F̂rx L sin θ [e]
As earlier, we assume that the impulsive friction force is related to the impulsive normal
force by F̂rx = µs F̂ry . Introducing this relationship to Equations [d] and [e] gives the angular
velocity at the end of impact as
F̂ry L
Ω = (cos θ − µs sin θ) [f ]
IO
The impulsive normal forces are related by the coefficient of restitution e
F̂ry = eF̂cy [g]
Introducing this to Equation [f] and considering Equation [c] gives the relationship between
the angular velocities before and after impact as
cos θ − µs sin θ
Ω = eω [h]
cos θ + µs sin θ
The above result is valid only when cos θ − µs sin θ > 0, which implies that the resultant
moment in the restitution stage should be a positive quantity. Otherwise, the rod will not
bounce back. The critical values of the angle θ and friction coefficient µ for this condition
is
1
= tan θ [i]
µs
For example, for µs = 0.8, the critical value of θ becomes θ = tan−1 (1/µs ) = 51.34◦ and
for µs = 0.5 we obtain θ = 63.43◦ . When there is no friction, the rod will always bounce
back. Another interesting special case is when θ = 0, in which case the pendulum is level
with the horizontal surface. In this special case, Ω = eω and whether there is friction or not
becomes immaterial, as there are no sliding velocities.
Note that solving this problem without differentiating between the compression and
restitution stages would give the result Ω = eω for all values of θ. This result is the same as
the no-friction case. Incorporating the compression and restitution stages into the formula-
tion takes into the consideration the friction forces.

5.13 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Benaroya, H., and Nagurka, M., Mechanical Vibration, 3rd Edition, CRC Press, 2009.
Bottega, W.J., Engineering Vibrations, CRC Press, 2006.
Karnopp, D. Vehicle Dynamics, Stability, and Control, 2nd Edition, CRC Press, 2013.
Moon, F., Applied Dynamics, with Applications to Multibody and Mechatronic Systems,
John Wiley, 1998.
Rao, S.S., Mechanical Vibration, 5th Edition, Prentice-Hall, 2010.
Wilson, D.G. (with Papadopoulos, J.), Bicycling Science, 3rd Edition, MIT Press, 2004.
Kinetics Applications 221

5.14 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 5.2—Rolling
5.1 (E) A sphere of mass m and radius R is released onto a flat surface with speed v and
no angular velocity and begins sliding upon contact. The coefficient of friction between the
sphere and the surface is µ. Calculate how long the sphere will continue sliding, until the
motion becomes roll without slip.
5.2 (M) Consider the previous problem and a bowling ball of weight 16 lb and diameter
8.6 in.; release speed of 10 ft/sec; and coefficient of friction of µk = 0.3, and calculate the
distance travelled. At what release speed will the ball continue sliding until it hits the pins?
A bowling alley is 60 ft long.

z
g  m  y

R x
G
C ax

s

FIGURE 5.45
Figure for Problem 5.3.

5.3 (M) The cylinder of mass m and radius R in Figure 5.45 is resting freely on the bed of a
truck. The truck is suddenly given an acceleration of 0.3g. Find the minimum value of the
coefficient of static friction µs so that the motion of the cylinder with respect to the truck
bed will be roll without slip.

!" #"

M g
g

M
 T
m D
R
v R $%
R G
G
FRR C

FIGURE 5.46
Figures for a) Problems 5.4, 5.5, and 5.7, b) Problems 5.9, 5.20, 5.31, and 5.32.

5.4 (M) The unicycle in Figure 5.46a is modeled as a point mass M over a disk of mass
222 Applied Dynamics

m and radius R. The top mass M moves together with the center of the disk. Given that
M = 70 kg, m = 10 kg and R = 45 cm, find the torque that is needed to propel the unicycle
with an acceleration of 2 m/s2 . Assume that friction is sufficient to prevent slip. The rolling
resistance force is FRR = 0 for this problem.
5.5 (M) The unicycle in Figure 5.46a is modeled as a point mass M over a disk of mass
m and radius R. The top mass M moves together with the center of the disk. Given that
M = 70 kg, m = 10 kg and R = 45 cm, find the torque that is needed to propel the unicycle
with an acceleration of 2 m/s2 . Assume that friction is sufficient to prevent slip. The rolling
resistance force is FRR = 0.02W .
5.6 (M) Show that for a disk of radius R that is propelled by pushing it at its center, similar
to Figure 5.2b; the amount of force Bx needed so that the wheels will slip is much larger
than T /R, where T is the torque needed to make the disk slip in Example 5.1.
5.7 (M) The unicycle in Figure 5.46a is modeled as a point mass M over a disk of mass
m and radius R. The top mass M moves together with the center of the disk. Given that
M = 70 kg, m = 10 kg and R = 45 cm, find the torque that will make the wheel slip for
when µs = 0.3. The rolling resistance force is FRR = 0 for this problem.

g z
o y

x
G
vo
R

FIGURE 5.47
Figure for Problem 5.8.

5.8 (M) The billiard ball of mass m and radius R in Figure 5.47 is struck with a cue stick
below the center of mass so that the ball acquires a horizontal velocity v0 and counterclock-
wise angular velocity Ω0 . Given friction coefficient µk , calculate the ratio of v0 to Ω0 so that
the ball eventually begins to roll backwards. Model the ball as a solid sphere.
5.9 (D) The imperfect disk in Figure 5.46b, whose center of mass is a distance e = R/5 that
is off center and whose mass moment of inertia is IG = 0.6mR2 , is at rest, when a torque
M is applied to it. If the coefficient of friction between the disk and the surface is 0.4, what
is the largest torque that can be applied so that the disk will still roll without slipping?

Section 5.4—Impulse and Momentum


5.10 (E) A vehicle of weight 2000 kg is traveling with speed of 90 kph when it hits a wall.
The collision with the wall lasts 0.5 seconds. Determine the average impact force on a front
seat passenger of weight 80 kg when a) the passenger is wearing a seat belt (deceleration of
passenger same as the vehicle) and b) the passenger is not wearing a seatbelt (passenger hits
the windshield in 0.001 seconds). Given that the human body does not sustain significant
injury at acceleration levels < 7g, assess the potential of injury.
5.11 (M) Consider the baseball (weight 5.125 oz, circumference 9 in) in Figure 5.48a. The
ball is coming to the batter at a horizontal speed of 80 ft/sec. With what impulsive force
and angle θ does the batter have to hit the ball, so that after the hit the ball will have
Kinetics Applications 223

!" #"
g g

G '()*+,-./ G
$ R
 R %&
F^ B B

F^

FIGURE 5.48
Figures for a) Problem 5.11, b) Problem 5.12.

a horizontal speed of 90 ft/sec and vertical speed of 72 ft/sec? Calculate the magnitude
of the force that the batter needs to apply, if the collision takes place in 0.001 sec. Then,
consider the case when the batter hits the ball with the same force and same angle, but for
a duration of 0.0007 seconds. Calculate the velocity of the baseball immediately after the
hit.
5.12 (C) Consider the baseball in Figure 5.48b. The ball is coming at the batter at a
horizontal speed of 85 ft/sec. The batter wants to hit the ball with an angle θ = 23◦ and
impulsive force of F̂ = 3 lb·sec. But rather than hitting the ball in the direction intended, the
batter hits the ball with an angle φ, while imparting the desired impulsive force. Calculate
and plot the resulting horizontal and vertical velocities after impact, as well as the resulting
angular velocity, as a function of the angle φ, in the range −30◦ ≤ φ ≤ 30◦ .
5.13 (C) Consider the baseball in the previous problem and calculate and plot the hang
time and distance travelled by the baseball. Consider the same range for φ as the previous
problem. Model aerodynamic drag but ignore the effect of the rotation of the ball (Magnus
effect). Use a drag coefficient of CD = 0.3.

y
a) B b) y
m, L
x
B

2m, L m, L m, L
o
120
C
A C A x
^
^ F
F (Impulsive force)

FIGURE 5.49
Figures for a) Problem 5.14, b) Problem 5.15.

5.14 (M) The T-shaped body in Figure 5.49a is resting on a frictionless table when it is hit
by an impulsive force F̂ . Find the velocity of point A immediately after the impulse.
5.15 (D) The two rods AB and BC in Figure 5.49b are connected by a pin joint. They are
224 Applied Dynamics

at rest on a frictionless table when an impulsive force F̂ is applied to point C. Calculate


the angular velocities of the two rods immediately after the impulse.

a) b)
2

0.75 m 2 kg u
O v v
R
0.5 m 30 gm C h
150 m/s

FIGURE 5.50
Figures for a) Problem 5.16, b) Problem 5.17.

5.16 (E) A rod of mass 2 kg and length 0.75 m pivots about point O, as shown in Figure
5.50a. A 30 gm bullet, traveling with speed 150 m/s impacts the rod at point A and gets
embedded in it. Assuming the rod does not suffer any damage, calculate the angular velocity
of the rod immediately after impact.

Section 5.5—Work, Energy and Power


5.17 (M) The disk in Figure 5.50b of radius R is rolling without slipping with angular
velocity Ω when it encounters a step of height h (h < R). Assuming that during contact
with point C the disk pivots about C, calculate the speed needed by the disk so it can climb
over the stair and begin rolling (Ω2 > 0, u > 0). Calculate also the energy loss.
5.18 (E) Consider a cyclist with a frontal area of 0.9 m2 and drag coefficient7 CD = 1.1. a)
How much power does the cyclist have to generate to maintain a speed of 9 kph? b) There
is a headwind of 2.5 m/s. Recalculate the power requirement. c) The rolling resistance force
(see Figure 5.46a) is 0.01 times the combined weight of the bicycle and rider. If the rider
has a mass of 80 kg and the bicycle is 16 kg, calculate the rolling resistance force and the
power that the rider needs to generate in order to travel at the conditions of case a and
case b.
5.19 (M) Consider a cyclist with KD = 0.4 kg/m (see Problem 4.29) and a combined cyclist-
bicycle mass of 92 kg. The cyclist can generate power of 110 Watts. Calculate how fast the
cyclist can travel uphill on a 4% grade, with a rolling resistance force FRR = 0.02W (see
Figure 5.46a) and headwind of 3 km/h. Hint: Solve the resulting equation numerically or
plot the power vs. speed curve.
5.20 (M) Find the smallest initial velocity that can be imparted on the imperfect disk in
Figure 5.46b so that the disk can complete at least one revolution. The initial position is
such that the center of mass is as close to the ground as possible (equilibrium position).
5.21 (M) The particle of mass m in Figure 5.51 is acted upon by a force expressed in polar
coordinates as F = − k/r2 er , where r is the distance from the origin to the particle.
Calculate the total energy and the angular momentum about O and show that these two
quantities are constant.
5.22 (M) A particle of mass m is acted upon by a force expressed in spherical coordinates
as F = −kReR , where R is the distance from the origin to the particle. Calculate the total
7 This value for C
D is larger than the drag coefficient of a truck. With the help of slick cycling gear and
leaning of the cyclist, the drag coefficient for bicycle racers reduces to below 0.9.
Kinetics Applications 225

Y
r

Fr
m


X
O

FIGURE 5.51
Figure for Problem 5.21.

energy and the angular momentum about the origin and show that these two quantities are
constant.
5.23 (E) The vehicle in Figure 5.52 has a total mass of 1800 kg. The vehicle is released from
rest at a 6◦ incline. The wheels roll without slip and the only force other than gravity and
friction that acts on the vehicle is rolling resistance, which is a force opposite the direction
of motion with a magnitude of FRR = 0.01W . Calculate the speed of the vehicle after the
vehicle has rolled 100 m down the incline.

a) b)
g
x, v
v
kV kB
FRR W
o
6

FIGURE 5.52
Figures for a) Problem 5.23, b) Problem 5.24.

5.24 (M) A vehicle of mass 1900 kg is traveling with a speed of 45 kph when it hits a barrier.
The barrier can be modeled as an axial spring of stiffness kB N/mm and the vehicle’s crush
resistance (see Example 5.15) can be modeled as a spring kV = 35 N/mm, as shown in
Figure 5.52b. After the crash, it is observed that the total deflection of the vehicle and the
barrier is 0.5 m. Calculate the crush stiffness kB of the barrier.
5.25 (D) The solid uniform sphere in Figure 5.53 of mass m and radius R is placed on top
of a fixed sphere of the same radius, and it is slightly tipped. Assuming that friction is
sufficient to prevent slipping, find the value of the angle θ at which point the spheres lose
contact.
5.26 (D) Consider the solid uniform sphere in Figure 5.53 of mass m and radius R which
is placed on top of a fixed sphere of the same radius. Assuming that friction coefficient is
µs = 0.6, find the value of the angle θ at which point the spheres slide against each other.
226 Applied Dynamics


g
R

FIGURE 5.53
Figure for Problems 5.25 and 5.26.

Section 5.6—Equations of Motion


5.27 (M) Consider the crane in Example 5.6 and now treat the motion of the trolley x as a
variable. The system now has two degrees of freedom. Obtain the two equations of motion,
considering that a horizontal force F controls the motion of the trolley, which is of mass
M . Discuss the circumstances under which modeling the crane as a one-degree-of-freedom
system is a reasonable assumption.
5.28 (M) Obtain the equations of motion of the masses with a pulley in Figure 1.9.
5.29 (M) Consider the sphere rolling without slipping inside a semicircle in Example 3.2
and obtain the equation of motion.
5.30 (M) Find the equation of motion of the bracket in Problem 4.40.

a) g b) g

B
m L/2
R
 k
x L/2
D R/3 O
y m x
G M
A
 k
C
x

FIGURE 5.54
Figure for Problems a) 5.31, b) 5.33.

5.31 (M) The imperfect disk in Figure 5.46b is given a small initial velocity that is not
sufficient for the disk to complete a revolution. Hence, the disk rocks back and forth. Obtain
the equation of motion of this rocking motion, using the angle θ as the motion variable, as
shown in Figure 5.54a. Friction is sufficient to prevent slip.
Kinetics Applications 227

5.32 (M) The imperfect disk with the center of mass away from the geometric center by
R/5 in Figure 5.46b is given an initial velocity that is sufficient for the disk to complete a
revolution. Friction is sufficient to prevent slip. Find the equation of motion of the disk for
the rolling motion using the position of the center of the disk as the motion variable.
5.33 (M) The rod in Figure 5.54b pivots about point O. It is supported by a spring on one
end and at the other end a mass is connected to it through a spring. Obtain the equations
of motion. The springs deform only vertically and are undeformed when θ = 0 and x = 0.
5.34 (M) Consider the bicycle model of a vehicle in Figure 3.9 and obtain the equation of
motion associated with the speed vA by summing moments about the instant center. There
is a force FT acting on the rear wheel in the x direction. Express the resulting equation in
terms of vA only by eliminating θ̇ by means of Equation (3.19).
5.35 (M) Consider the supermarket cart in Figure 8.7 and obtain the equations of motion
using a Newtonian approach. Use as motion variables the coordinates of point A, XA and
YA , and the yaw angle θ. Also express the velocities in terms of the speed of point A, vA ,
and angular velocity ω = θ̇. Note that, because of the no sliding constraint at point A,
the cart has only two degrees of freedom. Also, the caster at point B eliminates all friction
forces there.
5.36 (M) Obtain the equations of motion of the double pendulum in Figure 5.55a, assuming
that the cables are always taut.
5.37 (M) Obtain the equation of motion of the rod in Example 5.9 by differentiating the
energy.

Section 5.8—Linearization, Equilibrium, and Stability

a) b)

g
g
L1 O k
x

1
m1 z m, L m, L
L2
2
m2

FIGURE 5.55
Figures for a) Problem 5.36, b) Problem 5.38.

5.38 (M) Find the equilibrium position for the two links attached to a spring in Figure
5.55b. The spring is not stretched when the links are horizontal. Hint: Use potential energy
formulation.
5.39 (M) Find the equilibrium positions for the system in Problem 5.28.
5.40 (M) Find the equilibrium positions for the system in Figure 5.56a.
5.41 (M) Find the equilibrium position of the bracket in Problem 5.30. You can solve this
problem by calculating the potential energy and differentiating it.
5.42 (M) Find the equilibrium positions for the system in Figure 5.56b by using a) the
displacements of the masses as motion variables and b) the deflections of the springs as
motion variables. Comment on which set is preferable.
228 Applied Dynamics

!" #"

k
g g
m
k
m
-k

$!#%& -m
-k
'()*!+*
,m

FIGURE 5.56
Figures for a) Problem 5.40, b) Problem 5.42.

5.43 (C) Consider the system in Figure 5.57a and find the equilibrium positions of the
masses.
5.44 (C) Consider the rod in Figure 5.57b, which is suspended from two springs, using the
displacement of the center of mass and the rotation of the rod as motion variables. Obtain
the equations of motion and find the equilibrium positions, and linearize about the position
zG = 0, θ = 0. The springs are not stretched when zG = 0, θ = 0.

#$ %$
g
g
k
m !k
zG B
m x&
k

G L'(
!k
"L'(
A
x! "L'(
!m
L'(
"k

"m x"

FIGURE 5.57
Figures for a) Problem 5.43, b) Problem 5.44.

5.45 (M) A particle of unit mass is acted upon by an excitation F (x) = −x + x2 /4 − 0.1ẋ.
Find the equilibrium position(s) and ascertain their stability.
5.46 (M) A particle of unit mass is acted upon by an excitation F (x) = −x + x3 /9. Find
the equilibrium position(s) and ascertain their stability.
5.47 (M) Consider the rod and springs in Figure 4.57b. Here, the rod is of length L and
Kinetics Applications 229

mass m. Obtain the equation of motion and equilibrium position. Analyze the stability of
the equilibrium positions.
5.48 (M) Consider the double pendulum in Problem 5.36. Linearize the equations of motion
about the operating points θ1 = θ2 = 0.

Section 5.9—Motion in the Vicinity of the Earth


5.49 (C) A spherical missile of mass 2 kg and radius 12 cm is launched at Yankee Stadium
in the easterly direction with an initial speed of 600 km/h and with an angle of 30◦ with the
horizontal. The drag coefficient between the ball and the air is CD = 0.32. Write a program
to integrate the equations of motion to determine when and where the missile lands. Include
the Coriolis effect in your model.

Section 5.10—Collisions
5.50 (M) Two vehicles collide. Vehicle A has a weight of WA = 3000 lb and its crush
stiffness is kA = 7900 lb/in. Vehicle B has a weight of WB = 4500 lb and has a stiffness of
kB = 5100 lb/in. Vehicle A is stationary when it is rear-ended by vehicle B. The vehicles
are mangled and they skid together to a stop. Measurements after the accident show that
after the crash the two mangled vehicles skidded 22 ft before coming to a stop. The sliding
friction coefficient between the two crushed vehicles and the road is estimated as µk = 0.7.
Calculate the impact forces and accelerations experienced by the vehicles, as well as the
speed of vehicle B immediately before the crash.

Section 5.11—Impact of Rigid Bodies: Simple Solution


5.51 (M) Consider the basketball in Example 5.17 and an average women’s NCAA basketball
(dimensions in Table 5.4). Calculate the coefficient of restitution from the bounce. Given a
coefficient of friction of µs = 0.7 and the basketball passed forward from a height of 1.5 m,
calculate the maximum initial horizontal velocity for the ball to not slide upon impact.
5.52 (M) Consider the basketball in Example 5.16. Obtain the solution for the nondimen-
sional equations for the bouncing ball in the presence of an initial angular velocity. Calculate
the amount of backspin that would be necessary for the ball to have zero forward velocity
immediately after impact.
5.53 (M) Discuss the advantages of having a topspin on a ball by obtaining expressions for
the friction force and energy loss. Conduct this analysis for the case of no sliding. What
happens when the topspin is high enough so that the contact point C has zero horizontal
velocity immediately before impact?

Section 5.12—A More Accurate Model of Rigid Body Impact


5.54 (D) A solid sphere of mass m and radius R makes contact with a surface while its
center of mass velocity is vG = vx i − vz k and angular velocity is zero. Find the value of
the smallest coefficient of friction that will cause the sliding to end during the compression
stage.
6
Response of Dynamical Systems

6.1 Introduction
The previous chapters focused on the kinematics and kinetics of dynamical systems, and
they discussed the describing equations, linearization, and qualitative analysis of the re-
sponse. This chapter and the next consider the response of dynamical systems from a
quantitative perspective. The interest is in
• Solving the equations of motion for the response, analyzing how the system behaves
over time, and learning what the motion amplitudes are as a function of the excitation;
• Looking at the response of a system whose model is not accurately known and ascer-
taining the system properties from the response.
This chapter deals with the response of linear (or linearized) single-degree-of-freedom sys-
tems. The next chapter examines the response of multi-degrees-of-freedom dynamical sys-
tems.
The describing equations of a dynamical system are, for the most part, in terms of
differential equations; thus, this chapter looks at techniques for solving ordinary differen-
tial equations and develops physical explanations of the response. Both time-domain and
frequency-domain techniques are considered. While studying time-domain techniques, two
solution methods are discussed: the homogeneous plus particular solution approach and the
Laplace transform solution.

6.2 The Unit Impulse and Unit Step Functions


This section considers two mathematical functions that are useful in the study of the re-
sponse of systems. The unit impulse and unit step functions constitute a special class of
aperiodic functions. They are not traditional functions (which are defined by what they are);
rather, they are defined by what they do to other functions to which they are multiplied.

6.2.1 The Unit Impulse Function


The unit impulse function, also known as the Dirac delta function, is named after the
British physicist Paul Dirac (1902–1984), who shared the Nobel prize for physics in 1933
with Edwin Schrödinger and who held the Lucasian Chair of Mathematics at the University
of Cambridge. The Dirac delta function has the shape of a spike, whose duration approaches
zero and whose amplitude approaches infinity, as shown in Figure 5.11 and as described in

231
232 Applied Dynamics

the previous chapter. Denoted by δ(t − to ), the Dirac delta function is defined as
Z ∞
δ (t − to ) = 0 when t 6= to δ (t − to ) dt = 1 (6.1)
−∞

The unit of the Dirac delta function is 1/time. Because of its high amplitude and short
duration, the Dirac delta function is useful in modeling impulsive forces, such as forces that
occur during collisions, as was discussed in Chapter 5.
It is of interest to evaluate the integral of the product of the Dirac delta function
multiplied by another function, say, f (t)
Z ∞ Z ∞ Z ∞
f (t) δ (t − t0 ) dt = f (t0 ) δ (t − t0 ) dt = f (t0 ) δ (t − t0 ) dt = f (t0 ) (6.2)
0 0 0

The second term in the above equation arises because the Dirac delta function is zero when
t 6= t0 . Multiplying a function by the Dirac delta function at t = t0 and integrating gives
the value of the function at t = t0 , that is, f (t0 ). The process is similar to taking a snapshot
of f (t) at t = t0 , or a sampling of f (t) at t = t0 .
The Dirac delta function is not a discontinuous function. Rather, it can be shown to
arise from a limiting process. This means that its derivative exists. Denoting this derivative
by δ 0 (t − t0 ), multiplying it by an arbitrary function f (t), and integrating by means of
integration by parts gives
Z ∞ Z ∞

f (t) δ 0 (t − t0 ) dt = f (t) δ (t − t0 )|−∞ − f 0 (t) δ (t − t0 ) dt = −f 0 (t0 ) (6.3)
−∞ −∞
0
in which f (t) = df (t) /dt.

6.2.2 The Unit Step Function


The unit step function, also known as the Heaviside function, named after the self-taught
British scientist Oliver Heaviside (1850–1925), is denoted by
(
0 for t < t0
u (t − t0 ) = (6.4)
1 for t > t0

!!"#$#""#

$"" $"

FIGURE 6.1
Unit step function u (t − t0 ).

Depicted in Figure 6.1, the unit step function is discontinuous at t = t0 . It is dimension-


less. We can show that the unit step function is the integral of the unit impulse function
Z t
u (t − t0 ) = δ (τ − t0 ) dτ (6.5)
−∞
Response of Dynamical Systems 233

with τ as a dummy variable of integration. It follows that the unit impulse function is the
derivative of the unit step function

du (t − t0 )
δ (t − t0 ) = (6.6)
dt
The unit step function is helpful when expressing the transient response of a system and
when forces are applied to a system at different times.

6.3 Homogeneous Plus Particular Solution Approach


After obtaining the linear or linearized equation of motion, the next step is to solve the
resulting ordinary differential equation and to obtain the solution. This section discusses
the homogeneous plus particular solution approach. The Laplace transform solution will
be discussed in the next section. It should be noted that this and the next section discuss
solution procedures. They do not discuss the theory behind and the derivations of the
solution methods. The interested reader is referred to texts on applied mathematics.
Consider a linear, constant coefficient ordinary differential equation (ODE) Dx (t) =
F (t), where x (t) is the variable, D is the differential operator, and F (t) is the excitation.
Let the excitation be an explicit function of time and not a function of the variable x (t) or
its derivatives. The following steps outline the solution of an ordinary differential equation
by the homogeneous plus particular solution approach:
• The solution is written as the sum of a term called the homogeneous solution plus a
term called the particular solution; x (t) = xH (t) + xP (t).
• The homogeneous solution xH (t) is the solution to the differential equation in the
absence of excitation, that is, the solution to Dx (t) = 0. This solution, which will be
illustrated by an example shortly, is in terms of unknown coefficients.
• The particular solution xP (t) has the same form as the excitation. For example, if
the excitation is in the form F (t) = sin ωt, where ω is a parameter, then xP (t) =
E1 sin ωt + E2 cos ωt, or, if F (t) = tn , then xP (t) = E0 + E1 t + E2 t2 + . . . + En tn .
The coefficients Ei are obtained by plugging the particular solution into the differential
equation and collecting like terms. When the excitation has several terms in it, we can
calculate the particular solution associated with each term of the excitation separately
and combine these solutions by means of the principle of superposition.
• The homogeneous and particular solutions are combined, and the initial conditions are
invoked to find the undetermined coefficients. Note that the initial conditions satisfy the
entire solution x (t) = xH (t)+xP (t). It is a common mistake to evaluate the coefficients
in xH (t) by incorrectly thinking that the initial conditions satisfy xH (t).

Example 6.1
Obtain the solution to the differential equation ẍ (t) + 4x (t) = 2.5t − cos t, subject to the
initial conditions x (0) = 0.2, ẋ (0) = −0.1.
The differential operator is D = d2 /dt2 + 4 and the excitation is F (t) = 2.5t − cos t.
The homogeneous solution, xH (t), satisfies the differential equation

ẍH (t) + 4xH (t) = 0 [a]


234 Applied Dynamics

Considering a solution in the form xH (t) = Ceλt , introducing it into Equation [a], and
collecting terms gives
Ceλt λ2 + 4 = 0

[b]
We are seeking a nontrivial solution, that is, x(t) 6= 0. We note that the Ceλt cannot be
zero, so the terms in the brackets in Equation [b] must vanish. This relationship
λ2 + 4 = 0 [c]
is recognized as the characteristic equation.
The solution of the characteristic equation is two pure imaginary roots λ1,2 = ±2i. As
will be demonstrated in the next sections, the homogeneous solution can be written as
xH (t) = C1 sin 2t + C2 cos 2t [d]
with C1 and C2 yet to be determined.
The particular solution can be found by making use of the principle of superposition and
splitting the excitation into two, F (t) = F1 (t) + F2 (t), where F1 (t) = 2.5t and F2 (t) =
− cos t. It follows that xP (t) = xP1 (t) + xP2 (t), where xP1 (t) is the particular solution for
the excitation F1 (t) = 2.5t and xP2 (t) is the particular solution for F2 (t) = − cos t. The
individual particular solutions are
xP1 (t) = D0 + D1 t xP2 (t) = E1 sin t + E2 cos t [e]
Taking the second derivatives of these solutions and substituting them into the equation
of motion gives
4 (D0 + D1 t) = 2.5t [f ]
and
−E1 sin t − E2 cos t + 4 (E1 sin t + E2 cos t) = − cos t [g]
The coefficients D0 , D1 , E0 , E1 can be obtained by separating the above equations
to coefficients of powers of t in Equation [f] and as coefficients of the sinusoidal terms in
Equation [g]:
1st equation =⇒ t0 : 4D0 = 0 t1 : 4D1 = 2.5 [h]
2nd equation =⇒ sin t : 3E1 = 0 cos t : 3E2 = −1 [i]
which can be solved for the coefficients, with the result
5 1
D0 = 0 D1 = E1 = 0 E2 = − [j]
8 3
Substituting the results in Equation [j] into Equation [e], we obtain xP1 (t) = 5t/8, xP2 =
− cos t/3.
The total solution is the sum of the homogeneous and particular solution as
5 1
x (t) = xH (t) + xP (t) = C1 sin 2t + C2 cos 2t + t − cos t [k]
8 3
Next, the initial conditions are used to find C1 and C2 . At t = 0 the position x (0) has
the form
1
x (0) = C2 − = 0.2 =⇒ C2 = 0.5333 [l]
3
Differentiating Equation [k] with respect to time and evaluating at t = 0 gives
5
ẋ (0) = 2C1 + = −0.1 =⇒ C1 = −0.3625 [m]
8
so that the total response becomes
5 1
x (t) = −0.3625 sin 2t + 0.5333 cos 2t + t − cos t [n]
8 3
Response of Dynamical Systems 235

6.4 Laplace Transform Solution


The Laplace transform method is named after the French mathematician and astronomer
Pierre-Simon Laplace (1749–1827), who made contributions in statistics, spherical harmon-
ics, potential theory, and analyzing the shape of the Earth. The method transforms a dif-
ferential equation into an algebraic one, while automatically taking into consideration the
initial conditions. Manipulation of the resulting algebraic expressions leads to the solu-
tion. This section deals with the application of the Laplace transform to solve differential
equations and it does not represent an in-depth study.

6.4.1 General Formulation


Given a function of time f (t), its Laplace transform is denoted by F (s) = L (f (t)) and is
defined by
Z ∞
F (s) = L (f (t)) = e−st f (t) dt (6.7)
0

where s is known as a subsidiary variable, which, in general, is a complex number. The


Laplace transform has some interesting properties. Given a function f (t), the Laplace
transform of its derivative, L (df (t) /dt), can be obtained using integration by parts as
follows:
Z ∞
df (t) d
L = e−st f (t) dt
dt 0 dt


Z ∞
= e−st f (t) 0 − −se−st f (t) dt = −f (0) + sF (s)

(6.8)
0

The above result assumes that f (t) does not increase more rapidly than e−st decreases.
It follows that the Laplace transform of the second derivative of a function becomes
Z ∞
d2 f (t) d2
L 2
= e−st 2 f (t) dt = −sf (0) − f˙ (0) + s2 F (s) (6.9)
dt 0 dt

where f˙ (0) is the derivative of f (t) evaluated at t = 0.


The Laplace transform is a linear operation, thus permitting application of the principle
of superposition. That is, given a function f1 (t) whose Laplace transform is F1 (s) and a
function f2 (t), whose Laplace transform is F2 (s), then for any α and β the relationship

L (αf1 (t) + βf2 (t)) = αF1 (s) + βF2 (s) (6.10)

holds. The Laplace transform of a complicated expression can be obtained by breaking down
the expression into sums of simpler equations whose individual transforms can be obtained
separately and then by combining the individual transforms. Laplace transforms of common
functions are tabulated in Table 6.1.
Of interest in response analysis is an expression of the type
Z t Z t
x (t) = f1 (τ ) f2 (t − τ ) dτ = f1 (t − τ ) f2 (τ ) dτ (6.11)
0 0

which is known as the convolution integral. Denoting the Laplace transforms of x (t), f1 (t),
236 Applied Dynamics

and f2 (t) by X (s), F1 (s), and F2 (s), respectively, and calculating the Laplace transform
of the above equation gives
Z ∞ Z t 
X (s) = L (x (t)) = e−st f1 (τ ) f2 (t − τ ) dτ dt (6.12)
0 0

Evaluation of the above integral can be carried out by means of a series of substitutions
in the limits of the integration and in the order of integration. Details can be found in texts
on applied mathematics. The end result can be shown to be

X (s) = F1 (s) F2 (s) (6.13)

If the function G (s) is the Laplace transform of the function g (t), then g (t) is said to
be the inverse Laplace transform of G (s). While a formal definition of the inverse Laplace
transform exists, this definition is rarely used to calculate the inverse Laplace transform. We
usually rely on tables of Laplace transform pairs to evaluate the inverse Laplace transform.
When the expression whose inverse Laplace transform is sought does not look like one of
the commonly known pairs, the expression can be manipulated algebraically to write it
as a series of expressions whose inverse transform can readily be found. There are several
approaches for doing this. The partial fraction method is outlined in the example that
follows.

TABLE 6.1
Laplace transform pairs

Function f (t) Transform L(f (t)) = F (s)


δ (t) Dirac delta function 1
1
1 s
1
t s2
tf (t) −dF(s)/ds
n!
tn sn+1
e−at 1
s+a
te−at 1
(s+a)2
ω
sin ωt s2 +ω 2
s
cos ωt s2 +ω 2
e−at sin ωt ω
(s+a)2 +ω 2
e−at cos ωt s+a
(s+a)2 +ω 2
ω
sinh ωt s2 −ω 2
s
cosh ωt s2 −ω 2
1/2
√1 e−ζωt sin 1 − ζ 2 ωt 1
s2 +2ζωs+ω 2
ω 1−ζ 2
 
1/2 1/2
e−ζωt cos 1 − ζ 2 ωt + √ ζ sin 1 − ζ 2 ωt s+2ζω
s2 +2ζωs+ω 2
1−ζ 2
Response of Dynamical Systems 237

Example 6.2
s
Obtain the inverse Laplace transform of X (s) = (s+a)(s 2 +b2 ) by means of a partial fraction

expansion.
Let us write X (s) as the sum of two fractions

s A Bs + C
X (s) = = + 2 [a]
(s + a) (s2 + b2 ) s+a s + b2

To find the coefficients A, B, and C, the right side above equation can be rewritten as

A Bs + C A s2 + b2 + Bs (s + a) + C (s + a)
+ 2 = [b]
s+a s + b2 (s + a) (s2 + b2 )

Comparing the numerators of the left side of Equation [a] and the right side of Equation
[b] results in

s2 terms: 0 = A + B s1 terms: 1 = Ba + C s0 terms: 0 = Ab2 + Ca [c]

The three relationships in Equation [c] need to be solved for the three unknowns A, B, and
C. The first equation gives B = −A and the third gives C = −Ab2 /a. Introduction of these
two results into the second equation yields

Ab2
1 = −Aa − [d]
a
so that the coefficients are
a a Ab2 b2
A = − B = −A = C = − = 2 [e]
a + b2
2 a + b2
2 a a + b2
Introducing these values into Equation [a] gives

b2
 
s 1 a as
X (s) = = − + + [f ]
(s + a) (s2 + b2 ) a 2 + b2 s + a s2 + b2 s2 + b2

From the Laplace transform pairs in Table 6.1, the inverse Laplace transform is recog-
nized as
1
x (t) = L−1 (X (s)) = 2 −ae−at + a cos bt + b sin bt

2
[g]
a +b

6.4.2 Solving Differential Equations Using the Laplace Transform


The following procedure is used to obtain the solution of a differential equation by means
of the Laplace transform. Given a differential equation of the form

d2 d
a2 2
x (t) + a1 x (t) + a0 x (t) = f (t) (6.14)
dt dt
with initial conditions x (0) and ẋ (0), the first step is to obtain the Laplace transform of
the differential equation, which gives

a2 s2 X (s) − a2 sx (0) − a2 ẋ (0) + a1 sX (s) − a1 x (0) + a0 X (s) = F (s) (6.15)

where X (s) and F (s) are the Laplace transforms of x (t) and f (t), respectively.
238 Applied Dynamics

The next step is to collect like terms, which gives

a2 s2 + a1 s + a0 X (s) = x (0) (a2 s + a1 ) + ẋ (0) a2 + F (s)



(6.16)

and, dividing by a2 s2 + a1 s + a0 , we arrive at an expression for X (s)


a2 s + a1 a1 F (s)
X (s) = x (0) + ẋ (0) + (6.17)
a2 s2+ a1 s + a0 2
a2 s + a1 s + a0 2
a2 s + a1 s + a0
The solution is obtained by calculating the inverse Laplace transforms of each term
individually. If necessary, partial fraction expansions or another technique can be used to
express the above equation as a collection of terms whose inverse Laplace transform is
easily recognized. This solution procedure will be demonstrated in Sections 6.5 and 6.8,
while calculating the response of first- and second-order systems.

Example 6.3
Obtain the solution of the differential equation in Example 6.1 by means of the Laplace
transform.
The differential equation and boundary conditions are

ẍ (t) + 4x (t) = 2.5t − cos t, x (0) = 0.2, ẋ (0) = −0.1 [a]

Taking the Laplace transform we obtain


2.5 s
s2 + 4 X (s) = 2 − 2

+ 0.2s − 0.1 [b]
s s +1
Dividing both sides by s2 + 4 gives the expression for X as
2.5 s 0.2s 0.1
X (s) = − + − [c]
s2 (s2 + 4) (s2 + 1) (s2 + 4) (s2 + 4) (s2 + 4)
We use partial fraction expansions to put the terms above into forms whose inverse
Laplace transforms we can calculate. We write the Laplace transform of x as X = X1 +
X2 + X3 + X4 . For the first term, X1 , we obtain
2.5 As + B Cs + D
X1 (s) = = + 2 [d]
s2 (s2 + 4) s2 s +4
Equating the denominators we get

(As + B) s2 + 4 + s2 (Cs + d) = 1

[e]

Separation into the powers of s gives

s3 =⇒ A + C = 0 s2 =⇒ B + D = 0 s1 =⇒ 4A = 0 s0 =⇒ 4B = 1 [f ]

whose solution is A = C = 0, B = 1/4, D = −1/4. It follows that X1 (s) and its inverse
Laplace transform are
   
2.5 1 1 −1 5 1
X1 (s) = − x1 (t) = L X1 (s) = t − sin 2t [g]
4 s2 s2 + 4 8 2
The second term, X2 (s) can be expressed as
s As + B Cs + D
−X2 (s) = = + 2 [h]
(s2 2
+ 1) (s + 4) 2
s +4 s +1
Response of Dynamical Systems 239

Equating denominators yields

(As + B) s2 + 1 + (Cs + d) s2 + 4 = s
 
[i]

and separation into coefficients of the powers of s results in

s3 =⇒ A+C = 0 s2 =⇒ B+D = 0 s1 =⇒ A+4C = 1 s0 =⇒ B+4D = 0 [j]

whose solution is B = D = 0, A = −1/3, C = 1/3.


We conclude that X2 (s) and its inverse Laplace transform are
 
1 s s 1 1
X2 (s) = 2
− 2
x2 (t) = L−1 X2 (s) = cos 2t − cos t [k]
3 s +4 s +1 3 3

The inverse Laplace transforms of the next two terms are readily recognized as
   
−1 0.2s −1 −0.1
x3 (t) = L = 0.2 cos 2t x4 (t) = L = −0.05 sin 2t [l]
s2 + 4 s2 + 4

The solution can be expressed as x(t) = x1 (t) + x2 (t) + x3 (t) + x4 (t) and is the same as
the solution obtained in Example 6.1.

6.5 Response of First-Order Systems


Even though in dynamics we mostly encounter systems described by second-order differen-
tial equations, there exist systems that are described by first-order differential equations.
Example 4.6 casts the motion description in normal-tangential coordinates in terms of two
first-order equations. So are the stability equations for the lateral motion of a vehicle in
Chapter 14. Also, in numerical analysis, we deal with second- or higher-order differential
equations by converting them into a series of first-order differential equations.
Consider a system described by
1
ẋ (t) + x (t) = f (t) (6.18)
τ0
where x is the variable, t is time, and τ0 is a parameter known as the time constant, whose
unit is time. The initial condition is x (0) = x0 . Consider first the free response, by setting
the excitation f (t) = 0 and assuming a solution in the form x (t) = Ceλt , where C and λ
are unknowns. Substituting into the differential equation gives
 
1 1
λCeλt + Ceλt = 0 =⇒ λ+ Ceλt = 0 (6.19)
τ0 τ0

For a nontrivial solution, Ceλt 6= 0, so that the only way the above equation is satisfied
is by having
1
λ+ = 0 (6.20)
τ0
Equation (6.20) is the characteristic equation and the value(s) of λ that satisfy it are
240 Applied Dynamics

the characteristic values. For the first-order system here, the solution of the characteristic
equation is
1
λ = − (6.21)
τ0
so that the free response becomes
t
x (t) = Ce− τ0 (6.22)

The value of C is determined from the initial condition. Evaluating the above equation
at t = 0 gives
t
x (0) = x0 = Ce0 = C =⇒ x (t) = x0 e− τ0 u (t) (6.23)

1
x(t/ 0)/x0

0.368

0
0 1
Time t/ 0

FIGURE 6.2
Free response of a first-order system: x(t/τ0 )/x0 vs. t/τ0 .

The response is in the form of exponential decay, as shown in the nondimensionalized


plot in Figure 6.2. The rate of decay is determined by the time constant, which explains
the name. For example, when t = τ0 , the value of x (τ0 ) becomes x (τ0 ) = x0 /e = 0.368x0
and x (2τ0 ) = x0 /e2 = 0.135x0 . Note the unit step function u (t) in Equation (6.23). This
is done to indicate that the response is zero for t < 0. Though not essential for the simple
response term above, it is nonetheless a good idea to use the unit step function to denote
the time at which the response begins.
Next, consider a constant excitation that acts on the first-order system. We express the
excitation as f (t) = f0 /τ0 and consider a zero initial condition, x (0) = 0. The unit of f0 is
displacement. We will solve this equation two ways. Using the homogeneous plus particular
solution approach, the response has the form x (t) = xH (t) + xP (t). From the derivation of
t
the free response earlier, xH (t) = Ce− τ0 , with C unknown. The particular solution depends
on the nature of the excitation. Since the excitation is constant, so is the particular solution,
so that xP (t) = E. The value of E is ascertained by plugging the particular solution into
the differential equation, with the result
1 f0 1 f0
ẋP (t) + xP (t) = =⇒ E = =⇒ E = f0 (6.24)
τ0 τ0 τ0 τ0
Response of Dynamical Systems 241

so that the total solution becomes


t
x (t) = xH (t) + xP (t) = Ce− τ0 + f0 (6.25)
The value of C is found from the initial condition
x (0) = 0 = C + f0 =⇒ C = −f0 (6.26)
so that the response to a constant input with zero initial condition is
h t
i
x (t) = f0 1 − e− τ0 u (t) (6.27)

1
0.865

0.632
x(t/ 0)/f0

0
0 1 2
Time t/ 0

FIGURE 6.3
First-order system response to constant input.

The response is shown in Figure 6.3. The value of x (t) exponentially approaches the
steady-state value of x (∞) = f0 . The time constant dictates the rate at which x (t) ap-
proaches its steady-state value. When t = τ0 , the terms inside the square brackets in the
above equation have the value 1 − e−1 = 0.632, and when t = 2τ0 , the amplitude becomes
x (2τ0 ) = f0 1 − e−2 = 0.865f0 . A system with a larger time constant reaches its final (or
steady-state) value slower.
We next obtain the response of a first-order system using the Laplace transform. The
describing equation is
1
ẋ (t) + x (t) = f (t) x (0) = x0 (6.28)
τ0
The Laplace transform of the differential equation has the form
1
sX (s) − x0 + X (s) = F (s) (6.29)
τ0
where F (s) is the Laplace transform of f (t). Upon rearranging,
 
1
X (s) s + = x0 + F (s) (6.30)
τ0
Solving for X (s) gives
x0 F (s)
X (s) = + (6.31)
s + τ10 s + τ10
242 Applied Dynamics

The inverse Laplace transform of the first term of the right side of Equation (6.31) is
recognized as
!
x0 t
L −1
1 = x0 e− τ0 (6.32)
s + τ0

The second term on the right side of Equation (6.31) can be written as
F (s)
= F (s) S (s) (6.33)
s + τ10
1
in which S (s) = s+ τ1
, whose inverse Laplace transform is
0

t
L−1 (S (s)) = e− τ0 (6.34)
Using the convolution integral in Equation (6.11), the inverse Laplace transform of the
second term in Equation (6.31) becomes
! Z t
F (s) τ
L−1
= f (t − τ )e− τ0 dτ (6.35)
s + τ10 0

where we note that τ and τ0 are different quantities. The total response has the form
! Z t
− τt F (s) − τt τ
x (t) = x0 e 0 + L −1
= x 0 e 0 + f (t − τ )e− τ0 dτ (6.36)
s + τ10 0

Example 6.4
Using a Laplace transform formulation, obtain the response of the first-order system
1 f0
ẋ (t) + x (t) = x (0) = x0 [a]
τ0 τ0
where f0 is constant. Show that when the forcing is zero, the solution becomes the same as
Equation (6.27) and when the initial condition is zero, the solution is the same as Equation
(6.23).
Taking the Laplace transform of the differential equation yields
1 f0 1
sX (s) − x0 + X (s) = F (s) [b]
τ0 τ0 s
Upon rearranging,  
1 f0 1
X (s) s + = x0 + [c]
τ0 τ0 s
The next step is to solve for X (s), which gives
x0 f0 1
X (s) = 1 + τ
  [d]
s + τ0 0 s s+ 1
τ0

To find the inverse Laplace transform, we need to write the second term in Equation [d]
into a form whose inverse Laplace transform is easily recognizable. Expanding it as
 
1
f0 1 A B A s + τ0 + Bs
  = + 1 =   [e]
τ0 s s + 1 s s + τ0 s s+ 1
τ0 τ0
Response of Dynamical Systems 243

Equating the numerators, we obtain for order s0 terms A = f0 and for order s1 terms
B = −A = −f0 . Substituting these terms into the above equation and calculating the
inverse Laplace transforms from Table 6.1, the response is calculated to be
h t
i
x (t) = f0 + (x0 − f0 ) e− τ0 u (t) [f ]

When the initial condition x0 = 0, the solution is the same as Equation (6.23) and when
the forcing f0 = 0, the solution is the same as Equation (6.27).

6.6 Review of Complex Variables


It is convenient to use complex variables when analyzing the free or forced response of
systems. This section reviews basic complex variable mathematics.
A complex number is one that has a real and a complex part and is defined as c = a + ib,
where c is the complex number,
√ a is the real part, and b is the imaginary part. The complex
number i is defined as i = −1. Consider a plane, which will be referred to as the complex
plane, where the horizontal coordinate denotes the real axis and the vertical coordinate
denotes the imaginary axis, as shown in Figure 6.4. A complex number can be expressed as
a point on the complex plane. Figure 6.4 shows the complex number c = 4 + i3.

#$%&'(%)*!+",
!

" -.%/0+#,

FIGURE 6.4
The complex number c = 4 + i3 in the complex plane.

The complex conjugate of a complex number c = a + ib is denoted by an overbar (c̄)


and is defined as

c̄ = a − ib (6.37)

When a complex number is multiplied by its conjugate, the result is

cc̄ = (a + ib) × (a − ib) = a2 + b2 (6.38)

The magnitude M of a complex number is defined as


√ p
M = cc̄ = a2 + b2 (6.39)

The magnitude of a complex number is the length of the line joining the origin of the
complex plane and the complex number. A complex number can also be expressed in terms
244 Applied Dynamics

of its magnitude and the angle the line joining the origin of the complex plane makes with
the real axis. Denoting this angle by θ we conclude that
b
tan θ = (6.40)
a
Let us next evaluate the expression eiθ , which is widely used in analysis. Introducing the
Taylor series expansion of eiθ and recalling the Taylor series of the sine and cosine functions
2 3
(iθ) (iθ)
eiθ = 1 + iθ + + + ...
2! 3!

θ2 θ4 θ3 θ6
 
= 1− + + ... + i θ − + + ... = cos θ + i sin θ (6.41)
2! 4! 3! 6!

It follows from the derivation above that the complex conjugate has the form

e−iθ = cos θ − i sin θ = eiθ (6.42)

Considering Equations (6.41) and (6.42), the sine and cosine functions can be expressed
in terms of eiθ and e−iθ as
1 iθ 1 iθ
e + e−iθ e − e−iθ
 
cos θ = sin θ = (6.43)
2 2i
Because b = M sin θ and a = M cos θ, the complex number c can be expressed in terms of
its magnitude M and angle θ as

c = M cos θ + iM sin θ = M (cos θ + i sin θ) = M eiθ (6.44)


1
In response analysis, complex number terms in the form a+ib are frequently encountered.
We can rewrite such a number so that the complex quantity appears in the numerator:
1 1 a − ib a − ib
= × = 2 (6.45)
a + ib a + ib a − ib a + b2

6.7 Second-Order Systems


Most dynamical systems are described by second-order differential equations. This is be-
cause force or moment balances are used to obtain the equations of motion and accelerations,
which involve second derivatives. The previous chapter discussed how to obtain the equa-
tions of motion and to linearize the equation of motion about an operating point. As a
representative model of a linear or linearized second-order system, we will use the mass-
spring-damper system in Figure 6.5a.
A one-degree-of-freedom stable second-order dynamical system can be viewed as com-
posed of three components:
• A mass component which stores kinetic energy and generates the inertia forces (resis-
tance to acceleration).
• A stiffness component which stores potential energy and generates restoring forces (re-
sistance to displacement, thus pushing the system back to equilibrium).
Response of Dynamical Systems 245
!" #"
%
!"#$ %‫"ܪ‬#$
' !"#$
&("#$
&

("#$
("#$

FIGURE 6.5
a) Mass-spring-damper system, b) free-body diagram.

• A damping component which dissipates energy. The damping component approximates


the different types of energy dissipation studied in Chapters 4 and 5, such as viscous,
friction, drag, and hysteresis. The model used here is based on viscous damping.
The motion can be viewed as an exchange between the kinetic and potential energies,
during which time energy gets dissipated by the damping component. The system in Figure
6.5a is a generic linear or linearized model, with m, c, and k denoting the mass, stiffness,
and damping, respectively, with the associated damping force as cẋ (t) and the spring force
as kx (t). Any linear or linearized system can be cast in the form above, with the coefficients
referred to as equivalent mass meq , equivalent damping ceq , and equivalent stiffness keq .
The free-body diagram is shown in Figure 6.5b. The motion variable x (t) is measured
from the undeformed position of the spring. Summing forces along the line of motion gives
+
X
→ F = ma = mẍ (t) = −cẋ (t) − kx (t) + F (t) (6.46)

and, rearranging so that the motion variable and its derivatives are on the left side, the
equation of motion is obtained as
mẍ (t) + cẋ (t) + kx (t) = F (t) (6.47)
The initial conditions are denoted by x (0) = x0 , ẋ (0) = v0 . In view of the stability
results in Section 5.8, the interest is in the case when m > 0, c ≥ 0, and k ≥ 0. The free
motion of such systems will have either constant or decaying amplitudes.
Because of the several cases that are of interest, the solution of second-order systems will
be considered separately for different situations. We will begin with free response, that is,
response to zero external excitation, followed by response to forcing. The chapter concludes
with the important case of harmonic excitations and steady-state response.

Example 6.5
Consider the rod in Figure 5.17. Linearize the equation of motion about the equilibrium
position θe = 0 and identify the equivalent mass, stiffness, and damping coefficients.
In Example 5.5, the equation of motion was shown to be
L
IO θ̈ + kL2 sin θ cos θ − mg sin θ = 0 [a]
2
Because the linearization is about θe = 0, we can use the small angle approximation of
sin θ ≈ θ and cos θ ≈ 1, which results in the linearized equation of motion in the form
 
L
IO θ̈ + kL2 − mg θ = 0 [b]
2
246 Applied Dynamics

The equivalent mass and stiffness terms are identified as


 
2 mgL
meq = IO keq = kL − ceq = 0 [c]
2
It should be noted that the equivalent mass, stiffness, and damping are not unique
themselves but their ratio is unique. We can multiply the equation of motion by an arbitrary
constant b, and the equivalent mass and stiffness become bIO and b kL2 − mg L2 . When


b = I1O , the equivalent mass becomes unity, meq = 1, and the equivalent stiffness becomes

kL2 mgL
keq = − [d]
IO 2IO

a) b)
Lsin
c k

FD FS

m, L
G
mg

O O
Oh
Ov

FIGURE 6.6
a) Rod with added damper, b) free-body diagram.

Next, replace the spring on the right in Figure 5.17 by a damper, let the remaining
spring be of constant k and obtain the equivalent damping constant. The rod and free-body
diagram are shown in Figure 6.6. Let us concentrate only on the moment about the base
that the damping force causes. The damping force is
d
FD = c (L sin θ) = cLθ̇ cos θ [e]
dt
so that the moment of the damping force about the base becomes

MOD = −FD L cos θ = −cL2 θ̇ cos2 θ [f ]

and the equation of motion becomes


L
IO θ̈ + cL2 θ̇ cos2 θ + kL2 sin θ cos θ − mg sin θ = 0 [g]
2
Linearizing the damping expression about the equilibrium position θ = 0, θ̇ = 0 gives

MOD = −cL2 θ̇ [h]

and the equivalent damping constant is recognized as ceq = cL2 .


Response of Dynamical Systems 247

6.8 Free Response of Undamped Second-Order Systems


We begin the study of second-order systems by analyzing the free response, that is, response
in the absence of any external excitations. The undamped case, that is, c = 0, is the focus
of this section. The next section introduces damping.
In the absence of damping and external excitations, the equation of motion and initial
conditions become

mẍ (t) + kx (t) = 0 x (0) = x0 ẋ (0) = v0 (6.48)

It
p is customary to divide the equation of motion by m and to define the quantity ωn =
k/m, which is called the natural frequency or frequency of oscillation. The unit of natural
frequency is 1/time. The natural frequency is dependent only on the parameters of the
system and not on the initial conditions. The equation of motion is rewritten as

ẍ (t) + ωn2 x (t) = 0 (6.49)

Consider a solution in the form x (t) = Ceλt , where C and λ are not yet known. Intro-
ducing this solution into the equation of motion and collecting terms gives

λ2 + ωn2 Ceλt = 0

(6.50)

For the above equation to have a nontrivial (nonzero) solution, Ceλt 6= 0. Hence, the
only way the above equation can be satisfied is if the terms inside the brackets vanish. The
resulting equation

λ2 + ωn2 = 0 (6.51)

is called the characteristic equation. The characteristic equation is an algebraic equation in


the form of a polynomial of the same order as the differential equation of motion, in this
case two, in λ. Solving for the roots of the characteristic equation gives

λ = ±iωn (6.52)

There are two pure imaginary roots. The response can be written as

x (t) = C1 eiωn t + C2 e−iωn t (6.53)

Note that eiωn t and e−iωn t are complex conjugates. Hence, for the above equation to be
real-valued, C1 and C2 must also be complex conjugates. It is convenient to write them as
1 −iφ 1 iφ
C1 = Ae C2 = Ae (6.54)
2 2
in which A and φ are real-valued quantities, referred to as the amplitude and phase angle,
respectively. The response can then be expressed as
A h i(ωn t−φ) i
x (t) = e + e−i(ωn t−φ) = A cos (ωn t − φ) u (t) (6.55)
2
The response is in the form of a constant amplitude sinusoidal with a phase shift, in
essence vibration with constant amplitude. The values of A and φ depend on the initial
conditions. Setting t = 0 in the above equation gives

x (0) = x0 = A cos (−φ) = A cos φ (6.56)


248 Applied Dynamics

Differentiating Equation (6.55) with respect to time results in

ẋ (t) = −ωn A sin (ωn t − φ) u (t) (6.57)

and at time t = 0 the value of the derivative becomes

ẋ (0) = v0 = −ωn A sin (−φ) = Aωn sin φ (6.58)

We solve Equations (6.56) and (6.58) simultaneously to obtain the amplitude and phase
angle in terms of the initial conditions as
s  2  
2 v0 −1 v0
A = x0 + φ = tan (6.59)
ωn ωn x0

The inverse tangent equation has two solutions. You should select the value of φ such that
the values of x0 and v0 are obtained when Equation (6.55) and its time derivative are
evaluated at t = 0.

x(t)
x0 T=2 n

t0 t1 t2 t
n

FIGURE 6.7
Free response of an undamped second-order system.

The response, plotted in Figure 6.7, is harmonic. Let us denote its period of the motion
by T and find its value. Consider a point in time, say, t1 , and a second point t2 = t1 + T
one cycle later. Setting x(t1 ) = x(t2 ) in the response equation we obtain

A cos (ωn t1 − φ) = A cos (ωn t2 − φ) = A cos (ωn t1 + ωn T − φ) (6.60)

from which we conclude that ωn T = 2π, so the period of oscillation is recognized as



T = (6.61)
ωn
We can show that the time it takes to reach the first peak is t0 = φ/ωn . Another
commonly used quantity to describe harmonic motion is the frequency, fn , defined as
1 ωn
fn = = (6.62)
T 2π
The unit of frequency is cycles per second (cps) or Hertz (Hz), named after the German
physicist Heinrich Rudolf Hertz (1857–1894).
Next, derive the response in terms of the initial displacement and velocity, x0 and v0 .
Response of Dynamical Systems 249

We can do this by manipulating Equation (6.55) but, for illustrative purposes, let us use the
Laplace transform solution. Going back to Equation (6.49) and taking its Laplace transform
yields

L ẍ (t) + ωn2 x (t) = s2 X (s) − sx0 − v0 + ωn2 X (s) = 0



(6.63)

in which X (s) = L (x (t)) is the Laplace transform of x (t). Collecting like terms, the
expression for X (s) becomes
s 1
X (s) = x0 + v0 2 (6.64)
s2 + ωn2 s + ωn2

We obtain the response by the inverse Laplace transform x (t) = L−1 (X (s)). As dis-
cussed earlier, it is convenient to consult Laplace transform tables to ascertain the inverse
transform. The response has the form
v0
x (t) = x0 cos ωn t + sin ωn t (6.65)
ωn

Example 6.6
An undamped mass-spring system has mass of 10 kg and stiffness of 90 N/m. The initial
conditions are x0 = −0.1 m and v0 = 0.6 m/s. Calculate the amplitude and phase angle of
the ensuing motion.
The natural frequency is
r r
k 90
ωn = = = 3 rad/s [a]
m 10
From Equation (6.59), the amplitude of vibration is
s  2 s  2
2 v 0 2
0.6
A = x0 + = 0.1 + = 0.2236 m [b]
ωn 3

The phase angle is found from


   
v0 0.6
φ = tan−1 = tan−1 = tan−1 (−2) = −1.107, 2.034 rad [c]
ωn x0 3 × −0.1

We need to pick the correct value of the phase angle. Consider the first value, φ = −1.107
rad. Checking the relationship x0 = A cos φ, we get

x0 = −0.1 A cos φ = 0.2236 cos (−1.107) = 0.1 [d]

which are not the same. Hence, the correct phase angle is φ = 2.034 rad.

Example 6.7—Measuring Mass in Outer Space


A critical task of space scientists is to monitor the health of astronauts. Astronauts go
through frequent examinations, such as measuring pulse, body temperature, and blood
pressure. One physical property that cannot be directly measured in space is the mass of an
astronaut. Weight loss due to extended stay in space, especially in the form of bone mass
loss, is of concern. Skylab astronauts lost about three to six pounds of weight per month
while in space.
The mass of Skylab astronauts was measured by the body mass measurement device
250 Applied Dynamics

mA
k
mC

FIGURE 6.8
Schematic of body mass measurement device.

(BMMD), which consists of a chair of mass mC that travels on rails and is attached to a
rigid wall by a spring of constant k. A sketch is shown in Figure 6.8. The astronaut, of mass
mA , is strapped to the chair, and pthe chair is set into motion. The natural frequency of the
chair-astronaut assembly is ωn = k/(mC + mA ), and the period of oscillation is

2π 2π mC + mA
T = = √ [a]
ωn k
Once the period of oscillation is measured, the mass of the astronaut is calculated as
kT 2
mA = − mC [b]
4π 2
The derivations in the previous section assume that the mass is rigid and does not
change shape. Any movement, even the breathing of the astronaut, could affect vibrational
properties of the BMDD. The astronauts would be strapped tightly to the chair, body and
feet, they would hold a bar to remain stiff and were told to not breathe while the period of
oscillation was measured.
The International Space Station uses a different and more accurate device called Space
Linear Acceleration Mass Measurement Device (SLAMMD) to measure mass.1

6.9 Free Response of Damped Second-Order Systems


This section introduces damping to the equation of motion and obtains the free response
(no external excitation). The equation of motion and initial conditions are
mẍ (t) + cẋ (t) + kx (t) = 0 x (0) = x0 ẋ (0) = v0 (6.66)
Dividing the equation of motion by the mass gives
c k
ẍ (t) + ẋ (t) + x (t) = 0 (6.67)
m m
p
The natural frequency, ωn = k/m, was defined earlier. Here, we introduce the dimen-
sionless parameter ζ, referred to as the damping factor and defined as
c c c
= 2ζωn ζ = = √ (6.68)
m 2mωn 2 km
1 See http://www.nasa.gov/mission pages/station/science/experiments/SLAMMD.html for more details.
Response of Dynamical Systems 251

The equation of motion, in terms of the damping factor ζ, becomes

ẍ (t) + 2ζωn ẋ (t) + ωn2 x (t) = 0 (6.69)

Consider a response in the form x (t) = Ceλt , introduce it to the above equation and
collect like terms, with the result

λ2 + 2ζωn λ + ωn2 Ceλt = 0



(6.70)

Following the same line of argument that Ceλt 6= 0 for a nontrivial solution, the character-
istic equation is obtained as

λ2 + 2ζωn λ + ωn2 = 0 (6.71)

The characteristic equation is a polynomial of order two in λ. Its solution is


p
λ1,2 = −ζωn ± ωn ζ 2 − 1 (6.72)

and the general form of the response can be written as

x (t) = C1 eλ1 t + C2 eλ2 t (6.73)

The nature of the response depends on the roots of the characteristic equation. The
damping factor dictates the nature of the motion. The following different cases can be
identified:
• When ζ = 0, the roots are pure imaginary, λ1,2 = ±iωn . This case is referred to as the
undamped case which was discussed in the previous section. The response is oscillatory
with constant amplitude.

• When 0 < ζ < 1, the roots are complex with real negative parts. This case is referred
to as the underdamped case. The response is an exponentially decaying sinusoidal.
• When ζ = 1, the roots are real, negative, and repeated: λ1,2 = ± − ωn . The response
is a decaying exponential and has no oscillatory component. This case is known as the
critically damped case, as it is the limit at which oscillatory motion ceases.

• When ζ > 1, there are two real distinct negative roots. This case is called overdamped.
The motion is aperiodic, and it has the form of a decaying exponential.
• The case when ζ < 0 results in exponential growth (periodic or aperiodic) of the motion
and is descriptive of an unstable system, as discussed in Section 5.8. Having ζ < 0
corresponds to a negative damping coefficient. Unstable systems are sometimes called
negatively damped.

6.10 Underdamped Systems


The underdamped case is the most commonly encountered form of damped systems. Their
response is exponentially decaying oscillation. Vehicle suspension systems or any other vi-
bration absorption systems are designed such that there will be some oscillation in the
response.
252 Applied Dynamics

Because in underdamped systems ζ < 1, we can write the roots of the characteristic
equation as
p
λ1,2 = −ζωn ± iωn 1 − ζ 2 (6.74)
p
Define the damped natural frequency as ωd = ωn 1 − ζ 2 so that

λ1,2 = −ζωn ± iωd (6.75)

and write the response as

x (t) = C1 eλ1 t + C2 eλ2 t = C1 e(−ζωn +iωd )t + C2 e(−ζωn −iωd )t (6.76)

Noting the identity ea+b = ea eb , Equation (6.76) can be written as

x (t) = e−ζωn t C1 eiωd t + C2 e−iωd t



(6.77)

The terms inside the parentheses in Equation (6.77) are similar to Equation (6.53),
which was obtained when analyzing the response of undamped second-order systems. We
can make use of that analysis to conclude that C1 and C2 are complex conjugates and that
the terms inside the parentheses can be expressed in the form of Equation (6.55). Thus, the
response can be written as

x (t) = Ad e−ζωn t cos (ωd t − φd ) u (t) (6.78)

where Ad and φd are the amplitude and phase angle of the motion, respectively. Their values
depend on the initial conditions. At t = 0

x (0) = x0 = Ad cos (−φd ) = Ad cos (φd ) (6.79)

Differentiating Equation (6.78) and evaluating at t = 0 gives

ẋ (0) = v0 = ωd Ad sin (φd ) − Ad ζωn cos φd (6.80)

The above two equations can be solved for Ad and φd in terms of the initial conditions.
The result is
s  2  
ζωn x0 + v0 ζωn x0 + v0
Ad = x20 + φd = tan−1 (6.81)
ωd ωd x0

As discussed earlier, the inverse tangent has two solutions.


The response can be viewed as a sinusoidal that is enveloped by an exponentially de-
caying function, as shown in Figure 6.9. The amount of decay depends on the amount of
damping. The period of oscillation is denoted by Td , whose value is Td = ω2πd .
It is of interest to quantify the amount of decay during a cycle. To this end, denote the
time at which one of the peaks occurs by t = t1 . The next peak occurs at t = t2 = t1 + Td .
The values of the peaks from the response equation are

x(t1 ) = Ad e−ζωn t1 cos (ωd t1 − φd ) x(t2 ) = Ad e−ζωn t2 cos (ωd t2 − φd ) (6.82)

Noting that

cos (ωd t2 − φd ) = cos (ωd (t1 + Td ) − φd )

= cos (ωd t1 + 2π − φd ) = cos (ωd t1 − φd ) (6.83)


Response of Dynamical Systems 253

x(t)
Ad ––
x0
Ad e – nt

t2
t1
t

Td
–Ad –
0 1 2 3 4 5 6 7 8 9
Time

FIGURE 6.9
Free response of an underdamped (0 < ζ < 1) system.

the ratio between the two peaks becomes


x(t1 ) Ad e−ζωn t1 cos (ωd t1 − φd )
=
x(t2 ) Ad e−ζωn t2 cos (ωd t2 − φd )

e−ζωn t1 e−ζωn t1 ζωn Td


√2πζ
1−ζ 2
= = = e = e (6.84)
e−ζωn t2 e−ζωn (t1 +Td )
The ratio of any two consecutive peaks depends only on the damping factor ζ. It does
not depend on the mass, stiffness, period of oscillation, or initial conditions. This property
leads to a method for estimating the damping, as will be described shortly.
When designing a damper, such as in a suspension system, we select the damping factor
such that the response amplitude dies out in a desired number of cycles. Figure 6.10 plots
the free response (x0 = 1, v0 = 0) for different values of the damping factor ζ, using a
natural frequency of ωn = 1 rad/s. Figure 6.11 plots the value of the amplitude ratio of two
consecutive peaks
x(tj+1 ) − √2πζ
= e 1−ζ2 (6.85)
x(tj )
for different values of the damping factor ζ.
Suspensions of sports cars are designed so that ζ > 0.3, and passenger cars have damping
factors in the range ζ = 0.2–0.3. From Figures 6.10 and 6.11, there is quite a difference in
the number of cycles it takes for the motion amplitudes to become small. For damping levels
of 0.6 or higher, almost the entire motion amplitude subsides after half a cycle. This is not
desirable, except for race cars, as it makes the response almost aperiodic. It is preferable for
most suspension systems to have some periodicity, so they attenuate the vibrational motion
over a cycle or two.
Next, let us obtain the free response using the Laplace transform. Taking the Laplace
transform of Equation (6.69) and collecting like terms gives
s2 + 2ζωn s + ωn2 X (s) − sx0 − v0 − 2ζωn x0 = 0

(6.86)
254 Applied Dynamics

1
ζ = 0.05
0.8 ζ = 0.1
ζ = 0.2
0.6
ζ = 0.3
0.4 ζ = 0.5

0.2
x(t)

−0.2

−0.4

−0.6

−0.8
0 2 4 6 8 10 12 14 16 18 20
t

FIGURE 6.10
Response for different values of the damping factor ζ: 0.05, 0.1, 0.2, 0.3, 0.5.

in which X (s) = L (x (t)) is the Laplace transform of x (t). Solving for X (s) gives

s + 2ζωn 1
X (s) = x0 + v0 2 (6.87)
s2 + 2ζωn s + ωn2 s + 2ζωn s + ωn2

We can calculate the response x (t) = L−1 X (s), the inverse Laplace transform of X (s),
from Table 6.1, with the result
!
−ζωn t ζ 1 −ζωn t
x (t) = x0 e cos ωd t + p sin ωd t + v0 e sin ωd t (6.88)
1 − ζ2 ωd

Example 6.8
Given an underdamped system with ωn = 5 rad/s, ζ = 0.1, x0 = 1 cm, v0 = 0, obtain the
response amplitude Ad and phase angle φd .
The damped natural frequency is
p √
ωd = ωn 1 − ζ 2 = 5 0.99 = 4.975 rad/s [a]

The initial conditions give

x (0) = x0 = 1 = Ad cos φd ẋ (0) = v0 = 0 = Ad ωd sin φd − Ad ζωn cos φd [b]

The second equation above can be solved for the phase angle as
   
−1 ζωn −1 0.1 × 5
φd = tan = tan = 0.1005, 3.242 rad [c]
ωd 4.975
Response of Dynamical Systems 255

$1
"#+0.9
!! "#*0.8
!1
"#)0.7
"#(0.6
"#'0.5
"#&0.4
"#%0.3
"#!0.2
"#$0.1
" 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
" "#$ "#! "#% "#& "#' "#( "#)
ȗ

FIGURE 6.11
xj+1
Amplitude ratios xj for two consecutive peaks.

By checking with the initial conditions, we can show that the correct phase angle is
0.1005 rad. In degrees, φd = 0.1005 × 180/π = 5.74◦ . The amplitude can be obtained by
1 1
Ad = = = 1.005 cm [d]
cos φd cos 5.74◦
Noting that ζωn = 0.5, the response is

x (t) = 1.005e−0.5t cos (4.975t − 0.1005) cm [e]

6.11 Damping Estimation by Logarithmic Decrement


Many times, the mass, stiffness, or damping parameters of a system are not known accu-
rately. It is of interest to find ways of measuring or estimating these parameters. Measure-
ment of mass or stiffness can be carried out relatively easily, as both parameters can be
estimated using static (or simple dynamic) methods, such as weighing, or applying a known
force and measuring the displacement. By contrast, damping forces depend on velocity, so
we cannot use static methods to estimate damping.
The logarithmic decrement method is a convenient way of estimating the amount of
damping. The method works well when there are no external excitations and viscous damp-
ing is the only (or most significant) form of energy dissipation. The method takes advantage
of the results in the previous section which show that for free motion, the amplitude ratios
of any two consecutive peaks depend only on the damping factor.
Consider the plot of the free response of an underdamped system, such as the one shown
in Figure 6.12 and denote the peak values of the plot by x1 , x2 , . . . , and the times where the
peaks occur as t1 , t2 , . . . . For any two consecutive peaks, the amplitude ratios are given by
Equation (6.84). For any two consecutive peaks xj and xj+1 (j = 1, 2, . . . ) the logarithmic
256 Applied Dynamics

!!""

#$

!#

!$ !&
!%
"
"$ "% "&

FIGURE 6.12
Peak amplitudes of an underdamped system.

decrement δ is defined as the natural logarithm of the amplitude ratio


 
xj 2πζ
δ = ln = p (6.89)
xj+1 1 − ζ2
Once the amplitudes of any two consecutive peaks are measured, we can calculate the
logarithmic decrement and use Equation (6.89) to solve for the damping factor, with the
result
δ
ζ = √ (6.90)
4π 2 + δ2
When the damping factor is small, we can approximate the above equation by
δ
ζ ≈ (6.91)

In general, we prefer to conduct the logarithmic decrement analysis with the first few
peaks. Measurements of peaks with higher amplitudes are usually more accurate than peaks
with lower amplitudes. Also, we can check the validity of the viscous damping model by
estimating damping using several pairs of consecutive peaks and comparing the estimates.
The logarithmic decrement method can be extended to the case of nonconsecutive peaks.
Noting that the amplitude ratio of any two consecutive peaks is the same,
x1 x2 xj
= = ... = (6.92)
x2 x3 xj+1
for any two peaks separated by l cycles, we can write
 l
xj xj
= j = 1, 2, . . . (6.93)
xj+l xj+1
It follows that the logarithmic decrement for l cycles can be expressed as
 
1 xj
δ = ln (6.94)
l xj+l
Response of Dynamical Systems 257

The damping in a system, even for a well-designed damper, is almost never purely
viscous. Effects such as friction, hysteresis, and air resistance also play a role in dissipating
energy. Under such circumstances, the amplitude ratios will be different. A procedure to
estimate damping under such circumstances is to average the peak ratios. Denoting the
logarithmic decrement for two consecutive peaks j and j + 1 by
xj
δj = ln j = 1, 2, . . . (6.95)
xj+1
and measuring n pairs of peaks and calculating their ratios, the logarithmic decrement can
be approximated by
n
1X
δ = δj (6.96)
n j=1

Another approach to estimate the logarithmic decrement is to plot the natural logarithms
of the peak amplitudes and conduct a curve fit analysis to obtain an average decay rate.

Example 6.9
After three cycles, the peak amplitude of a mass-spring-damper system drops to 80% of its
original value. Calculate the amplitude ratio after five cycles. Given m = 4 kg and k = 128
N/m, calculate the damping coefficient c.
Denoting the initial peak amplitude by x1 , after three cycles the amplitude is x4 = 0.8x1 .
From Equation (6.94) the logarithmic decrement for l = 3 is
 
1 x1 1 1
δ = ln = ln = 0.07438 [a]
l xl+1 3 0.8
Because the logarithmic decrement is quite small, the damping constant can be found using
δ 0.07438
ζ = = = 0.01184 [b]
2π 2π
The exact value, obtained from Equation (6.90), also turns out to be
δ 0.07438
ζ = √ = √ = 0.01184 [c]
4π 2 + δ2 4π 2+ 0.074382
To find the amplitude ratio in terms of the logarithmic decrement, we can rearrange
Equation (6.94) with the result
x1
= eδl [d]
xl+1
For l = 5, the above ratio becomes
x1
= e5×0.07438 = 1.4505 [e]
x6
so the response amplitude ratio after five cycles becomes
1
x6 = x1 = 0.6894x1 [f ]
1.4505
q
c k
Using the relationships m = 2ζωn and ωn = m , the damping coefficient is calculated
as r r
k 128
c = 2mζωn = 2mζ = 2 × 4 × 0.01184 × = 0.5358 N·s/m [g]
m 4
258 Applied Dynamics

Example 6.10
The measured values of four consecutive peak amplitudes are given as 1, 0.8, 0.65, 0.50.
Estimate the damping factor.
The logarithmic decrements for the three cycles are
1 0.8 0.65
δ1 = ln = 0.2231 δ2 = ln = 0.2076 δ3 = ln = 0.2624 [a]
0.8 0.65 0.50
The average for the three cycles gives an estimate for the logarithmic decrement as
0.2231 + 0.2076 + 0.2624
δ ≈ = 0.2301 [b]
3
so that the estimated damping factor is ζ ≈ δ/2π = 0.0366.
Let us compare the accuracy of this approximation by calculating the peak amplitudes
for when x1 = 1 and δ = 0.2301. From Equation (6.89) the amplitudes of two consecutive
peaks become
xj+1 = xj e−δ j = 1, 2, . . . [c]
In the case here, e−δ = e−0.2301 = 0.7945, so that

x2 = 0.7945x1 = 0.7945 x3 = 0.7945x2 = 0.6312 x4 = 0.7945x3 = 0.5015 [d]

which are quite close to the measured values.

6.12 Response to an Impulsive Force


We next consider the response of dynamical systems that are subjected to external excita-
tions and terms on the right side of the equations of motion. We begin with a special type
of forcing, which is in the form of a very large force applied over a very short time period.
Such an excitation is known as an impulsive force, as discussed in Section 5.4.1. Bouncing
of a ball and impact of two bodies are examples of impulsive motion.
R In dynamics, the term
impulse is defined as the integral of force over time, that is, F (t) dt.

!""#
ǻ
!!

"! "!ǻ "

FIGURE 6.13
Approximation of an impulsive force.

Consider Figure 6.13, which depicts a large force F0 applied at t = t0 for a very short
Response of Dynamical Systems 259

period of time, denoted by ∆. The resulting impulsive force F̂ has the form F̂ = F0 ∆.
The units of impulsive force are force×time or mass×velocity. To obtain the response to
an impulsive force, use is made of the developments of Chapter 5, where we show that an
impulsive force causes a sudden change in velocity, with no change in position.
The impulse-momentum theorem for a particle of mass m moving in one direction (x)
becomes
Z t2 X
mẋ(t1 ) + F (t) dt = mẋ(t2 ) (6.97)
t1

For a very large force F0 applied over a very short period of time, we can neglect
the effects of other forces acting on the body, such as springs, damping, or gravity, as the
magnitudes of such forces are small compared to the impulsive force. Also, t2 = t0 +∆ ≈ t+0,
which gives
Z t0 +∆ X
F (t) dt ≈ F0 ∆ = F̂ (6.98)
t0

so that the impulse-momentum relationship for an impulsive force applied at t = t0 becomes

F0 ∆ F̂
ẋ(t+
0 ) − ẋ (t0 ) = = (6.99)
m m
Consider a mass-spring-damper system

F (t)
ẍ (t) + 2ζωn ẋ (t) + ωn2 x (t) = , x (0) = 0, ẋ (0) = 0 (6.100)
m

and subject the system to an impulsive force F (t) = F̂ δ (t) at t = 0. Immediately after the
impulsive force is applied, that is, at t = 0+ , the initial conditions become


x 0+ ẋ 0+
 
= 0 = v0 = (6.101)
m
and, since there no longer is an applied force, F (t) = 0, t > t+ . Hence, the response of a
system at rest that is subjected to an impulsive force F̂ can be treated as a free motion

problem with initial velocity v0 = m .
From Section 6.10, the response to an initial velocity is x (t) = ωv0d e−ζωn t sin ωd t. Substi-

tuting the initial velocity v0 = m, the response to an impulsive force has the form

F̂ −ζωn t
x (t) = e sin ωd t u (t) (6.102)
mωd
p
where ωd = ωn 1 − ζ 2 is the damped natural frequency. This result can be extended to
the case when the amplitude of the mass-normalized impulsive force fˆ = m
F̂ F̂
is unity, m = 1,
by calling it the impulse response g (t) as
1 −ζωn t
g (t) = e sin ωd t (6.103)
ωd
For undamped systems the impulse response has the form
1
g (t) = sin ωn t (6.104)
ωn
260 Applied Dynamics

1
undamped
damped
g(t)ωn

−1
0
t

FIGURE 6.14
Impulse response g (t) at t = t0 = 0 for an undamped system and a damped system.

The impulse response g (t), for an impulse applied at t = t0 , is shown in Figure 6.14 for
an undamped system and a damped system.
The idealized mathematical model of an impulsive force is in terms of the Dirac delta
function. From the discussion in Section 6.2, we can express an impulsive force F̂ applied
at time t = t0 as a continuous function of time by means of the Dirac delta function:
F (t) = F̂ δ (t − t0 ) = F0 ∆δ (t − t0 ) (6.105)
Next, let us obtain the response to an impulsive force by the Laplace transform method.
The equation of motion and initial conditions are

ẍ (t) + 2ζωn ẋ (t) + ωn2 x (t) =
δ (t) , x (0) = 0, ẋ (0) = 0 (6.106)
m
The Laplace transform of the equation of motion is

s2 + 2ζωn s + ωn2 X (s) =

(6.107)
m
where X (s) = L(x (t)). Solving for X (s) gives
F̂ 1
X (s) = (6.108)
m s2 + 2ζωn s + ωn2
The Laplace transform of the impulse response g (t), which is denoted by G (s), has the
form
1
G (s) = 2 (6.109)
s + 2ζωn s + ωn2
so that X (s) = G (s) F̂ /m. It follows that the response x (t), which is the inverse Laplace
transform of X (s), becomes
F̂ F̂ 1 −ζωn t
x (t) = L−1 X (s) =g (t) = e sin ωd t (6.110)
m m ωd
The impulse response equations developed above are based on the assumption that the
duration of the impulse is very short. When dealing with real-life problems, we must be
careful to check that the assumption is realistic. Especially for systems whose response
is periodic, verify that the duration of the impulse is much shorter than the period of
T
oscillation. A rule of thumb is to use ∆ < 10 or, for damped systems, ∆ < T10d .
Response of Dynamical Systems 261

6.12.1 Response to Multiple Impulses

80

60
F(t)

40

20

0
0 1 2 3 4 5 6
t

FIGURE 6.15
Two impulsive excitations (applied at t1 = 1, t2 = 4.5).

An interesting application of the principle of superposition is the case where more than
one impulse is applied. Consider a second-order system that is at rest to which an excitation
consisting of two impulses, such as the one shown in Figure 6.15, is applied. The excitation
is expressed as
F (t) = F̂1 δ (t − t1 ) + F̂2 δ (t − t2 ) (6.111)
where F̂1 and F̂2 are the magnitudes of the impulses.
The excitation can thus be written as a summation of two excitations
F (t) = F1 (t) + F2 (t) (6.112)
in which F1 (t) = F̂1 δ (t − t1 ) and F2 (t) = F̂2 δ (t − t2 ). It follows from the superposition
principle that the response x (t) can be written as the superposition of the responses to the
individual forces as
x (t) = x1 (t) + x2 (t) (6.113)
where xi (t) (i = 1, 2) are the individual responses. For the first impulsive force F̂1 , the
corresponding response is x1 (t) = F̂m1 g (t − t1 ) u (t − t1 ).
The response to the second impulse, x2 (t), must reflect the fact that the impulse was
applied at time t = t2 . This response can be written as
F̂2
x2 (t) = g (t − t2 ) u (t − t2 ) (6.114)
m
Each response term has t − ti (i = 1, 2) both as the argument of the impulse response g
and the unit step function u. This is to indicate that x1 (t) begins at time t = t1 and x2 (t)
begins at time t = t2 . The total response is
F̂1 F̂2
x (t) = x1 (t) + x2 (t) = g (t − t1 ) u (t − t1 ) + g (t − t2 ) u (t − t2 ) (6.115)
m m
The response is plotted in Figure 6.16 for the case of m = 1, ωn = 1, ζ = 0, t0 = 0,
F̂1 = 1, t1 = 4.5, and F̂2 = 1.4. Notice the change in x (t) at t = 4.5 (rather, in its
derivative), when the second impulse is applied.
262 Applied Dynamics

1
x(t)

−1

−2
0 5 10 15
t

FIGURE 6.16
Response to two impulses.

Example 6.11
An undamped mass-spring system (W = 50 lb, k = 100 lb/in) that is at rest is subjected
to a force of magnitude F0 = 1, 000 lb for a time period of ∆ = 0.01 seconds. Calculate
the response assuming the force is impulsive and assess the validity of the impulsive force
assumption.
The magnitude of the impulsive force is
F̂ = F0 ∆ = 1000 × 0.01 = 10 lb·sec [a]
so that the mass normalized impulsive force becomes
F̂ 10 lb·sec
fˆ = = 1 sec2
= 6.434 ft/sec [b]
m 50 lb 32.17 ft

Hence, the initial velocity is


v0 = fˆ = 6.434 ft/sec [c]
The free response of an undamped system to an initial velocity is given by x (t) =
v0
ωn sin ωn t. The natural frequency is
r s
lb
k 100 in × 12 in
ft
ωn = = 1 sec2
= 27.786 rad/sec [d]
m 50 lb 32.17 ft

Hence, the response becomes


v0 6.434
x (t) = sin ωn t = sin (27.786t) = 0.2315 sin (27.786t) u (t) ft [e]
ωn 27.786
Next, let us examine the validity of the impulsive motion assumption. The period of
oscillation is
2π 2π
T = = = 0.2261 sec [f ]
ωn 27.786
and the ratio of the duration of the impulse to the period of oscillation is
∆ 0.01 1 1
= ≈ < [g]
T 0.2261 23 10
so that the impulsive motion assumption is realistic.
Response of Dynamical Systems 263

6.13 Step Response


Consider an undamped mass-spring system that is initially at rest which is subjected to a
force in the form of step input at t = 0 defined by F (t) = F0 u (t) and shown in Figure 6.17.

"!!"

"#

FIGURE 6.17
Step input.

The equation of motion and initial conditions are

mẍ (t) + kx (t) = F (t) = F0 u (t) x (0) = 0, ẋ (0) = 0 (6.116)

Let us obtain the response using the homogeneous plus particular solution approach, x (t) =
xH (t) + xP (t). Rewrite the equation of motion as
F0
ẍ (t) + ωn2 x (t) = u (t) x (0) = 0, ẋ (0) = 0 (6.117)
m
From earlier developments in this chapter, xH (t) = C1 sin ωn t + C2 cos ωn t. The par-
ticular solution has the same form as the forcing and can be written as xP (t) = E. The
amplitude of the particular solution, E, is determined by introducing the particular solution
to the equation of motion.
F0
ẍP (t) + ωn2 xP (t) = 0 + ωn2 E = (6.118)
m
which is solved for E as
F0 F0
E = 2
= (6.119)
mωn k
Because the force F0 is constant, the particular solution is, in essence, the static de-
flection, as in the static deflection of a spring on which a weight acts. The total response
becomes
F0
x (t) = xH (t) + xP (t) = C1 sin ωn t + C2 cos ωn t + (6.120)
k
The initial conditions are used to find the constants C1 and C2 . At t = 0
F0
x (0) = C2 + = 0 (6.121)
k
264 Applied Dynamics

so that C2 = −F0 /k. Differentiating Equation (6.120) gives

ẋ (t) = ωn C1 cos ωn t − ωn C2 sin ωn t (6.122)

and evaluating at t = 0 leads to

ẋ (0) = ωn C1 = 0 (6.123)

with the result C1 = 0. The total response becomes


F0 F0
x (t) = (1 − cos ωn t) u (t) = (1 − cos ωn t) u (t) (6.124)
k mωn2

1.5
s(t) ω2n

0.5

0
0 2 4 6 8 10 12 14 16 18 20
t (sec)

FIGURE 6.18 
Step response s(t) ωn2 of an undamped second-order system for ωn = 1.

The above result can be generalized by setting the amplitude of the mass-normalized
force Fm0 to unity and by calling the resulting response the step response s (t) as

1
s (t) = (1 − cos ωn t) u (t) (6.125)
ωn2

The step response is plotted in Figure 6.18. Note that the impulse response g (t) is the
derivative of the step response. Indeed, differentiating the step response gives

ds (t) 1 1
= 2 (ωn sin ωn t) u (t) = sin (ωn t) u (t) = g (t) (6.126)
dt ωn ωn

The step response for a damped second-order system is found in a similar way. The
procedure is lengthy so it is omitted here. The counterpart of Equation (6.124) for damped
systems is
  
F0 ζωn
x (t) = 1 − e−ζωn t cos ωd t + sin ωd t u (t) (6.127)
k ωd

The steady-state value of the displacement (that is, at infinity) is xss = x(∞) = F0 /k.
As discussed earlier, the steady-state value turns out to be the static deflection. The step
Response of Dynamical Systems 265

s(t) ω2n 1.5

0.5

0
0 2 4 6 8 10 12 14 16 18 20
t (sec)

FIGURE 6.19 
Step response s(t)ωn2 of an underdamped second-order system for ωn = 1, ζ = 0.15.

response for a damped system is obtained by dividing the above equation by the mass-
normalized force Fm0 and has the form
  
1 ζωn
s (t) = 2 1 − e−ζωn t cos ωd t + sin ωd t u (t) (6.128)
ωn ωd

The step response for an underdamped system is plotted in Figure 6.19.


Next, let us obtain the step response of an undamped system using the Laplace transform
method. Taking the Laplace transform of Equation (6.117) leads to

F0 1
s2 + ωn2 X (s) =

(6.129)
ms
Solving for X (s) gives

F0 1
X (s) = (6.130)
m s (s + ωn2 )
2

One can use partial fractions to expand the term on the right as
1 A Bs + C
S (s) = = + 2 (6.131)
s (s2 + ωn2 ) s s + ωn2

which can be rewritten as



1 A s2 + ωn2 + Bs2 + Cs
S (s) = = (6.132)
s (s2 + ωn2 ) s (s2 + ωn2 )

Equating like terms in the numerator gives

s0 terms: Aωn2 = 1 s1 terms: C = 0 s2 terms: A+B = 0 (6.133)

which can be solved for A, B, and C as


1 1
A = B = − C = 0 (6.134)
ωn2 ωn2
266 Applied Dynamics

Introduction of the values of A, B, and C into Equation (6.131) gives


 
1 1 1 s
S (s) = = − (6.135)
s (s2 + ωn2 ) ωn2 s s2 + ωn2

and the Laplace transform of the response becomes


 
F0 1 1 s
X (s) = − (6.136)
m ωn2 s s2 + ωn2

The time response x (t) is obtained by using the inverse Laplace transform pairs from
Table 6.1, with the result
F0 1
x (t) = L−1 X (s) = (1 − cos ωn t) u (t) (6.137)
m ωn2

which, of course, is the same expression obtained in Equation (6.124).


Similar to the approach in the previous section, in the presence of several step excita-
tions, the response is obtained by superposition. The example that follows illustrates the
procedure.

Example 6.12
Consider an undamped mass-spring system with m = 4 kg, k = 36 N/m, subjected to the
constant force F0 = 7.2 N that is applied during the time period 0 < t < 5 seconds, as
shown in Figure 6.20. Obtain the response to zero initial conditions.

F(t)

F0

t0 t

FIGURE 6.20
Step excitation.

We can solve this problem two ways. The first approach is to obtain the response between
0 < t < t0 and then use the position and velocity at t = t0 = 5 s as p initial conditions for
the response when t > t0 . Noting that the natural frequency is ωn = k/m = 3 rad/s, in
the range 0 < t < t0 the response is
F0
x (t) = (1 − cos ωn t) u (t) = 0.2 (1 − cos 3t) u (t) 0<t<5 [a]
k
so that the position and velocity at t = t0 = 5 seconds are
F0
x (t0 ) = x(5) = (1 − cos ωn t0 ) = 0.2 (1 − cos 15) = 0.3519 m
k
Response of Dynamical Systems 267

F0
ẋ (t0 ) = ẋ(5) = ωn sin ωn t0 = 0.2 × 3 sin 15 = 0.3902 m/s [b]
k
The response for t > t0 is free response with initial conditions given in Equation [b]
 
ẋ (t0 )
x (t) = x (t0 ) cos ωn (t − t0 ) + sin ωn (t − t0 ) u (t − t0 ) m t>5s
ωn
 
0.3902
= 0.3519 cos 3 (t − 5) + sin 3 (t − 5) u (t − 5) m t>5s [c]
3
The second approach makes use of superposition. The total force is treated as the su-
perposition of two forces as shown in Figure 6.21 and described by

F(t) F1(t) F2(t)

F0 F0
t0
= +
t0 t t0 t t
– F0

FIGURE 6.21
Superposition of forces.

F (t) = F0 u (t) − F0 u (t − t0 ) = 7.2u (t) − 7.2u (t − 5) [d]

The response is obtained as the superposition of the responses to the two excitations
and we have
F0 F0
x (t) = (1 − cos ωn t) u (t) − (1 − cos ωn (t − t0 )) u (t − t0 )
k k

= 0.2 (1 − cos 3t) u (t) − 0.2 (1 − cos 3 (t − 5)) u (t − 5) [e]


We can show that the response expressions in Equations [c] and [e] are identical. The
response is plotted in Figure 6.22.
While in this example the difference in effort between the two ways of calculating the
response is not large, in general the superposition method is preferable, as it represents a
more general solution.

Example 6.13—Coulomb Friction


An interesting example of the step response is vibrating systems that are subjected to
Coulomb-type frictional forces, also known as dry friction. As discussed in Chapter 4, friction
forces act in the opposite direction of the velocity of the contact point and they dissipate
energy. Recall that the other forces that dissipate energy include viscous damping forces,
aerodynamic or other drag, and internal forces due to hysteresis.
A system subjected to frictional forces is shown in Figure 6.23a, which depicts a block
of mass m attached to a wall via a spring of constant k. The static and kinetic coefficients
268 Applied Dynamics

0.4

x(t) 0.2

−0.2

−0.4
0 1 2 3 4 5 6 7 8 9 10
t

FIGURE 6.22
Response of an undamped system to two step inputs.

a) b) c)
 
mg mg
x
k kx
F kx F F
m

k N k N
(s , k )
N N

FIGURE 6.23
a) System subjected to frictional forces, b, c) free-body diagrams.

of friction between the block and the surface it slides on are µs and µk . The friction force
always opposes velocity (or impending velocity). When the block is moving, the amplitude
of the friction force is the kinetic friction coefficient multiplied by the normal force.
The free-body diagrams for the moving block are shown in Figures 6.23b–c. Two free-
body diagrams are needed because the direction of the friction force is different depending on
the direction of the velocity (or impending velocity). There are three possible configurations:
1. The block is not moving. The velocity is ẋ = 0, even though x may not be zero and
the spring is deflected. The maximum value of the friction force is Ff ≤ µs N = µs mg
and its direction depends on the combined magnitude of F − kx. When F − kx > 0,
the friction force is to the left, and when F − kx < 0 the friction force is to the right.
The magnitude of the friction force is found from static equilibrium with the result
|Ff | = |F − kx|.
2. The block is moving with velocity ẋ > 0 (to the right), as shown in Figure 6.23b. Here,
the friction force is Ff = −µk N = −µk mg and it is directed to the left.
3. The block is moving with velocity ẋ < 0 (to the left), as shown in Figure 6.23c. Here,
the friction force is Ff = µk N = µk mg and it is directed to the right.
Response of Dynamical Systems 269

Using the signum function, whose plot is shown in Figure 6.24, (sign a = 1 when a > 0
and sign a = −1 when a < 0), the friction force for when the block is moving can be
expressed as Ff = −µk mg sign(ẋ). It follows that the equation of motion can be written as

sign(a)

0
a

–1

FIGURE 6.24
Signum function.

mẍ + kx = F − µk mg sign(ẋ) [a]


The equation of motion is nonlinear, as the signum function is discontinuous at ẋ = 0.
However, the motion can be split into two regimes, depending on the sign of the velocity ẋ.
In each regime, the equation of motion is linear, and it can be integrated in the range of its
validity. Consider the case when the external force F = 0, so the two equations of motion
can be written as
when ẋ > 0, ẍ + ωn2 x = −µk g [b]
when ẋ < 0, ẍ + ωn2 x = µk g [c]
A third relationship is needed for the instant when ẋ = 0:

when ẋ = 0, ωn2 |x| ≤ µs g or ωn2 |x| > µs g [d]

If at the instant when the velocity is zero and if ωn2 |x (0) | > µs g, that is, the spring force
is larger than the maximum value of the static friction force, the block will have subsequent
motion, otherwise it will not move.
Consider as initial conditions an initial displacement x0 and no initial velocity, and that
ωn2 x0 > µs g. This relationship indicates that at the onset of motion, the spring force is
greater than the maximum value of the friction force, so the block will move. The initial
motion of the block is in the negative x direction, as x0 is positive and the spring force acts
to restore the block to its equilibrium position. Therefore, ẋ < 0, Figure 6.23c is the correct
free-body diagram, and Equation [c] is the equation of motion, whose solution is
 
µk g µk g µk g
x (t) = x0 cos ωn t + 2 (1 − cos ωn t) = + x0 − 2 cos ωn t [e]
ωn ωn2 ωn

The velocity becomes zero after a half cycle, t = T /2 = π/ωn . The displacement at this
point is  
π 2µk g
x = −x0 + 2 [f ]
ωn ωn
270 Applied Dynamics

and x is negative at this instant.


In the next phase (next half cycle) of the motion, the position is negative, so that the
spring will move the mass in the positive x direction. Hence ẋ > 0 and Equation [b] becomes
the equation of motion, with initial condition x( ωπn ) = −x0 + 2µ kg
2 . As long as the spring
ωn
force is larger than µs mg, the block will move. It follows that the response has the form
     
π µk g µk g π π 2π
x t− = − 2 + x0 − 3 2 cos ωn t − <t≤ [g]
ωn ωn ωn ωn ωn ωn

At the end of the second half cycle, t = 2π/ωn , the velocity is again zero, at which point
the displacement has the value
 
2π 4µk g
x = x0 − 2 [h]
ωn ωn

It follows that at each half cycle the amplitude of vibration is reduced by 2µ kg


2 . Hence,
ωn
the response subject to Coulomb friction is in the form of a decaying curve, with the lines
connecting the peak amplitudes in the form of two straight lines of slope ± 2µ kg
πωn . A typical
response curve is shown in Figure 6.25. The motion ends when the spring force is less
than the static friction force at the end of a half cycle. The straight lines mentioned above
resemble decay envelopes but, in reality, they are not. At the peak amplitudes, where the
response curve contacts the straight lines, the contacting curves are not tangent to each
other.

x(t)

x0 2k g
x0 – n t

s g
n2

/n 2/n 3/n t

– x0

FIGURE 6.25
Response to Coulomb friction.

The problem with Coulomb friction is that the solution has to be obtained for every half
cycle separately, making the analysis cumbersome. We often replace the Coulomb friction
model with an equivalent viscous damping model, using a method such as the one in Exam-
ple 6.10. While this is a gross simplification, considering the uncertainties in determining
friction forces and coefficients, decay rates and damping factors, the simplification becomes
justifiable, especially for low levels of friction.
Response of Dynamical Systems 271

6.14 Response to General Excitations—Convolution Integral


The response to an arbitrary excitation can be obtained by expressing the excitation as
the sum of impulsive forces, using the impulse response equations in Section 6.12, and
by invoking the principle of superposition. In the process, important results involving the
convolution integral are obtained.
Consider an excitation, such as the one in Figure 6.26. This general force can be expressed
as a summation of impulses. To this end, consider an arbitrary time t = τ and a small time
increment ∆τ . The shaded area denotes the impulse, which can be approximated as
F̂ (τ ) ≈ F (τ ) ∆τ (6.138)
as long as ∆τ is small. By making use of the Dirac delta function, the forcing between
τ < t < τ + ∆ can be expressed in terms of the impulsive force as
F (t, τ ) = F̂ (τ ) δ (t − τ ) ≈ F (τ ) ∆τ δ (t − τ ) (6.139)

!!""

!!IJ"

# IJ ǻIJ "

FIGURE 6.26
A general force and impulse generated at t = τ .

Recall that the response to an unit impulse applied at t = τ , δ (t − τ ), is g (t − τ ), as


obtained in Section 6.12. Denoting the component of the response due to the excitation
f (t, τ ) = F (t, τ )/m at t = τ by x (t, τ ), which has the units of force/mass, the response to
an impulse at t = τ can be written as
x (t, τ ) = f (τ ) ∆τ g (t − τ ) (6.140)
Next, consider the entire excitation. Equation (6.139) can be extended to the rest of the
excitation F (t), by representing it as a sum (superposition) of impulses in the form
X
F (t) = F (τ ) ∆τ δ (t − τ ) (6.141)

The response can be viewed as the superposition of the individual impulse responses
X X
x (t) = x(t, τ ) = f (τ ) ∆τ g (t − τ ) (6.142)

As the time period is made smaller, becoming infinitesimal, we replace the summation
by integration and ∆τ by the infinitesimal element dτ , which gives
Z t
x (t) = f (τ ) g (t − τ ) dτ (6.143)
0
272 Applied Dynamics

The above equation is known as the convolution integral. By a proper change of variables,
we can show that
Z t Z t
x (t) = f (τ ) g (t − τ ) dτ = f (t − τ ) g (τ ) dτ (6.144)
0 0

We base the selection of the form of the convolution integral to use on the ease with
which the convolution integral is evaluated. For example, if the excitation is of constant
magnitude, the form with f (t − τ ) is preferable.
Recalling the discussion in Section 6.12, the response obtained from the convolution
integral is the response of the system to zero initial conditions. To find the total response,
it is necessary to superpose the convolution integral with the response to initial conditions
only (the free response), with the result
 
−ζωn t v0 + ζωn
x (t) = e x0 cos ωd t + sin ωd t
ωd

Z t
1
+ f (t − τ ) e−ζωn τ sin ωd τ dτ (6.145)
ωd 0

For an undamped system, the response has the form


Z t
v0 1
x (t) = x0 cos ωn t + sin ωn t + f (t − τ ) sin ωn τ dτ (6.146)
ωn ωn 0
We can also use the Laplace transform approach to obtain the general solution. The
equation of motion and initial conditions are

ẍ (t) + 2ζωn ẋ (t) + ωn2 x (t) = f (t) x (0) = x0 , ẋ (0) = v0 (6.147)

Taking the Laplace transform gives

s2 + 2ζωn s + ωn2 X (s) = F (s) + (s + 2ζωn ) x0 + v0



(6.148)

where X (s) and F (s) are the Laplace transforms of x (t) and f (t), respectively. Recall from
Section 6.12 that the Laplace transform of the impulse response g (t) is
1
G (s) = (6.149)
s2 + 2ζωn s + ωn2
Solving for X (s), we obtain

X (s) = G (s) ((s + 2ζωn )x0 + v0 + F (s))

s + 2ζωn 1
= x0 + v0 2 + F (s) G (s) (6.150)
s2 + 2ζωn s + ωn2 s + 2ζωn s + ωn2
The inverse Laplace transform of the above equation gives Equation (6.145) for damped
systems and Equation (6.146) for undamped systems. Note that the inverse Laplace trans-
form of F (s) G (s) is
Z t Z t
−1
L (F (s) G (s)) = f (t − τ ) g (τ ) dτ = g (t − τ ) f (τ ) dτ (6.151)
0 0

When evaluating the convolution integral, we can


Response of Dynamical Systems 273
Rt
• Perform a direct integration of 0
g (t − τ ) f (τ ) dτ , or
• Take the Laplace transform of the convolution integral, which is F (s) G (s), express this
term in a simpler form, and then obtain the inverse Laplace transform.
Let us compare the convolution integral approach with the homogeneous plus particular
solution approach. When using the homogeneous plus particular solution, we first calculate
the homogeneous solution. This solution is in terms of unknown coefficients. Then, the
particular solution is obtained. There are no unknown coefficients in the particular solution.
The next step is to add the two solutions. We then invoke the initial conditions and solve
for the unknown coefficients.
When using the convolution integral approach, we do not solve for unknown coefficients.
Rather, we directly calculate the response from the initial conditions and by evaluating an
integral. Convolution is a more direct approach even though calculating the convolution
integral may be more difficult than finding the particular solution xP . On the other hand,
the convolution integral lends itself to easier computer implementation.

Example 6.14
Given an undamped mass-spring system with mass m and stiffness k, obtain the response
to the excitation F (t) = F0 sin (ωt) u (t), with ω 6= ωn , subject to the initial conditions
x (0) = x0 , ẋ (0) = v0 .
The equation of motion is
F0
ẍ (t) + ωn2 x (t) = sin ωt [a]
m
and, from Equation (6.146), the response is
Z t
v0 F0
x (t) = x0 cos ωn t + sin ωn t + sin ω (t − τ ) sin ωn τ dτ [b]
ωn mωn 0

We identify two methods for solving for the convolution integral. The first is to perform
the integration by expanding the sin ω (t − τ ) term, carrying out the algebra, and evaluating
the integral. The second is to make use of the the convolution theorem and Laplace transform
ω
tables. Recalling that L (sin ωt) = s2 +ω 2 , the Laplace transform of convolution integral

becomes  Z t 
1 ω 1
L sin ω (t − τ ) sin ωn τ dτ = 2 × 2 [c]
ωn 0 s + ω2 s + ωn2
The partial fraction expansion approach can be used to evaluate the inverse Laplace
transform. First, express the Laplace transform of the convolution integral as
ω 1 As + B Cs + D
× 2 = 2 + 2 [d]
s2 + ω 2 s + ωn2 s + ωn2 s + ω2
Equating the denominators of the right side of the above equation gives
 
ω 1 (As + B) s2 + ω 2 + (Cs + D) s2 + ωn2
× 2 = [e]
s2 + ω 2 s + ωn2 (s2 + ω 2 ) (s2 + ωn2 )
The coefficients A, B, C, and D are found by equating like powers in the numerator,
which gives
s0 : Bω 2 + Dωn2 = ω [f ]
s1 : Aω 2 + Cωn2 = 0 [g]
274 Applied Dynamics

s2 : B+D = 0 [h]
s3 : A+C = 0 [i]
From Equations [h] and [i] D = −B and C = −A. Substituting these values into
Equation [g] results in
A ω 2 − ωn2 = 0

[j]
As the case of resonance (ω = ωn ), for which the solution is different, is not being
considered here, Equation [j] leads to the conclusion that A = C = 0. Introducing Equation
[h] into Equation [f] and solving for B gives
ω ω
B = = − 2 [k]
ω 2 − ωn2 ωn − ω 2
and the partial fraction expansion becomes
 
ω 1 ω 1 1
× 2 = 2 − 2 [l]
s2 + ω 2 s + ωn2 ωn − ω 2 s2 + ω 2 s + ωn2

The next step is to obtain the inverse Laplace transform of Equation [l]. Introducing the
dimensionless frequency ratio ω̄, defined as ω̄ = ωωn , the inverse transform becomes
   
ω 1 ω sin ωt sin ωn t
L−1 × = −
s2 + ω 2 s2 + ωn2 ωn2 − ω 2 ω ωn
1
= (sin ωt − ω̄ sin ωn t) [m]
ωn2 (1 − ω̄ 2 )
Introduction of the above equation into Equation [b], gives the response as
 
v0 F0 1
x (t) = x0 cos ωn t + sin ωn t + (sin ωt − ω̄ sin ωn t) u (t) [n]
ωn k (1 − ω̄ 2 )
The general response to systems subjected to harmonic excitation is analyzed later in
this chapter. The approach in this example should be compared with those approaches.

6.15 Time-Domain vs. Frequency-Domain Analysis


The response analysis in this chapter has so far been concerned with obtaining the response
as a function of time. To this end, we solved the differential equations describing the motion,
and we developed expressions for position, velocity, and acceleration as a function of time.
Also, we analyzed plots of the response. This type of analysis is known as time-domain
analysis.
Another and equally important analysis of response is by means of frequency-domain
analysis. The frequency content of a signal yields valuable information beyond what is
learned in time-domain analysis. Frequency-domain analysis can broadly be classified into
two categories:
• Given a signal or measured response, we obtain the frequency spectrum of the signal by
taking its Fourier transform. We examine the frequency content, that is, the contribution
of different frequencies to the signal, as well as amplitudes associated with the different
frequencies.
Response of Dynamical Systems 275

4
x (t)
1

2 x (t)
x (t), x (t) 2
2

0
1

4
0 1 2 3 4 5 6 7 8 9 10
t (sec)

FIGURE 6.27
Plot of x1 (t) = 1.2 sin 2t and x2 (t) = 1.2 sin 2t − 2 cos 10.4t.

• Given the equations of motion, we solve these equations in the frequency domain. This
type of analysis is especially useful when the excitation is harmonic or periodic.
Shown in Figure 6.27, the signal x1 (t) = 1.2 sin 2t is sinusoidal with a constant am-
plitude. This signal is characterized by its amplitude and its frequency. The frequency is
measured in terms of radians/s or cycles/s (Hertz), or in terms of the period of oscillation.
For the signal above, the frequency is ω = 2 rad/s or f = ω/2π = 2/2π = 1/π Hz.
Next, consider a slightly different signal,

x2 (t) = 1.2 sin 2t − 2 cos 10.4t (6.152)

which is the previous signal to which a higher frequency component has been added. This
signal is also plotted in Figure 6.27. In the presence of several more frequencies in the
description of a signal, it becomes more difficult to identify the frequencies that contribute
to the signal by just looking at the time plot.
The Fourier transform is a useful way to analyze the frequency characteristics of a signal.
The Fourier transform of a time signal x (t) is defined by2 X (ω) and has the form
Z ∞
X (ω) = x (t) eiωt dt (6.153)
−∞

The Fourier transform converts a time signal into the frequency domain. The Fourier
transform has a real and an imaginary component, X (ω) = XRe (ω) + iXIm (ω), where
the subscripts denote real and imaginary components, respectively. Noting that eiωt =
cos ωt + i sin ωt, XRe (ω) and XIm (ω) can be expressed as
Z ∞ Z ∞
XRe (ω) = x (t) cos ωt dt XIm (ω) = x (t) sin ωt dt (6.154)
−∞ −∞

When generating the frequency spectrum, we first measure the signal at a number of
discrete points in time, a process called sampling. The Fourier transform is then obtained by
means of the Discrete Fourier transform. This process has become computationally efficient
by the development of the Fast Fourier transform (FFT). The FFT is so widely used that
every data analysis package contains an FFT routine.
2 Some prefer to define the Fourier transform in terms of the frequency f = ω/2π, using e2πif t .
276 Applied Dynamics

1.5
Amplitude

0.5

0
0 10 20 30 40 50 60 70 80
Frequency (rad/s)

FIGURE 6.28
Fast Fourier transform of x2 (t) = 1.2 sin 2t − 2 cos 10.4t.

The real and imaginary parts of the Fourier transform can be combined to describe the
amplitude and phase angles of the frequency spectrum as
 
XIm (ω)
q
2 (ω) + X 2 (ω) −1
XA (ω) = XRe Im φ (ω) = tan (6.155)
XRe (ω)

The Fourier transform of the signal x2 (t) = 1.2 sin 2t − 2 cos 10.4t can be shown to be
X2 (ω) = 1.2δ (ω − 2) − 2δ (ω − 10.4). It consists of two peaks at the two frequencies 2 and
10.4 rad/s. Figure 6.28 shows the frequency spectrum of this signal, obtained computation-
ally by the Fast Fourier transform. The plot consists of two peaks, supplemented with some
noise.
In frequency response plots, the frequency axis usually is in terms of Hertz (Hz). Also,
a majority of frequency plots are logarithmic (log-log or log-linear), as many times there is
a large frequency range of interest as well as a large amplitude range.

Example 6.15
Speaker manufacturers (also microphone manufacturers) usually provide data sheets asso-
ciated with the performance of their products, containing plots of the frequency response
and phase angles. Manufacturers test the performance of the speakers by inputting to them
audio signals at different frequencies (but with the same amplitudes). The outputs of the
speakers or microphones are measured; their Fourier transforms are calculated and com-
pared with the input, which is a straight line. The frequency range usually is from 20 Hz
(low base) to 20,000 Hz (20 kHz, very high treble).
The decibel (dB) is a commonly used logarithmic unit that measures signal strength
relative to a reference amplitude. For sound levels, a 0 dB corresponds to sound pressure of
0.0002 microbar. This sound level represents the threshold that humans can hear. Human
perception of the intensity of sound is nearly proportional to the logarithm of the sound
intensity; thus, the decibel correlates well with human hearing. In the United States, the
Environmental Protection Agency has identified 70 dB to be a disruptive level. Someone
frequently exposed to 85 dB sound is likely to suffer hearing loss.
A good speaker should have the same response at every frequency that is excited, so that
its frequency response curve should resemble a horizontal line. Also, the phase difference
between the input and output signals should be zero. Depending on the quality of the
Response of Dynamical Systems 277

!"

"

34
í

í

í
!"! !"# !"$ !"%
&'()*(+,-./012

FIGURE 6.29
Frequency response curve for a speaker.

speaker, the frequency response curve deviates from a horizontal line. The amplitude of
the deviation and the frequencies at which the deviation takes place are indications of the
quality of the speaker. Figures 6.29 and 6.30 show normalized frequency response curves
(frequency response of speaker divided by a reference signal, so the desired dB value is 0).
The frequency response curve in Figure 6.30 resembles a straight line, which is indicative
of a higher quality speaker.

!"

"
34

í

í

í
!"! !"# !"$ !"%
&'()*(+,-./012

FIGURE 6.30
Frequency response curve for a better speaker.

Be careful in using frequency response curves alone to determine speaker (or microphone)
quality, as the frequency response will be different at different amplitude levels. Quality
speakers will have performance as in Figure 6.30 for all sound amplitudes. Also, some
manufacturers provide frequency response curves for individual components of speakers,
such as woofers and tweeters.
278 Applied Dynamics

6.16 Response to Harmonic Excitation


The previous section examined the frequency content of a signal. This section analyzes
the response of systems acted upon by excitations that repeat themselves periodically. A
periodic signal can be expressed in terms of a summation of sines and cosines, with all terms
having periods that are integer multiples of the period of the signal. A signal that can be
represented by a single sinusoidal is called harmonic.
This section considers response to forcing that is harmonic. Response to general periodic
excitation is studied by taking the Fourier series expansion of the excitation, obtaining the
response to each term of the excitation, and using the principle of superposition to combine
the individual excitations.
Response to harmonic excitation is studied separately from the general system response
for the following reasons:
• Harmonic or periodic excitation is very commonly encountered, including rotating ma-
chinery, cars with unbalanced tires, travel on roads that are not smooth, and vortex
formation due to air flow over a body.
• It is of interest to analyze motion amplitudes for different values of the excitation fre-
quency. The need to avoid resonance is paramount.
• When analyzing response to harmonic excitation, we are not interested in initial con-
ditions or the amplitude of motion as a function of time. Rather, our interest is in
motion amplitudes and phase angles and how these quantities change as a function of
the excitation frequency.
For systems subjected to harmonic or periodic excitation, the primary interest is in the
motion characteristics after all transient effects, such as those due to initial conditions as
well as those due to short-term loads, have died out. This remaining motion is called the
steady-state response. Recalling the discussion of homogeneous and particular solutions for
differential equations, for harmonically excited systems the steady-state solution is recog-
nized as the particular solution.

6.16.1 General Formulation for Second-Order Systems


Consider the linear or linearized single-degree-of-freedom model represented by a mass-
spring-damper and depicted in Figure 6.5 and let the excitation be described by

F (t) = F0 cos ωt or F (t) = F0 sin ωt (6.156)

where F0 is the amplitude of the force and ω is the frequency of harmonic excitation.
Whether the excitation is expressed as a sine or cosine is immaterial, as the two functions
are separated by a phase angle of π/2 and we are not interested in the motion amplitudes
as a specific time instant.
It is convenient to use complex variable notation to describe harmonic excitation. Re-
calling that eiωt = cos ωt + i sin ωt, the excitation can be expressed in the general form
F (t) = F0 eiωt . This way, we can solve for the response for a cosine or sine excitation using
a single formulation. The equation of motion becomes

mẍ (t) + cẋ (t) + kx (t) = F0 eiωt (6.157)


Response of Dynamical Systems 279

Divide the equation of motion by the mass m and note that


F0 F0 2
= ω (6.158)
m k n
F0
Introducing the term A = k , where the dimension of A is displacement, the equation
of motion can be rewritten as
ẍ (t) + 2ζωn ẋ (t) + ωn x (t) = Aωn2 eiωt (6.159)
The steady-state response has the form of the particular solution and can be written as
xss (t) = Xss eiωt (6.160)
where Xss is the amplitude of the steady-state response. Introducing the above solution to
the equation of motion and collecting like terms gives
−ω 2 + 2iζωn ω + ωn2 Xss eiωt = Aωn2 eiωt

(6.161)
Dividing both sides by ωn2 eiωt and recalling the dimensionless frequency ratio ω̄ = ωωn ,
which describes the ratio of the excitation frequency to the natural frequency, we obtain
the steady-state response amplitude Xss as
1
Xss = A (6.162)
1 − ω̄ 2 + 2iζ ω̄
The steady-state amplitude is characterized by the excitation amplitude multiplied by
a nondimensional ratio. This ratio is dependent on the ratios of the system and excitation
frequencies. It is defined as
1
G (iω̄) = 2
(6.163)
1 − ω̄ + 2iζ ω̄
where G (iω̄) is called the frequency response. Recalling√from Section 6.6 that a complex
number a + ib can be represented as M eiψ , where M = a2 + b2 and ψ = tan−1 (b/a), we
can express the frequency response as
G (iω̄) = |G (iω̄)| e−iψ (6.164)
in which
1 2ζ ω̄
|G (iω̄)| = p ψ = tan−1 (6.165)
(1 − ω̄ 2 )2 + (2ζ ω̄)2 1 − ω̄ 2

The quantity |G (iω̄)| is referred to as the magnification factor and ψ is the phase angle,
not to be confused with the phase angles defined earlier in this chapter. The steady-state
response can be written as
xss (t) = A|G (iω̄) |ei(ωt−ψ) (6.166)
For a cosine excitation F (t) = F0 cos ωt, we use the real part of the above equation, and
for a sine excitation, we use the imaginary part. Hence,
For a cosine excitation: xss (t) = A |G (iω̄)| cos (ωt − ψ)

For a sine excitation: xss (t) = A |G (iω̄)| sin (ωt − ψ) (6.167)


The value of the magnification factor is affected by the ratio of the excitation frequency
to the natural frequency, as well as by the amount of damping. Figure 6.31 plots |G (iω̄) |
for various values of the damping factor. The magnification factor becomes smaller with
increasing amounts of damping. The magnification factor becomes larger as the frequency
ratio ω̄ approaches unity. The condition of ω̄ = 1 is known as resonance, and it manifests
itself in high amplitude vibrations, as will be discussed in the next section.
280 Applied Dynamics

!
"#$ ȟ!()'#&
"
ȟ!()'#%
*$+%Ȧ,* %#$
ȟ!()'#"
%
ȟ!()'#$
&#$
ȟ!()&
&
ȟ!()%
'#$
'
' '#$ & &#$ % %#$ "
Ȧ "!ȦȦ#

FIGURE 6.31
The magnification factor |G (iω̄)|.

6.16.2 Quality Factor


The maximum value of the magnification factor is obtained near resonance. To calculate
the maximum value of the magnification factor, it is necessary to differentiate |G (iω̄)| with
respect to ω̄. After some algebra, we can show that the peak value of the magnification
factor occurs at
p
ω̄ = 1 − ζ2 (6.168)

The maximum value of the magnification factor, known as the quality factor and denoted
by Q, can be shown to have the value
1
Q = |G (iω̄)|max = p (6.169)
2ζ 1 − ζ2

For small amounts of damping, such as ζ < 0.1, the peak value of |G (iω̄)| occurs almost at
resonance, and the quality factor can be approximated as
1
Q ≈ (6.170)

The name “quality factor” may be misleading to a person studying vibrations of struc-
tures and vehicles as in such cases we wish to stay as much away from resonance as possible.
On the other hand, engineers working in circuitry, such as the tuning circuit in a radio, are
interested in having as large amplitudes as possible. As discussed in the previous section, it
is preferable to design circuitry and components so that there is a faithful reproduction of
the input signal. In an electrical circuit, the resistor serves the same function as a damper
in a mass-spring system. It dissipates energy, which is undesirable.
The quality factor can also be used to measure damping in a system. Measuring the
steady-state response Xss at different frequencies, we can plot |G (iω̄)| (as long as we know
A = F0 /k), and compare with Figure 6.31. The measurement of the peak gives Q, from
which we calculate the damping factor. The problem with this approach is that, when taking
experimental measurements, the value of A may not be known accurately.
Response of Dynamical Systems 281

To alleviate this problem, the bandwidth method, also known as the half-power method,
has been proposed for lightly damped systems. Here, one first plots Xss (ω̄) from experi-
mental measurements. The maximum value Xssmax is
A
Xssmax = AQ = (6.171)


When the damping factor ζ is small, the two values of ω̄ when Xss (ω̄) = Xssmax / 2
can be shown to be

ω̄1 ≈ 1 − ζ − ζ 2 ω̄2 ≈ 1 + ζ − ζ 2 (6.172)

These values are obtained by noting that Xss (ω̄) = A |G (iω̄)| and solving for the two
frequencies ω̄1 and ω̄2 from
Xssmax A 1
Xss (ω̄) = √ = √ = A |G (iω̄)| (6.173)
2 2ζ 2
which eliminates the need to know the value of A. Addition of the two values for ω̄ in
Equation (6.171) gives the solution for damping as
1
ζ = (ω̄2 − ω̄1 ) (6.174)
2

6.16.3 Phase Angle

ȟ"+,#*! ȟ"+,#*& ȟ"+,#*( ȟ"+,#*%


!"#$

!%#$
ȟ"+,#
!&#$

ȥ )#$ ȟ"+,! ȟ"+,&

'#$

(#$

#$
# #*% ! !*% & &*% (
Ȧ !"ȦȦ#

FIGURE 6.32
Phase angle for harmonic excitation.

The phase angle ψ is plotted in Figure 6.32 for varying values of the damping factor. All
the phase angle curves reach the value of π2 when ω̄ = 1. For frequencies beyond resonance,
ω̄ > 1, the phase angle is greater than π/2 and this changes the sign of the response. When
the excitation frequency is less than the natural frequency, the phase angle ψ < π2 and
the response and excitation are in the same direction. This is known as being in phase.
By contrast, when ω > ωn , the phase angle is greater than 90◦ and the response is in the
opposite direction of the excitation. This is known as being out of phase.
For undamped systems, the phase angle is zero when ω̄ < 1 and it is 180◦ after resonance.
282 Applied Dynamics

Example 6.16
An undamped mass-spring system has a weight of 4 lb. It is subjected to a harmonic load
of amplitude F0 = 6 lb and period of T = 2 seconds. Design the spring constant k such that
the steady-state amplitude is Xss = 11 inches.
The steady-state amplitude is
F0 1
Xss = A |G (iω̄)| = [a]
k |1 − ω̄ 2 |

where the driving frequency is


2π 2π
ω = = = π rad/sec [b]
T 2
From Equation [a], we have two possibilities for the magnification factor:
1 1
|G (iω̄)| = or |G (iω̄)| = [c]
1 − ω̄ 2 ω̄ 2 − 1

Let us use the first possibility, so that Xss = Fk0 1−ω̄


1
2 . Multiplying and dividing this equation
2 2
by ωn and noting that k/m = ωn , after some algebra we arrive at

F0 6 4
k = + mω 2 = + π 2 = 7.772 lb/ft [d]
Xss 11/12 32.17

The natural frequency and dimensionless frequency ratios become


r s
k 7.772 ω π
ωn = = = 7.906 rad/sec ω̄ = = = 0.3974 [e]
m 4/32.17 ωn 7.906

F0 1
Let us now consider the second possibility, so that Xss = k ω̄ 2 −1 . It follows that the
expression for the spring constant becomes
F0 6 4
k = − + mω 2 = − + π 2 = −5.318 lb/ft [f ]
Xss 11/12 32.17

Because the right side is less than zero, no physically realizable solution is possible. Hence,
this problem has only one solution, k = 7.772 lb/ft.

6.17 Resonance
Dynamical systems subjected to harmonic excitation are designed such that resonant fre-
quencies are avoided, or encountered as briefly as possible. A system subjected to periodic
excitation very close to that of its natural frequency can experience dangerous levels of
vibration, which can cause physical damage, discomfort to occupants, and reduction in pre-
cision. Even one or two cycles of such excitation may lead to dangerous amplitude levels.
The Tacoma Narrows Bridge, which collapsed in 1940 as a result of wind-induced lon-
gitudinal and torsional vibrations, is a classic example of poor design that did not properly
take into account aerodynamic effects. Flow separation at the trailing edge, as discussed in
Response of Dynamical Systems 283

Chapter 4, led to the formation of vortices which initiated periodic excitations. Vibration
amplitudes became so high that the bridge collapsed.3
In 1850, a bridge in France collapsed because soldiers marching on it in step created
a harmonic excitation that matched the natural frequency of the bridge. The Millennium
Bridge in London in 2000 and Brooklyn Bridge in New York City in 2003 experienced high
amplitude vibrations because of the large volume of pedestrian traffic. The latter was due
to a power outage in New York City.
For undamped systems, setting ζ = 0 in Equation (6.166) the steady-state response
becomes
A
xss (t) = eiωt = A |G (iω̄)| ei(ωt−ψ) (6.175)
1 − ω̄ 2
where the magnification factor and phase angle takes the form
(
1 0 when ω̄ < 1
|G (iω̄)| = ψ= (6.176)
|1 − ω̄ 2 | π when ω̄ > 1

At resonance, that is, when the frequency ratio ω̄ = 1, the steady-state amplitude is
infinity for an undamped system. In practice, vibration amplitudes never reach infinity for
a variety of reasons:
• The system may sustain damage due to the high vibration amplitudes and collapse.
• The assumptions leading to the linearized model approximation lose their validity when
vibration amplitudes become high (necessitating the use of a more accurate mathemat-
ical model).
• All man-made or natural systems have a damping or other energy dissipation character-
istic, which prevents vibration amplitudes from becoming infinity. However, even though
there is damping, the magnification factor in Figure 6.31 can still become quite large.

It is of interest to examine how the vibration amplitudes increase under resonance. To


this end, write the excitation as F (t) = F0 eiωn t , so that the equation of motion becomes

F0 iωn t
ẍ (t) + ωn2 x (t) = e = Aωn2 eiωn t (6.177)
m
where A = F0 /k. A particular solution in the form x (t) = Ceiωn t does not work in this
case, as the solution attempted is the same as the homogeneous solution. Instead, try a
solution in the form x (t) = Cteiωn t . Differentiating this expression twice gives

ẍ (t) = 2iωn Ceiωn t − ωn2 tCeiωn t (6.178)

Introducing the above expression into the equation of motion results in

2iCωn eiωn t − Cωn2 teiωn t + Cωn2 teiωn t = Aωn2 eiωn t (6.179)

which, upon division by eiωn t , reduces to

2iCωn = Aωn2 (6.180)


3 You can view video of the spectacular collapse of the Tacoma Narrows Bridge by going to a public

domain video site, such as youtube.com.


284 Applied Dynamics

with the conclusion that C = − iAω


2 . Hence, the response has the form
n

A A
x (t) = − iωn teiωn t = ωn t (sin ωn t − i cos ωn t) (6.181)
2 2
which represents a sinusoidal whose amplitude increases linearly. Consider, without loss
of generality, the case when the excitation is F (t) = F0 cos ωn t, that is, the real part of
F0 eiωn t . The response is the real part of x (t), which is

A F0
Re (x (t)) = ωn t sin ωn t = t sin ωn t (6.182)
2 2mωn
The excitation and response have a phase difference of π/2. Figure 6.33 plots the re-
F0
sponse. The rise envelopes are given by ± 2mω n
t = ± A2 ωn t. After each cycle, that is, after a
time of T = 2π/ωn , the amplitude rises by Aπ. Sometimes, even after one cycle of excitation,
amplitude levels reach undesirable or uncomfortable values.

!!"" # Ȧ"
$
#

$
"

FIGURE 6.33
Response to resonant excitation.

To obtain the response using the Laplace transform, consider, without loss of generality,
the excitation F (t) = Fm0 cos ωn t and zero initial conditions

F0
ẍ (t) + ωn2 x (t) = cos ωn t (6.183)
m
The Laplace transform of the above equation of motion is
s
s2 + ωn2 X (s) = Aωn2

(6.184)
s2 + ωn2

so that
s
X (s) = Aωn2 2 (6.185)
(s2 + ωn2 )

A theorem from Laplace transform theory states that if F(s) is the Laplace transform
of f (t), then the Laplace transform of tf (t) is L (tf (t)) = −F 0 (s) where the prime denotes
differentiation with respect to s. For f (t) = sin ωn t, the Laplace transform of tf (t) becomes
 
d ωn s
L (t sin ωn t) = − = 2ωn 2 (6.186)
ds s2 + ωn2 (s + ωn2 )
2
Response of Dynamical Systems 285

which leads to Equation (6.182) as the response.


Resonance is more dangerous for systems with lower frequencies. More energy is dissi-
pated by damping in a system with a higher natural frequency. Also, forces needed to pro-
duce a persistent harmonic excitation at higher frequencies are larger than forces needed at
lower frequencies. In multi-degrees-of-freedom systems, we usually are most worried about
resonance of the lowest natural frequency.
The natural frequency of a system subjected to harmonic excitation is often smaller
than the excitation frequency. For example, rotating machinery usually have natural fre-
quencies much lower than their operating speeds. When the machine is started from rest,
the rotational speed will go through the resonant frequency before it reaches its operating
value, subjecting the machine to resonance.
The way to deal with such a situation is to make sure that resonance is encountered for
as short an amount of time as possible and that the operating frequency is reached rapidly.
Similarly, bridges have lower natural frequencies than car engine speeds. Drivers stuck on
suspension bridges are advised not to turn their engines off and on.

6.18 Transmitted Force

'" ("
!!"" !!""

% %
'
#
$ $ !#$&%&$'&F‫ܪ‬
) )

FIGURE 6.34
a) System acted upon by a harmonic force, b) free-body diagram.

Consider a mass-spring-damper system connected to a fixed surface and subjected to a


harmonic force such as the one that we studied in Section 6.16. The body and its free-body
diagram are shown in Figure 6.34.
This system is representative of a machine with rotating components that is on a factory
floor and rests on a base, also known as a mount, that is modeled by a spring and a damper.
The interest is in examining the amplitude of the force transmitted to the base. From earlier
developments, when the force has the form F (t) = F0 eiωt , the response has the form

x (t) = AG (iω̄) eiωt (6.187)

where A = F0 /k. The free-body diagram indicates that the force transmitted to the base,
denoted by Ftr (t), is

Ftr (t) = kx (t) + cẋ (t) (6.188)


286 Applied Dynamics

Introducing Equation (6.187) into Equation (6.188), the transmitted force is calculated
as
 
2iωc
Ftr (t) = F0 1 + G (iω̄) eiωt = F0 (1 + 2iζ ω̄) G (iω̄) eiωt (6.189)
k
so that the ratio of the transmitted force to the applied force becomes
Ftr (t)
= (1 + 2iζ ω̄) G (iω̄) = H (iω̄) (6.190)
F (t)
where
1 + 2iζ ω̄
H (iω̄) = (6.191)
1 − ω̄ 2 + 2iζ ω̄
is the associated frequency response function. We can express H (iω̄) in terms of its magni-
tude and phase angle as

H (iω̄) = |H (iω̄)| eiφh (6.192)

where the magnitude and phase angle are

2ζ ω̄ 3
q
2
|H (iω̄)| = |G (iω̄)| 1 + (2ζ ω̄) φh = tan−1 (6.193)
1− ω̄ 2
+ (2ζ ω̄)2

"#$
ȟ#+,'#&
"
ȟ#+,'#%
(!)ȓȦ*( %#$
ȟ#+,'#"
%
ȟ#+,'#$
&#$

&
ȟ#+,% ȟ#+,&
'#$

'
' '#$ & &#$ % %#$ "
Ȧ "#ȦȦ$

FIGURE 6.35
Transmissibility.

The term |H (iω̄)| is known as the transmissibility, a measure of how much force is
transmitted to the base. Another reason for the name transmissibility is discussed in the
next section. A plot of |H (iω̄)| is given in Figure 6.35 for different values of the damping
factor. As expected, lower levels of damping lead to higher amounts of transmissibility.
Response of Dynamical Systems 287

This can be explained by noting that dampers absorb energy and more damping means
more absorbed energy and less transmitted energy. Also, as expected, the transmissibility
increases in the vicinity of resonance.
Note that when the damping is zero, the transmissibility and magnification factor be-
come the same quantities.
|H (iω̄)|ζ=0 = |G (iω̄)|ζ=0 (6.194)
An√ interesting feature of transmissibility is that all plots cross the |H (iω̄)| = 1 line at
ω̄ = 2. In addition, unlike in the plot for the magnification factor |G (iω̄)|, the transmissi-
bility√values decay faster for lower levels of damping than for excitation frequencies beyond
ω̄ = 2. The significance of this property is discussed in Chapter 15, within the context of
ride properties and damper design for vehicles.

6.19 Base Excitation


The previous two sections considered a mass-spring-damper system subjected to a force
whose amplitude varies harmonically. This section discusses the very commonly encountered
case of base excitation. A building subjected to seismic excitation, a vehicle moving over an
imperfect road, and any object connected to a base that shakes or vibrates are examples of
base excitation.

a) b) mg

z(t)
m m

c
k k
2 2
y(t) k(z–y) c( – )

FIGURE 6.36
a) Base excitation of mass-spring-damper, b) free-body diagram.

Consider the mass-spring-damper attached to a moving base in Figure 6.36a. The coor-
dinate y (t) denotes the motion of the base, which is in the form y (t) = y0 eiωt . The motion
of the base is treated as a prespecified or known quantity; hence, the base does not add a
degree of freedom.
The motion of the mass is denoted by z (t) and is measured from the undeformed position
of the spring. Hence, the deflection of the spring connecting the base to the mass is z (t) −
y (t). Using the free-body diagram in Figure 6.36b and summing forces in the vertical
direction gives
X
+↑ F = −k (z (t) − y (t)) − c (ż (t) − ẏ (t)) − mg = maz (6.195)

where az = z̈, so the equation of motion becomes


mz̈ (t) + cż (t) + kz (t) = −mg + cẏ (t) + ky (t) (6.196)
288 Applied Dynamics

As discussed in Chapter 5, the −mg term leads to the static equilibrium position of
zst = −mg/k that lowers the mass. Alternatively, we can measure z (t) from the static
equilibrium position, which eliminates the −mg term from the equation of motion. The
gravity term does not affect the response to the harmonic excitation.
Dividing the equation of motion by m, considering a steady-state solution in the form
that is similar to the excitation, zss (t) = Z (iω̄) eiωt , and introducing this solution to the
equation of motion gives
−ω 2 + 2iζωωn + ωn2 Zeiωt = 2iζωωn + ωn2 y0 eiωt
 
(6.197)
ω
Dividing both sides by ωn2 and using the frequency ratio ω̄ = ωn , we can solve for the
amplitude Z as
1 + 2iζ ω̄
Z (iω̄) = y0 = y0 G (iω̄) (1 + 2iζ ω̄) = y0 H (iω̄) (6.198)
1 − ω̄ 2 + 2iζ ω̄
where the amplitude ratio H (iω̄) is defined in the previous section. The absolute value of
the steady-state amplitude is related to the transmissibility by
|Z (iω̄)| = |y0 | |H (iω̄)| (6.199)
where the transmissibility |H (iω̄)| is defined in Equation (6.193). For base excitation, trans-
missibility is a measure of how much of the base motion is transmitted to the mass.

Example 6.17—Motion over Wavy Terrain


An interesting application of base excitation is the motion of a vehicle over an imperfect
road. The imperfection can be due to poor construction, wear, age, shifting of the ground
below the road surface, or buckling of the road surface. Consider a vehicle, modeled as a
mass-spring-damper, traveling with constant speed v over a road whose shape is approxi-
mated as a sinusoidal of period L, as shown in Figure 6.37.4

#"$#
! ) ("'#
*
" "
! !
#"'#

$
&

FIGURE 6.37
Single-degree-of-freedom model of a vehicle over wavy terrain.

The road surface can be modeled as


 
2πx
y (x) = Y sin [a]
L
4 In general, the amplitude of the unevenness is related to its period. Large amplitude imperfections

usually have long wavelengths and small amplitude imperfections have small periods.
Response of Dynamical Systems 289

where Y and L denote the amplitude and length of the wave. Because we assume the
vehicle is moving with constant speed (amplitudes of road imperfections, when small, do
not affect the horizontal speed too much), its location on the x axis is represented by x = vt.
Substituting this relationship into Equation [a], the harmonic motion of the base is obtained
as  
2πvt
y (t) = Y sin = Y sin ωt [b]
L
where the excitation frequency ω is related to the vehicle speed by
2πv
ω = [c]
L
The motion over wavy terrain problem is essentially a base excitation problem, with the
same equation of motion as Equation (6.196). The response has the form

z (t) = Y H (iω̄) eiωt [d]

The vehicle amplitude is Y |H (iω̄)|, where the transmissibility |H (iω̄)| is shown in Figure
6.35.
Let us analyze the motion amplitudes in terms of the vehicle speed. From Equation
[c], as the vehicle speed increases so does
p the excitation frequency. The critical velocity is
reached at resonance, where ω = ωn = k/m, which gives the value of the critical speed as
r
L L k
vcr = ωn = [e]
2π 2π m
As the
√ speed increases beyond resonance, the vibration amplitude drops and after the value
ω̄ > 2 is reached, the transmissibility becomes lower than 1. The drop rate is higher for
lower levels of damping.
We can observe these results when driving a vehicle. As a vehicle goes faster, the effects
of uneven terrain and potholes are felt less, as long as the vehicle maintains contact with the
road surface. It should be cautioned that when a vehicle goes faster over potholes or rough
terrain, the damping forces become larger. Such driving wears out the dampers more rapidly.
While the displacement amplitude is Y |H (iω̄)|, the velocity amplitude is ωY |H (iω̄)| and
ω increases with speed, as indicated by Equation [c].
When a vehicle travels on a wavy road, the forces transmitted to the tires, that is, the
normal forces acting on the tires (wheel loads), vary as a result of the undulations of the
road. This variation affects the lateral forces generated by the tires and may reduce lateral
stability of the vehicle. Chapter 14 analyzes lateral instability of vehicles.
As discussed in Chapter 15, a ballpark figure for the bounce frequency of a suspension
system is fn = ω2πn ≈ 1 Hz, so that the natural frequency is ωn ≈ 2π rad/sec. This leads to
an expression for the frequency ratio as
ω 1 2πv v
ω̄ = ≈ = [f ]
ωn 2π L L
and the critical speed becomes
vcr = L [g]
Note that the units of v and vcr in the above equations are the unit of L divided by
seconds. For example, if the length of the period of road imperfection is 10 m (about four
times the wheelbase of a passenger car), the critical speed becomes vcr = 10 m/s or 36 kph.
If the vehicle speed is 45 kph (12.5 m/s), the frequency ratio becomes ω̄ = 12.5/10 = 1.25.
290 Applied Dynamics

Recall from the transmissibility plot in Figure 6.35 that after ω/ωn > 2 the transmis-
sibility is less than 1, |H(iω̄)| < 1, the minimum desired speed is

vdesired > 2L [h]

so that for a road√imperfection wavelength of 10 m, the minimum speed for a transmissibility


of less than 1 is 2 × 10 ≈ 14 m/s.

6.20 Harmonic Excitation Due to Imbalances and Eccentricity


Rotating systems are designed to be symmetric about the rotation axis. Asymmetry about
the rotation axis, such as having eccentricity, results in imbalances and harmonic excita-
tion due to these imbalances. This section discusses harmonic excitation due to commonly
encountered imbalances.

6.20.1 Rotating Imbalances

me

e
O
x m – me

k c

FIGURE 6.38
Rotating imbalance.

Internal imbalances arise from rotating components that are not completely symmetric,
leading to centrifugal forces. A generic model of a system with an imbalance is shown in
Figure 6.38. The system has a total mass of m and consists of a primary mass, mounted on
a spring and a damper, and an imbalance. The imbalance can be modeled as a massless rod
of length e connected to a point mass me , with the rod and mass rotating with constant
angular velocity ω. The primary mass is m − me .
Separating the imbalance from the primary mass, the free-body diagrams are shown in
Figure 6.39. Horizontal forces between the two masses do not contribute to the dynamics
when the horizontal motion of the primary mass is restrained. Assuming that this is the
case, the acceleration of the imbalance in the vertical direction is

d2
ax = (x + e sin ωt) = ẍ − ω 2 e sin ωt (6.200)
dx2
Response of Dynamical Systems 291

a) b)
(m – me)g

me g Oh
x

Oh O
Ov

Ov kx

FIGURE 6.39
Free-body diagrams: a) imbalance, b) primary mass.

where x is measured from the undeformed position of the spring. Summing forces for the
imbalance gives
X
F = me ax =⇒ me ẍ − ω 2 e sin ωt = Ov − me g

+↑ (6.201)

The force balance for the primary mass is


X
+↑ F = (m − me ) ẍ =⇒ (m − me ) ẍ = −cẋ − kx − (m − me ) g − Ov (6.202)

Adding the above two equations eliminates Ov , and rearranging terms gives the equation
of motion as

mẍ + cẋ + kx = −mg + me eω 2 sin ωt (6.203)

where the harmonic excitation force is recognized as me eω 2 sin ωt. As noted previously, the
gravity term does not contribute to the harmonic motion. The force of gravity affects only
the static equilibrium position. Equation (6.203) is similar to Equation (6.159), with the
amplitude A as

me ω 2 me 2
A = e = eω̄ (6.204)
m ωn2 m

The steady-state motion amplitude can be shown to be


me 2
|Xss | = e ω̄ |G (iω̄)| (6.205)
m
with the phase angle the same as Equation (6.165). If the rotation frequency ω is close to
the natural frequency, or ω̄ ≈ 1, the primary mass attains high vibration amplitudes.

6.20.2 Whirling of Rotating Shafts


In machinery, such as engines and turbines, disks and gears are mounted on shafts. The
elastic deformation of the shaft causes the axis of rotation of the disk to move. If, in addition
to the elastic deformation, there are imperfections in the disk, the disk ends up rotating
about a point that moves and thus is not its center of mass. The rotation of a vibrating
shaft is known as whirling.
292 Applied Dynamics
#
!"#$ %&'()

"

FIGURE 6.40
Disk on an elastic shaft.

Figure 6.40 illustrates a disk that is mounted on a shaft. The side view of the disk is
shown in Figure 6.41a, where point O denotes the undeformed position of the shaft, C
denotes the point on the shaft to which the disk is attached (center of rotation), and G
denotes the center of mass of the disk. The distance between C and G is denoted by  and
it represents the eccentricity of the disk. The disk is rotating with angular velocity ω with
respect to the fixed axes XY . Figure 6.41b shows a close-up of the angles involved.

#1 21
Ȧ
!
3'/4
&
%
! İ ȦW
% ș İ $ $
ȦW ij
" " ș
( #
)"
ș +(,'(
'( ș
* #

0()*$+,-*)& !"#$%&'(&
/"#$% )*$+,-*)&
.+/'%'+(

FIGURE 6.41
a) Side view of disk, b) close-up and acting forces.

The position of the center of the disk due to the deformation of the shaft is denoted by
the deformation d and angle θ. The angular velocity of the shaft, also known as whirling
speed, is θ̇. The angular velocities of the shaft and the disk are not necessarily the same.
The condition when they are, ω = θ̇, is known as synchronous whirl.
Consider a moving coordinate system xy that is attached to the shaft and is obtained
by rotating the inertial XY coordinate system by θ. The deformation of the shaft at point
Response of Dynamical Systems 293

C is

rC = di (6.206)

It follows that the position of the center of mass of the disk can be expressed as

rG = rC + rG/C = di +  cos ωtI +  sin ωtJ (6.207)

where ω is the angular velocity of the disk. We assume that ω is constant, so ωt is the angle
between the X axis and line CG. This mixed description is convenient for the purpose of
calculating the velocity and acceleration of G. The velocity and acceleration of the center
of mass can be obtained by means of the transport theorem. Denoting the angular velocity
of the reference frame xyz by Ω = θ̇k, the velocity and acceleration of point C become
˙ + dθ̇j
vC = ṙCrel + Ω × rC = di (6.208)

   
aC = v̇Crel + Ω × vC = d¨ − dθ̇2 i + dθ̈ + 2d˙θ̇ j (6.209)

Because ω is assumed to be constant, the relative acceleration aG/C is simply

aG/C = −ω 2 rG/C (6.210)

After some algebra, the acceleration of G, aG = aC + aG/C , can be expressed in terms


of the xy coordinates as
   
aG = d¨ − dθ̇2 − ω 2 cos (ωt − θ) i + dθ̈ + 2d˙θ̇ − ω 2 sin (ωt − θ) j (6.211)

The force acting on the shaft is due to the stiffness and damping of the shaft and can
be written as
 
F = −krC − cvC = − kd + cd˙ i − cdθ̇j (6.212)

While this force, shown in Figure 6.41b, exerts a moment about the center of mass of the
disk, its effect is small, as the eccentricity is small. For the purpose of the analysis here, we
assume that the rotational motion of the disk is not affected by the movement of the shaft.
Using Newton’s Second Law, and with d and θ as the motion variables, the two equations
of motion are the components of F = maG in the x and y directions
 
m d¨ − dθ̇2 + cd˙ + kd = mω 2 cos (ωt − θ)

 
m dθ̈ + 2d˙θ̇ + cdθ̇ = mω 2 sin (ωt − θ) (6.213)

The equations of motion can be simplified if synchronous whirl is assumed and ω = θ̇ =


const. Further simplification is possible for steady-state behavior, where d is also constant,
so that d˙ = 0, d¨ = 0. For steady-state, the equations of motion reduce to
c
ωn2 − ω 2 d = ω 2 cos φ ωd = ω 2 sin φ

(6.214)
m
where φ = ωt − θ is the angle that the line joining C and G makes with the x axis (Figure
294 Applied Dynamics

6.41b). Dividing the second equation above by the first, and making use of the damping
factor ζ and frequency ratio ω̄ = ω/ωn , gives the steady-state value for the angle φ as
2ζ ω̄
tan φ = (6.215)
1 − ω̄ 2
Further manipulation of the above equations provides an expression for the steady-state
value of the deformation d as
ω̄ 2
d = q = ω̄ 2 |G (iω̄)| (6.216)
2 2
(1 − ω̄) + (2ζ ω̄)
The critical speed of the disk is related to the natural frequency of the shaft. Indeed,
resonance occurs when
r
k
ωc = ωn = (6.217)
m
When  = 0, that is, if there is no eccentricity of the shaft, there are no harmonic
terms in the equations of motion for the disk. Harmonic terms appear when calculating the
reactions at the supports of the shaft, as will be demonstrated in Chapter 11, whether there
is eccentricity or not.

6.21 Bibliography
Benaroya, H., and Nagurka, M., Mechanical Vibration, 3rd Edition, CRC Press, 2009.
Bottega, W.J., Engineering Vibrations, CRC Press, 2006.
Dimaragonas, A., Vibration for Engineers, 2nd Edition, Prentice-Hall, 1996.
Greenberg, M.D., Advanced Engineering Mathematics, 2nd Edition, Prentice-Hall, 1998.
Irvine, T., vibrationdata.com Newsletter, October 2001.
Meirovitch, L., Fundamentals of Vibrations, Waveland Press, 2010.
Milliken, W.L., and Milliken, D.F., Race Car Vehicle Dynamics, SAE Publications (R146),
1995.
Thomson, W.T. and Dahleh, M.D., Theory of Vibration with Applications, 5th Edition,
Prentice-Hall, 1993.

6.22 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 6.3—Homogeneous Plus Particular Solution Approach


6.1 (M) Obtain the solution of the differential equation 2ẋ + 0.9x = 2 + 0.3t − 0.2t2 , subject
to the initial condition x (0) = −1.2.
6.2 (M) Obtain the solution of the differential equation 2ẍ + 18x = 2 + 0.3t, subject to the
initial condition x (0) = −1.2, ẋ (0) = 0.2.
6.3 (M) Obtain the solution of the differential equation 2ẍ + 18x = 0.3 sin 2t, subject to the
initial condition x (0) = 2.2, ẋ (0) = 0.
Response of Dynamical Systems 295

Section 6.4—Laplace Transform Solution


6.4 (M) Obtain the Laplace transform of the function f (t) = t sin 3t by direct integration.
1
6.5 (M) Calculate the inverse Laplace transform of the function F (s) = (s+2)(s 2 +4) by

partial fractions.
1
6.6 (M) Calculate the inverse Laplace transform of the function F (s) = (s2 +9) 2 by partial

fractions.
6.7 (M) Obtain the solution to the differential equation in Problem 6.1 using the Laplace
transform method.
6.8 (M) Obtain the solution to the differential equation in Problem 6.2 using the Laplace
transform method.
6.9 (M) Obtain the solution of the differential equation 2ẍ + 18x = 0.3 sin 2t, subject to the
initial condition x (0) = 2.2, ẋ (0) = 0 using the Laplace transform method.

Section 6.5—First-Order Systems


6.10 (M) A damper of constant c = 20 N·s/m is used to reduce the closing speed of the
5 kg kitchen drawer in Figure 6.42a. If the drawer is slammed with a speed of 0.2 m/sec,
calculate the speed of the drawer after 0.6 seconds. Also, find the time it takes for the speed
of the drawer to reach 0.1/sec. What is the time constant? Hint: Use the velocity as the
motion variable.
a) b)

x L/2 g
L/2
c O 
m
G
k B

k c

FIGURE 6.42
Figures for a) Problem 6.10, b) Problem 6.12.

6.11 (M) Consider the supermarket cart in Examples 8.2 and 8.14, and the equations of
motion. Now, set the acceleration of point A to zero, u̇1 = v̇A = 0, so that the system has
one degree of freedom. Calculate the time constant associated with the equation of motion.
Comment on the stability of the supermarket cart.

Section 6.7—Second-Order Systems


6.12 (M) The rod in Figure 6.42b is supported by two springs and a damper. Obtain the
equation of motion, linearize about θ = 0, and identify the equivalent mass, spring, and
damping coefficients.
6.13 (M) The mass m in Figure 6.43a is suspended from a string of constant tension T .
Using x as the motion variable, and assuming x is small relative to the string length L,
obtain the equation of motion and identify the equivalent mass and stiffness.
296 Applied Dynamics

a) b)
g g L/3
B
2L/3 O
x
T T A
m 2k
L–a a
k
L

FIGURE 6.43
Figures for a) Problem 6.13, b) Problem 6.14.

Section 6.8—Free Response of Undamped Second-Order Systems


6.14 (E) A beam of mass m and length L is pivoted at a distance L/3 from the right, as
shown in Figure 6.43b. Two springs of constants k and 2k act on the ends. Calculate the
natural frequency.
6.15 (E) You are designing the suspension of a car, modeled as a mass-spring system. The
desired frequency is 1.12 Hz. If the vehicle weighs 1300 lb, what should the spring constant
be?
6.16 (M) Calculate the natural frequency of the compound pendulum in Figure 6.44a. The
pendulum rotates freely about point O and has the following parameters: m1 = 0.5 slugs,
m2 = 0.2 slugs, m3 = 0.8 slugs, R = 2 ft, L = 4 ft, r = 1.5 ft.

a) b)
1.5 g O
O m1 g

m2 L0
2m, 2L
4 G
mL2
m3 G m, IG = 24
2

FIGURE 6.44
Figures for a) Problem 6.16, b) Problem 6.17.

6.17 (M) A body of mass m = 5 kg and mass moment of inertia IG = mL2 /24 (L = 75 cm)
is suspended from a pivot by a rod of mass 2m and length 2L so that the distance between
the center of mass and the pivot is L0 , as shown in Figure 6.44b. Find the natural frequency
of the system when a) L0 = 50 cm, b) L0 = 25 cm.
6.18 (M) The response of an undamped mass-spring system is measured and the peaks are
observed to be at 1.3, 5.7, and 9.9 seconds. Given that the spring constant is k = 5 N/mm,
estimate the mass of this system.
6.19 (C) Consider a simple pendulum with m = 1 kg, L = 2 m. Write a computer program
to integrate the equation of motion θ̈ (t) + Lg sin (θ (t)) = 0 for the set of initial conditions
Response of Dynamical Systems 297

at t = 0, 1) θ (0) = 40◦ , θ̇ (0) = 0, 2) θ (0) = 20◦ , θ̇ (0) = 0. Plot θ (t) vs. time and the
tension in the cable vs. time for both sets of initial conditions. Next, linearize the equation
of motion about θ = 0◦ and integrate the resulting equation with the same initial conditions
as above. Generate and compare the response for the linear and nonlinear model.
6.20 (D) A mass m1 is resting on a spring k. A second mass m2 is dropped onto the mass m1
from a height h, as shown in Figure 6.45a. Assuming the impact is plastic (the two masses
move together), obtain the response of the combined masses. The variable x is measured
from the static equilibrium position of m1 .

"# $#

& "$
*+

% (
)
'
"# ,-

! ! ș
! !

FIGURE 6.45
Figures for a) Problem 6.20, b) Problem 6.21.

6.21 (M) Torsional vibration of shafts is commonly modeled by considering a disk of mass
moment of inertia ID attached to a light shaft of length L, area polar moment of inertia J,
and torsion constant G, as shown in Figure 6.45b. A torque T acts on the disk. Denoting
the rotation of the disk by θ, obtain the equation of rotational motion and identify the
equivalent mass and stiffness.

Sections 6.9 and 6.10—Free Response of Damped Systems, Underdamped Sys-


tems

"
!

&

#% #% #$

'

FIGURE 6.46
Figure for Problem 6.22.

6.22 (M) Obtain the equation of motion for the mass-spring-damper system in Figure 6.46
and find the value of c for critical damping in terms of m, k1 and k2 .
298 Applied Dynamics

6.23 (E) Consider Problem 6.15. Now, you need to add a damper to the system so that
the damping factor will be 0.22. Calculate the damping constant of the damper. Then, all
of a sudden, the vehicle weight is increased to 1500 lb. What happens to the values of the
natural frequency and damping factor in view of this added mass?
6.24 (M) A mass-spring-damper system vibrates freely and has the following properties:
m = 3 kg, k = 6 N/m, c = 1.8 N·s/m. The initial condition is x (0) = 1.2, ẋ (0) = 0.
Calculate the damped natural frequency, amplitude Ad , and phase angle φd and calculate
the response amplitude at t = 2.4 seconds. Verify your answer by using Equation (6.88).
6.25 (M) A freely vibrating mass-spring-damper system has the following properties: m = 2
kg, k = 6 N/m, c = 1.6 N·s/m. The initial condition is x (0) = 0, ẋ (0) = 0.8. Calculate the
damped natural frequency, amplitude Ad and phase angle φd and calculate the response
amplitude at t = 3.5 seconds. Verify your answer by using Equation (6.88).

Section 6.11—Damping Estimation by Logarithmic Decrement


6.26 (E) The amplitude ratio of two consecutive peaks is measured as 0.8, with the time
interval between the two peaks as 1.4 seconds. What is the amplitude ratio between the
first and fourth peaks? If this mass-spring-damper system weighs 10 kg, what are the spring
and damping constants?
6.27 (M) You are designing a damper for a system of weight 2 lb and spring constant 5
lb/in. The peak amplitude of free motion should be between 50% and 65% of the original
amplitude after four cycles. Calculate the range of the damping constant of the damper
that will allow you to maintain this range.
6.28 (M) The peak amplitudes of the cycles of a second-order system are measured as 1,
0.8, 0.6, 0.4, 0.2. Assuming that the damping model is viscous, calculate an estimate for the
damping factor using the logarithmic decrement method.
6.29 (M) You need to to design an underdamped mass-spring-damper system for a body
weighing 200 kg. The damped period of oscillation is to be Td = 2 seconds and the amplitude
to be reduced by a factor of 2.5 in one cycle. That is x(t0 + Td ) = x (t0 ) /2.5, where t0 is
the time when the first peak is observed. Use the logarithmic decrement method to design
the stiffness k and damping constant c.

Section 6.12—Impulse Response


6.30 (E) A mass-spring system with m = 3 kg and k = 243 N/m is at rest when an
impulsive force is applied to it for 0.002 seconds. As a result of the impulsive force, the
velocity becomes 0.3 m/s. Find the amplitude of the applied force assuming the impulsive
force vs. time plot is a) rectangular, b) triangular.
6.31 (M) A wire pendulum consisting of weight W = 2 lb and taut wire of length L = 3 ft
is at rest when it is struck by a horizontal impulsive force of 350 lb for a period of 0.0007
seconds. Calculate the amplitude of the pendulum 2.3 seconds after the impulsive force is
applied.
6.32 (M) A mass-spring system weighing 1.5 kg and spring constant k = 2.3 N/cm is at rest
when it is subjected to two impulsive forces: one at t = 0 of magnitude F̂1 = 2 N·sec and
a second impulse at t = 3 seconds of magnitude F̂2 = −1 N·sec. Find the velocity at t = 8
sec.
6.33 (M) The rod pendulum in Figure 6.47 weighs 250 lb and is of length 3 ft. A damper with
viscous damping coefficient of c = 3 lb sec/ft connects it to a wall, as shown. The system
is at rest with θ = 0 when it is struck by a horizontal impulsive force of magnitude 2,500
lb at point B which lasts for a period of 10−3 seconds. a) Calculate the response using the
impulsive motion assumption. b) Check the validity of the impulsive motion assumption.
Response of Dynamical Systems 299

! &

%#$
$!"%
" %#&
ș (%
'
#

FIGURE 6.47
Figure for Problem 6.33.

c) The system is hit by another impulse of the same magnitude and time duration at time
t = 5 seconds. Calculate the response.

Section 6.13—Step Response


6.34 (M) A mass-spring system with m = 3 kg and k = 243 N/m is at rest when it is excited
by a constant force of 13 N for seven seconds. Find the position and velocity of the mass at
time 42 seconds.
6.35 (M) The peak amplitudes of the cycles of a second-order system are measured as 1,
0.8, 0.6, 0.4 ft, indicating Coulomb damping. The period of oscillation is measured to be
1.1 seconds. Calculate the coefficient of friction.
6.36 (M) Given the following values of a mass-spring system acted upon by friction, calculate
the number of cycles it takes for motion to stop; x0 = 0.8 m, ωn = 3 rad/sec, µk = 0.16,
µs = 0.20.

" #
!

$ &

FIGURE 6.48
Figure for Problem 6.37.

6.37 (D) Figure 6.48 is a schematic of a payload of mass m = 200 kg, being transported
by a pickup truck of mass M = 4000 kg, where the payload is connected to the bed of the
truck by an energy absorbing support, modeled as a spring of constant k = 1250 N/m and
damper c = 180 N·s/m. The truck begins from rest and accelerates at a rate of 0.15g for
6 seconds, after which it moves with constant speed. Calculate the amplitude of x at a)
t = 10 s, b) t = 15 s.
6.38 (C) Given the mass-spring-damper system resting on a rough surface with coefficient of
300 Applied Dynamics

friction µ, where m = 2 kg, k = 18 N/m, write a computer program to obtain the response
for the following coefficients of friction µ = 0.01, 0.02, 0.04, with c = 0. The initial conditions
are x (0) = 0.4 m, ẋ (0) = 0.7 m/s. Then, generate the same plots when c = 1.6 N·s/m.
Calculate the values of the first three peaks of each of the six plots. You can do this either
by plotting and measuring from the plot or by printing the response and selecting the peaks
from the plot. Compare the way the peaks reduce in value for the damped and undamped
cases.

Section 6.14—Response to General Excitations


6.39 (M) A mass-spring system weighs 32.17 lb and has a stiffness of 192 lb/in. It is subjected
to an excitation of F (t) = 3t lb, with zero initial conditions. Obtain the response by
evaluating the convolution integral.
6.40 (M) A mass-spring system with m = 2 kg and k = 18 N/m is subjected to a constant
excitation of F (t) = 5 N, with the initial conditions x0 = 0, v0 = 0.3 m/s. Obtain the
response by evaluating the convolution integral.
6.41 (M) A mass-spring system of mass 12 kg and stiffness 192 N/m is subjected to an
excitation of F (t) = 10e−2t N. The initial conditions are zero. Calculate the response
(using convolution integral or Laplace transform). Plot the response for 12 seconds.
6.42 (M) Consider the previous problem and add a damper that provides a damping factor
of ζ = 0.2. Calculate the damping constant and obtain an expression for the response. Plot
the response for 11 seconds.

Section 6.16—Response to Harmonic Excitation


6.43 (M) An undamped mass-spring system has mass of 200 gm and stiffness of 0.6 N/m. The
system is acted upon by a harmonic excitation of magnitude 0.3 N with period 3 seconds.
Calculate the steady-state amplitude. Then, add a damper to the system. Calculate the
damping coefficient so that the steady-state amplitude of the damped system is 85% of the
amplitude of the undamped system.
6.44 (M) An undamped mass-spring system has a mass of 2 kg and stiffness coefficient of
18 N/m. The system is subjected to an external force F (t) = 0.8 sin(2t) + 0.2 sin(4t). Find
the amplitude of the steady-state response. Then, plot the excitation F (t) as a function of
time for 12 seconds.
6.45 (M) You are designing the base of a 321.7 lb machine. The machine has a rotating
component that produces a harmonic excitation of amplitude F = 30 lb and excitation
frequency of ω = 5 rad/sec. The base is being modeled as a spring and you have to design
the spring constant. Find the range of values of the spring constant k, such that the vibration
amplitude of the machine is less than 2 inches.
6.46 (M) A body of weight 32.17 lb is connected to a base with a spring whose constant is
k = 81 lb/ft and a damper of constant c = 1.8 lb sec/ft. The mass is excited harmonically
with an amplitude of F = 8 lb. Find the steady-state displacement amplitude of the mass
when the excitation frequency ω is a) ω = 2 rad/sec, b) ω = 4 rad/sec, c) ω = 6 rad/sec.
6.47 (C) A mass-spring-damper system with m = 10 kg, k = 90 N/m is acted upon by
a harmonic force F = F0 eiωt , with F0 = 18 kg. Plot the frequency response in the range
0 ≤ ω ≤ 5 rad/s for the following cases: a) c = 6 N·s/m, b) c = 16 N·s/m, c) c = 24 N·s/m.
6.48 (C) Consider an undamped mass-spring system of m = 1.3 kg and k = 6.5 N/m.
The system is at rest when a harmonic force F (t) = 1.1 sin(2.05t) is applied to it. Write a
computer program to calculate the response as a function of time and plot it for 15 seconds.
Then, add a damper of c = 1.05 N·s/m to the system and plot the damped response.
Comment on the nature of the plot.
6.49 (C) A mass-spring system has mass 2500 kg and natural frequency 9 rad/s. The system
Response of Dynamical Systems 301

is acted upon by a harmonic force F0 eiθ of magnitude F0 = 1200 N. The operating frequency
of the excitation is ω = 180 rpm. The spin speed of the force is ω = θ̇. The excitation starts
with θ (0) = 0, ω (0) = 0 and accelerates at a constant rate of θ̈ = ω̇ = α radians/sec2
until the operating angular velocity is reached. Write a computer program to integrate the
equation of motion while the harmonic force revs up and the operating speed is reached in
a) 3 seconds, b) 5 seconds, c) 7 seconds. Plot the vibration amplitudes for the three cases.

Section 6.17—Resonance
6.50 (C) A mass-spring-damper system with m = 10 kg, k = 90 N/m, and c = 6 N·s/m is
at rest when a harmonic force F = F0 cos ωt, with F0 = 18 N acts on it for five seconds.
Plot the response of the system for 12 seconds for the following values of the excitation
frequency ω = 2.5, 2.8, 3, 3.2, and 3.5 rad/sec.
6.51 (C) A mass-spring-damper system with m = 321.7 lb, k = 90 lb/ft, and c = 2 lb·sec/ft
is at rest when a harmonic force F = F0 sin ωt, with F0 = 25 lb and ω = 0.8ωn acts on it for
11 seconds. Then, there are two seconds of no forcing, followed by five seconds of resonant
excitation (ω = ωn ) with the same forcing of 25 lb. The excitation ends after that. Plot
the response for 30 seconds. Comment on how much the motion amplitude increases during
resonance.

Sections 6.18 and 6.19—Transmitted Force and Base Excitation


6.52 (M) Consider Problem 6.43 and calculate the transmitted force for the undamped as
well as damped cases.

3m 3m 3m

FIGURE 6.49
Figure for Problem 6.53.

6.53 (M) A roadblock is designed as a series of asphalt humps stretching across the road
surface of the road and spaced 3 m apart, as shown in Figure 6.49. A vehicle with undamped
frequency of 1 Hz and damping factor of ζ = 0.2 is traveling over the roadblock. Calculate
the speed range for the vehicle such that the transmissibility of the vehicle stays over 1.25.
6.54 (M) A vehicle of weight (quarter car) 800 lb and undamped natural frequency of 6
rad/s and damping factor of ζ = 0.1 is traveling over a wavy road. The road is modeled as
a sinusoidal with period of 10 m and height 13 cm. Which speed ranges should the vehicle
avoid, so that the amplitude of the car is less than 25 cm? How will this range change, if
the damping factor becomes ζ = 0.2?
6.55 (M) A support in the form of a spring is installed under a precision machine of mass
m = 2500 kg on a factory floor because the floor is moving due to the motion of other
machinery nearby. The support has stiffness k = 40,000 N/m. The floor moves harmonically
with amplitude 0.1 cm and frequency Ω. Calculate the force transmitted to the equipment
when a) Ω = ωn , b) Ω = 2ωn , c) Ω = ωn /2.
6.56 (C) Consider Problem 6.53. Calculate the wheel load (you are considering a one
302 Applied Dynamics

d.o.f. model) and plot the steady-state values of the wheel load as a function of vehicle
speed, considering that the mass of the vehicle is 450 kg. Consider a speed range of 0–24
m/s.

Section 6.20—Harmonic Excitation Due to Imbalances and Eccentricity


6.57 (M) A mass-spring system has total weight of 2000 lb and natural frequency 8 rad/s.
The system has an imbalance, in the form of a rotating component of mass We = 60 lb and
a moment arm of e = 6 in, which spins at an operating frequency of ω = 200 rpm. Calculate
the steady-state amplitude of the mass.
6.58 (M) The mass ratio of a rotating imbalance is me /m = 0.07. The vibrating system has
a natural frequency of ωn = 9 rad/s and damping factor of ζ = 0.15. The driving frequency
of the imbalance is 700 rpm. The steady-state amplitude of the primary mass is observed to
be 0.01 m. Calculate a) the size of the imbalance arm e and b) the steady-state amplitude
of the primary mass when the driving frequency of the imbalance is 300 rpm.
6.59 (C) A mass-spring system has total mass of 2500 kg and natural frequency of 9 rad/s.
The system has an imbalance, in the form of a rotating component of mass me = 50 kg
and a moment arm of e = 12 cm, which spins at an operating frequency of ω = 200 rpm.
The spin speed of the imbalance is ω = θ̇. The imbalance starts at rest with θ (0) = 0 and
accelerates at a constant rate of θ̈ = ω̇ = α radians/sec until the operating speed is reached.
Write a computer program to integrate the equation of motion while the imbalance revs
up and the operating speed is reached in a) 3 seconds, b) 5 seconds, c) 7 seconds. Plot the
vibration amplitudes for the three cases.
7
Response of Multi-Degrees-of-Freedom Systems

7.1 Introduction
The previous chapter developed procedures for modeling and obtaining the response of
single-degree-of-freedom systems. Here, those developments are extended to systems with
more than one degree of freedom. We will follow the same approach for problem formulation
and solution. We will make use of matrix algebra.
A single-degree-of-freedom system has one natural state of motion when it is subjected to
an excitation or an initial condition. A multi-degrees-of-freedom system has as many natural
states as the number of degrees of freedom. In each natural state, the system assumes a
certain shape. This chapter develops procedures for finding these natural states.
The equations of motion of multi-degrees-of-freedom (m.d.o.f.) systems are usually in
the form of coupled equations, which cannot be solved independently of each other. We
develop ways of transforming a set of coupled equations into a set of uncoupled equations.
The uncoupled equations can then be solved independently of each other. It turns out that
the procedure that obtains the natural states also decouples the equations of motion.
The chapter develops an approximate model to deal with damped systems. It also and
discusses a relevant application of two-degrees-of-freedom systems, vibration absorbers.

7.2 Modeling of Multi-Degrees-of-Freedom Systems


We begin with the modeling of multi-degrees-of-freedom systems. The approach builds on
the developments in Chapters 5 and 6, which is as follows:
1. Determine the number of degrees of freedom and select a set of variables to describe the
motion. Define the positive directions of the motion variables, identify the points from
which they are measured, and select coordinate systems.

2. Separate the system into individual components and draw free-body diagrams.
3. Write force and moment balances.
4. Combine the force and moment balances, as well as kinematic relationships, to obtain
the equations of motion. The equations of motion should not have any constraint and
redundant forces, or unknown parameters. There should be as many equations of motion
as there are d.o.f.1
1 There are techniques where we describe the response by using a set of equations whose number is larger

than the degrees of freedom. Such approaches are discussed in Chapter 8.

303
304 Applied Dynamics

5. If necessary, linearize the equations of motion about an operating point, such as an


equilibrium position or the unstretched position of elastic members.
6. Write the equations of motion in matrix form.
The difference in this chapter is the matrix formulation. The general form of the equa-
tions of motion of a linear multi-degrees-of-freedom system with n degrees of freedom is

[m] {ẍ (t)} + [c] {ẋ (t)} + [k] {x (t)} = F (t) (7.1)

where {x (t)} is the position vector of order n, {F (t)} is the force vector of order n, and the
n × n matrices [m], [c], and [k] are the mass, damping, and stiffness matrices, respectively.
When the equations of motion are written using the procedure outlined below, and when
dealing with a stable or critically stable system, the coefficient matrices have the following
properties:
• The mass matrix [m] is positive definite.
• The damping matrix [c] is positive definite or positive semi-definite.
• The stiffness matrix [k] is positive definite or positive semi-definite.

The preferred procedure for writing the equations of motion in matrix form is to write
them in the order that the coordinates in the position vector are selected. For example, if
we select {x (t)} as {x (t)} = [x1 (t) x2 (t)]T , the equation of motion associated with x1 (t)
is written first and then the equation of motion associated with x2 (t) is written next. If the
order is reversed, the resulting matrix equations will not be incorrect, but the symmetry
properties of the coefficient matrices will be lost. We refer to writing the equations of motion
such that the coefficient matrices are symmetric as the zeroth convenience.
We can actually prove that the mass, stiffness, and damping matrices can be written in
symmetric form. It will be shown in Chapter 8, when studying Lagrangian mechanics, that
the kinetic and potential energies can be written in quadratic form as
1 1
T = {ẋ (t)}T [m] {ẋ (t)} V = {x (t)}T [k] {x (t)} (7.2)
2 2
which lead to symmetric mass and stiffness matrices.
To ascertain the equation of motion that is primarily associated with a certain motion
variable, it is helpful to examine the second derivatives and the units and dimensions of
the second derivative terms. For example, for a translational variable, say, x, the force
balance expression should have a term such as mẍ, where m is a mass parameter, and the
dimension of mẍ is mass × acceleration = force. For a rotational variable, say, θ, we look
for a moment equation that results in a term such as mL2 θ̈, which has the units of mass ×
length2 / time2 = moment.

7.2.1 Definition of Sign Definiteness


A symmetric matrix is said to be positive definite (positive semi-definite) if it has the
following properties:
1. All diagonal elements are > 0 (≥ 0), and

2. All minor determinants are > 0 (≥ 0), and


3. The main determinant is > 0 (≥ 0).
Response of Multi-Degrees-of-Freedom Systems 305

Note that positive definiteness is defined only for symmetric matrices. A nonsymmetric
matrix cannot be positive definite. Given a positive definite (positive semi-definite) matrix
[A] of order n × n and a nonzero vector {v} of order n, the scalar product {v}T [A] {v} has
the following property:
T
{v} [A] {v} > 0 (≥ 0) (7.3)

Positive definite (positive semi-definite) matrices have the property that all their eigen-
values are > 0 (≥ 0). Consider the eigenvalue problem for a symmetric matrix [A] in the
form

[A] {v} = λ {v} (7.4)

Left multiplying the above equation by {v}T gives


T T
{v} [A] {v} = λ {v} {v} (7.5)

For nonzero {v}, the product {v}T {v} is always greater than zero. Using the result in
Equation (7.3) that {v}T [A] {v} > 0, we conclude that the eigenvalues are λ > 0 (λ ≥ 0 for
positive semi-definite [A]).
For the eigenvalue problem defined by two matrices

[A] {v} = λ [B] {v} (7.6)

in which [B] is positive definite and [A] is positive definite (positive semi-definite), it turns
out that all the eigenvalues are λ > 0 (≥ 0). To show this, left multiply the above equation
by {v}T and solve for λ, which gives
T
{v} [A] {v}
λ = T
(7.7)
{v} [B] {v}

The numerator {v}T [A] {v} is always > 0 (≥ 0). The denominator is always positive,
{v}T [B] {v} > 0, leading to the conclusion that λ > 0 when [A] is positive definite and
λ ≥ 0 when [A] is positive semi-definite.

Example 7.1
Determine if the matrices below are positive definite.
 
4 2 −1  
4 3
[A] =  2 4 −2  [B] =
5 6
−1 −2 3
Matrix [B] is not positive definite, as it is not symmetric. Matrix [A] is symmetric and
its diagonals are all > 0. The next step is to calculate the minor determinants, beginning
at the upper
 left corner. The first minor determinant is 4. The second is the determinant
4 2
of , which is 16 − 4 = 12. The main determinant is
2 4

det [A] = 4 × 4 × 3 + 2 × (2 × −2 × −1) − (−1 × 4 × −1)

−(4 × −2 × −2) − (2 × 2 × 3) = 24 [a]


Matrix [A] satisfies all three criteria and hence is positive definite.
306 Applied Dynamics

g
k

O F
M y 
c x
x m, 2L


FIGURE 7.1
Cart and rod.

$ "#!"#ș
$
!

&#"'"() ș
*%!"'"%# "#$%!ș
ș

FIGURE 7.2
Displaced positions of cart and rod.

Example 7.2
A rod of mass m and length 2L is connected by a pin joint to a cart of mass M , as shown
in Figure 7.1. Find the equations of motion, linearize about equilibrium, and write the
equations in matrix form. Check the sign definiteness of the coefficient matrices.
Step 1. There are two degrees of freedom. The motion variables are selected as the dis-
placement of the cart x (positive to the right) and the angle θ (positive counterclockwise)
that the rod makes with the vertical. The displaced positions of the cart and rod are shown
in Figure 7.2.
Step 2. The second step involves separating the cart and rod and drawing the free-body
diagrams, as shown in Figure 7.3, together with the displaced coordinates. There are two
reaction forces at the point of separation O, Ox , and Oy . These forces act in opposite
directions on each component.
Step 3. The force and moment balances are written next. For the cart, the sum of forces
is written in the horizontal and vertical directions. In the horizontal direction, the sum of
forces is
+
X X
→ F = Ma a = ẍ F = −kx − cẋ + Ox + F [a]
which can be rewritten as
M ẍ + cẋ + kx = Ox + F [b]
Response of Multi-Degrees-of-Freedom Systems 307

Mg
Oy Oy
kx F
O Ox O
c Ox
y
G
N1 N2

x 2L
mg

FIGURE 7.3
Free-body diagrams.

In the vertical direction, the force balance is


X
+↑ F = 0 = N1 + N2 − M g + Oy [c]

The above equation gives the values of the reaction forces between the ground and the cart.
Next, consider the rod and write the force balances in the horizontal and vertical direc-
tions, as well as the sum of moments about the center of mass G. To this end, expressions
need to be developed for the acceleration of the center of mass in the horizontal and vertical
directions. Noting that the position vector for the center of mass is
rG = (x + L sin θ) i − L cos θj [d]
the accelerations are obtained by differentiating the above equation twice with respect to
time2 as
d2 d  
aGx = (x + L sin θ) = ẋ + L θ̇ cos θ = ẍ + Lθ̈ cos θ − Lθ̇2 sin θ [e]
dt2 dt
d2 d  
aGy = (−L cos θ) = Lθ̇ sin θ = Lθ̈ sin θ + Lθ̇2 cos θ [f ]
dt2 dt
Summing forces for the rod in the horizontal direction gives
 
+
X
→ F = maGx =⇒ m ẍ + Lθ̈ cos θ − Lθ̇2 sin θ = −Ox [g]

and in the vertical direction


X  
+↑ F = maGy =⇒ m Lθ̈ sin θ + Lθ̇2 cos θ = −Oy − mg [h]

Next, consider the moment balance for the rod about the center of mass and write
X X
MG = IG θ̈, MG = Ox L cos θ + Oy L sin θ [i]
1
where IG = 12 m(2L)
2
= 13 mL2 , so that the above equation can be written as
1
mL2 θ̈ = Ox L cos θ + Oy L sin θ [j]
3
2 We can also obtain the acceleration of the center of mass of the rod by using the relative acceleration

expression aG = aO + α × rG/O + ω × ω × rG/O , with aO = ẍi, rG/O = L sin θi − L cos θj, ω = θ̇k, α = θ̈k.
308 Applied Dynamics

There are five force and moment balances, Equations [b], [c], [g], [h], and [j]. The five
unknowns are x, θ, Ox , Oy , and N1 + N2 .3 Note that the moment equation about a fixed
P
point O, MO = IO θ̈, is not valid for this problem, as point O is not stationary.
Step 4. This step involves combining the force and moment balances to obtain the two
equations of motion. There are several ways of doing this, but the approach should take
into consideration that it is preferable to write the equations of motion in terms of symmetric
matrices. Since Equation [b] has Ox on its right side and Equation [g] has −Ox on its right
side, adding these two equations eliminates Ox and results in
 
M ẍ + cẋ + kx + m ẍ + Lθ̈ cos θ − Lθ̇2 sin θ = Ox − Ox + F [k]

The above equation can be rewritten as

(M + m) ẍ + mLθ̈ cos θ − mLθ̇2 sin θ + cẋ + kx = F [l]

Equation [l] is recognized as one of the equations of motion, namely the one associated
with x, as both ẍ, m and M appear in it. Also, the units of every term in the above equation
are force, or mass×acceleration. Equation [j] is used to obtain the second equation of motion.
Note that Equation [j] is a moment balance, and it has Ox and Oy in it. Substituting for
the values of these reaction forces from Equations [g] and [h] yields
1  
mL2 θ̈ = −mL cos θ ẍ + Lθ̈ cos θ − Lθ̇2 sin θ
3
 
−mL sin θ Lθ̈ sin θ + Lθ̇2 cos θ − mgL sin θ [m]
which reduces to
4
mL2 θ̈ + mLẍ cos θ + mgL sin θ = 0 [n]
3
Equation [n] is the second equation of motion, and it is associated with θ. Note that
the unit of each term in the above equation is moment, or mass moment of inertia times
angular acceleration. The equations of motion are nonlinear.
Step 5. The equilibrium positions are found and the equations of motion are linearized
about equilibrium. All time derivatives are zero at equilibrium and the external forcing is
set to zero, or F = 0. The first equation of motion, Equation [l], and second equation of
motion, Equation [n], calculated at equilibrium give

kx = 0 mgL sin θ = 0 [o]

The equilibrium equations are uncoupled and can be solved individually, with the result

xe = 0 θe = 0, π [p]

The equilibrium position θe = 0 is when the rod is below the cart and is the stable position.
In the position θe = π, the rod is upright and the equilibrium position is unstable.
Let us linearize about the stable equilibrium configuration, xe = 0, θe = 0. Since at
this equilibrium position all variables are zero, we can use the small angles assumption
(sin θ ≈ θ, cos θ ≈ 1), which gives the linearized equations of motion as
4
(M + m) ẍ + mLθ̈ + cẋ + kx = F mL2 θ̈ + mLẍ + mgLθ = 0 [q]
3
3 Note that because the cart only translates and its dimensions are not given, it can be treated as a

particle. Hence, we cannot distinguish between N1 and N2 and we can only calculate their sum.
Response of Multi-Degrees-of-Freedom Systems 309

Step 6. This step involves writing the equations of motion in matrix form. Introducing the
 T
position vector {x} = x θ , the equations of motion are expressed by first writing the
equation of motion for x and next the equations of motion for θ as
          
M +m mL ẍ c 0 ẋ k 0 x F
4 2 + + = [r]
mL 3 mL θ̈ 0 0 θ̇ 0 mgL θ 0
or
[m] {ẍ} + [c] {ẋ} + [k] {x} = {F } [s]
All coefficient matrices are symmetric. The mass and stiffness matrices are positive
definite and the damping matrix [c] is positive semi-definite.

Example 7.3
Consider the previous example and linearize the equations of motion about the unstable
equilibrium point, θe = π. Compare the coefficient matrices with the stable equilibrium
case.
Introduce the variable , so that the angular displacement can be expressed in terms of
the equilibrium position as
θ = θe +  = π +  [a]
Next, expand the sine and cosine of θ by using a small angles approximation for . Noting
that sin π = 0, cos π = −1,

sin θ = sin (π + ) = sin π cos  + cos π sin  = − sin  ≈ − [b]

cos θ = cos (π + ) = cos π cos  − sin π sin  = − cos  ≈ −1 [c]


It also follows that θ̇ = ,
˙ θ̈ = ¨. Introducing these values to the equations of motion,
Equations [l] and [n] in the previous example, and eliminating terms nonlinear in  (θ̇2 =
˙2 ≈ 0) gives
4
(M + m) ẍ − mL¨
 + cẋ + kx = F mL2 ¨ − mLẍ − mgL = 0 [d]
3
Introducing the position vector {x} = [x ]T , we can write the linearized equations in
standard form, where
     
M + m −mL c 0 k 0
[m] = 4 2 [c] = [k] = [e]
−mL 3 mL 0 0 0 −mgL

The mass matrix is positive definite and the damping matrix is positive semi-definite,
as in the linearized equations for the stable equilibrium position. However, the stiffness
matrix has a negative element in one of the diagonals and hence has no sign definiteness.
Indefiniteness of the stiffness matrix is an indication of instability. The equilibrium position
is unstable. An eigenvalue analysis, which will be conducted in subsequent sections, indicates
that the equilibrium position θe = π is indeed unstable.

Example 7.4
Consider the mass-spring-damper system in Figure 7.4. Derive the equations of motion for
when c3 = c4 = 0 and write them in matrix form.
Step 1. There are three degrees of freedom and the motion variables are selected as the
displacements of the masses and denoted by x1 , x2 , and x3 . The motion variables are
310 Applied Dynamics

F1 F2 F3
k1 k2 k3 k4

m1 m2 m3

c1 c2 c3 c4

x1 x2 x3

FIGURE 7.4
Mass-spring-damper system.

F1 F2 F3
k2(x2 – x1)
k1x1
m1 m2 k3(x3 – x2) m3 k4x4
c1 1
c2( 2 – 1)

FIGURE 7.5
Free-body diagrams for when c3 = c4 = 0.

measured from the unstretched positions of the springs. The positive direction for each
coordinate is to the right.
Step 2. The bodies are separated and the free-body diagrams are drawn, as shown in Figure
7.5. We can look at the force generated by a spring between two masses in two different
ways. Consider k2 , which is between m1 and m2 . If we hold m2 fixed and move m1 in the
positive direction by x1 , the spring k2 is compressed by x1 and the force acting on m1 is
k2 x1 and it is acting to the left (←). An equal and opposite force acts on m2 , towards the
right (→).
If the second mass is moved by a distance x2 in the positive direction, the stretch of
spring k2 becomes x1 − x2 and the spring force becomes k2 (x1 − x2 ) acting to the left on
m1 . The same force acts to the right on m2 .
If we view the spring force associated with k2 by looking at the displacement x2 first
and initially holding m1 fixed, the spring force can be shown to have the form k2 (x2 − x1 )
but now the spring force that acts on m2 is leftward (←). The same force acts to the right
(→) on m1 .
It is not important which approach you select for writing the spring forces. What is
important is to be consistent.
Step 3. The next step is to sum forces for each mass. Noting that the acceleration of each
mass is ai = ẍi , (i = 1, 2, 3), the force balances become
+
X
Mass 1: → F = m1 ẍ1 = −k1 x1 + k2 (x2 − x1 ) − c1 ẋ1 + c2 (ẋ2 − ẋ1 ) + F1 [a]

+
X
Mass 2: → F = m2 ẍ2 = −k2 (x2 − x1 ) + k3 (x3 − x2 ) − c2 (ẋ2 − ẋ1 ) + F2 [b]

+
X
Mass 3: → F = m3 ẍ3 = −k4 x3 − k3 (x3 − x2 ) + F3 [c]
Response of Multi-Degrees-of-Freedom Systems 311

Steps 4–6. There are no reaction forces to eliminate and no terms to linearize, so all that
is needed is to rewrite the above equations in standard form and to put them in matrix
form. Rearranging, the equations of motion are obtained as

m1 ẍ1 + (c1 + c2 ) ẋ1 − c2 ẋ2 + (k1 + k2 ) x1 − k2 x2 = F1 [d]

m2 ẍ2 − c2 ẋ1 + (c1 + c2 ) ẋ2 − k2 x1 + (k2 + k3 ) x2 − k3 x3 = F2 [e]

m3 ẍ3 − k3 x2 + (k3 + k4 ) x3 = F3 [f ]
The equations are ready to be cast into matrix form. Defining the position vector {x} =
[x1 x2 x3 ]T and force vector {F } = [F1 F2 F3 ]T , the equations of motion can be written in
the standard form of Equation (7.1), with the coefficient matrices as
   
m1 0 0 c1 + c2 −c2 0
[m] =  0 m2 0  [c] =  −c2 c2 0 
0 0 m3 0 0 0
 
k1 + k2 −k2 0
[k] =  −k2 k2 + k3 −k3  [g]
0 −k3 k3 + k4
We can show that all matrices are symmetric, [m] and [k] are positive definite, and
[c] is positive semi-definite. An interesting property of the damping and stiffness matrices
emerges from Equation [g]. Take, for instance, the second equation of motion, Equation [e].
The spring k2 connects m2 to m1 and k3 connects m2 to m3 . The coefficient of x2 is k2 + k3
and it is positive. The coefficients of x1 (−k2 ) and of x3 (−k3 ) are negative quantities. The
same holds true for the damping matrix. This is a property of systems modeled as a series
of masses, springs, and dampers. It also is a good way to spot errors when deriving the
equations of motion and calculating the coefficient matrices.

Example 7.5
Consider the mass-spring-damper system in the previous example and write the kinetic and
potential energies. Then, express them in quadratic form.
The kinetic energy is
1 1 1
T = m1 ẋ21 + m2 ẋ22 + m3 ẋ23 [a]
2 2 2
and the potential energy is
1 1 2 1 2 1
V = k1 x21 + k2 (x2 − x1 ) + k3 (x3 − x2 ) + k4 x23 [b]
2 2 2 2
2
Consider the second term in the potential energy, 12 k2 (x2 − x1 ) , which can be expanded
as
1 2 1
k2 (x2 − x1 ) = k2 x21 − 2x1 x2 + x22

[c]
2 2
Introducing the vector {y} = [x1 x2 ]T , we can express Equation [c] as
 
1 2 1 T k2 −k2
k2 (x2 − x1 ) = {y} {y} [d]
2 2 −k2 k2
312 Applied Dynamics
1 2
Note that we can express 2k2 (x2 − x1 ) also as
 
1 2 1 T k2 −2k2
k2 (x2 − x1 ) = {y} {y} [e]
2 2 0 k2
where the coefficient matrix is not symmetric. We see once again that writing the coefficient
matrices in symmetric form, both for equations of motion and energy expressions, is done
for convenience and for taking advantage of the properties of eigenvalue problems associated
with symmetric matrices.
Extending the procedure above to all the terms in the potential energy and using the
position vector {x} = [x1 x2 x3 ]T , we can express the kinetic and potential energies as
Equation (7.2), with the mass and stiffness matrices defined in Equation [g] of the previous
example. For a nonlinear system, we can also write the kinetic and potential energies in
their general form, expand them about equilibrium, and retain the quadratic terms.

7.3 Coupling
As demonstrated in the preceding examples, for a multi-degrees-of-freedom system, the
equations of motion, for the most part, are coupled to each other. They cannot be solved
individually, that is, independent of each other.
We can observe the existence of coupling in the equations of motion by the presence of
off-diagonal terms in the coefficient matrices. In Example 7.2, coupling is present through
off-diagonal terms in the mass matrix and in Example 7.4, through off-diagonal terms in
the stiffness and damping matrices.
It turns out that coupling is not a property of a dynamical system. Rather, coupling is
a function of the variables selected to describe the motion. It is possible to select a set of
coordinates such that the coupling is transferred from one coefficient matrix to another.
The question can be asked whether it is possible to find a set of motion variables that
will enable us to write the equations of motion in uncoupled form, so all coefficient matrices
will be diagonal. If so, it becomes possible to solve the equations of motion independently
of each other, making use of the methods developed in the previous chapter. The answer
to this question is affirmative, and this chapter develops methods that identify the motion
variables that lead to uncoupled equations of motion.
In the age of powerful computers and computational methods we can ask: Why not
integrate the coupled equations of motion numerically, without worrying about analytical
solutions or coupling? After all, we can integrate nonlinear as well as linear equations, thus
eliminating the need to find equilibrium positions and to linearize. The disadvantage of
numerical integration is that it represents a totally quantitative approach, it gives little
insight into the nature of the motion and its results are valid only for the given set of initial
conditions and excitation. Numerical integration becomes more valuable when combined
with qualitative analysis.

7.4 Free Motion of Undamped Multi-Degrees-of-Freedom Systems


As in the case of single-degree-of-freedom systems, it is instructive to begin the analysis of
multi-degrees-of-freedom systems by first considering the free motion and no damping. The
Response of Multi-Degrees-of-Freedom Systems 313

equations of motion reduce to

[m] {ẍ (t)} + [k] {x (t)} = {0} (7.8)

with initial conditions {x (0)} and {ẋ (0)}.


The synchronous motion assumption will be used to analyze the motion. The syn-
chronous motion assumption basically states that different components of a system move in
a similar fashion (the components are in sync with each other). Their amplitudes may be dif-
ferent, but their behavior over time is similar. For example, if one component comes to rest,
so do the other components. Or, if a component begins to vibrate, the other components
soon follow. You can demonstrate this concept by slightly tipping one of the ornaments
(e.g., pendant, prism) of a chandelier. Soon, all the other ornaments will also be swinging.
Considering synchronous motion, the evolution of the system in time can be expressed
as

{x (t)} = {u} eλt (7.9)

where {u} is an amplitude vector and eλt denotes the time dependence. At this stage,
both {u} and λ are unknown. Recall that for a single-degree-of-freedom system, the general
free response is x (t) = Ceλt . Introducing Equation (7.9) into the equations of motion and
collecting terms gives

λ2 [m] + [k] {u} eλt = {0}



(7.10)

For a nonzero solution, the term eλt cannot be zero, so we must have

λ2 [m] + [k] {u} = {0}



(7.11)

For the above equation to have a nontrivial solution, that is, {u} 6= {0}, the coefficient
matrix must be rank deficient. In other words, its determinant must be zero. Setting the
determinant to zero gives

det λ2 [m] + [k] = 0



(7.12)

Equation (7.12) is a polynomial of order n in λ2 . It is known as the characteristic


polynomial, or characteristic equation, or characteristic determinant. Its solution consists
of n roots, λ21 , λ22 , . . . , λ2n , which are known as eigenvalues. For the case when both [m]
and [k] are positive definite, all eigenvalues (roots) are real and negative. This conclusion
follows from the discussion in Section 7.2.1 regarding eigenvalue problems for two matrices.

Example 7.6
Consider the matrix equation
    
1 2 x 5
= [a]
3 4 y 6

Because the matrix is nonsingular its inverse exists4 and the solution becomes
   −1       
x 1 2 5 1 4 −2 5 −4
= = = [b]
y 3 4 6 4−6 −3 1 6 4.5
   
a b 1 d −b
4 Given a matrix [A] = , the inverse of the matrix is [A]−1 = .
c d ad−bc −c a
314 Applied Dynamics

The solution of the problem


    
1 2 x 0
= [c]
3 4 y 0

is x = 0, y = 0. The only way a set of matrix equations [A]{x} = {0} can have a nonzero
solution for {x} is when det[A] = 0. For example, for the equation
    
1.5 2 x 0
= [d]
3 4 y 0

the coefficient matrix is singular, as the second row is twice the first row. Thus, there is
only one equation to solve, 1.5x + 2y = 0, whose solution is y = −0.75x. It follows that
there is an infinite number of solutions that satisfy the relationship y = −0.75x.

7.4.1 Natural Frequencies


It is more convenient to deal with positive quantities for the solution of the characteristic
equation. To this end, define the quantity ω such that ω 2 = −λ2 and rewrite the character-
istic equation, which is also referred to as the frequency equation, as

det −ω 2 [m] + [k] = 0



(7.13)

The solution of the above equation consists of the natural frequencies ωj2 , (j =
1, 2, . . . , n) so that

ωj2 = −λ2j j = 1, 2, . . . , n (7.14)

The n natural frequencies describe the n unique ways (or modes)that the system can
oscillate. For a single-degree-of-freedom system, there is one natural frequency, and an n-
degrees-of-freedom system has n natural frequencies.
 The natural frequencies are the only
quantities that make the matrix −ω 2 [m] + [k] singular. This property is a useful tool in
spotting errors when calculating the natural frequencies.
For the purposes of convenience, the natural frequencies will be ordered in ascending
order, so that ω1 ≤ ω2 ≤ . . . ≤ ωn . We refer to this ordering scheme as the first convenience.
The ordering scheme takes into consideration the property that the lower natural frequencies
contribute more to the response (as will be demonstrated later in this chapter). The lowest
natural frequency, ω1 , is known as the fundamental frequency.
Natural frequencies of most systems are distinct. Cases in which repeated natural fre-
quencies (ωj = ωj+1 ) are encountered are referred to as degenerate. They mainly arise in
two-dimensional (planar) structures with geometric symmetry. The reader is referred to
texts on vibrations or mathematical analysis for further details. We will see in Chapter 10
an interesting example of degenerate systems.

Example 7.7
Consider the linearized equations about stable equilibrium (xe = 0, θe = 0) of the cart-rod
system in Example 7.2 and obtain the natural frequencies for M = 4, m = 3, L = 1, k =
8, g = 10, c = 0.
From Equation [r] in Example 7.2 the mass and stiffness matrices are
       
M +m mL 7 3 k 0 8 0
[m] = 4 2 = [k] = = [a]
mL 3 mL 3 4 0 mgL 0 30
Response of Multi-Degrees-of-Freedom Systems 315

and the characteristic equation becomes

8 − 7ω 2 −3ω 2
 
det −ω 2 [m] + [k] = det

= 0 [b]
−3ω 2 30 − 4ω 2

which, when expanded, leads to the characteristic equation

19ω 4 − 242ω 2 + 240 = 0 [c]

Solution of the characteristic equation yields the natural frequencies



2 242 ± 2422 − 4 × 19 × 240
ω1,2 = = 1.084, 11.653 (rad/sec)2 [d]
2 × 19
and, using the first convenience, the natural frequencies are ordered as
√ √
ω1 = 1.084 = 1.041 rad/sec ω2 = 11.653 = 3.414 rad/sec [e]

Example 7.8
Consider the linearized equations about unstable equilibrium (xe = 0, θe = π) of the cart-
rod system in Example 7.3 and assess stability for M = 4, m = 3, L = 1, k = 8, g =
10, c = 0.
From Equation [e] in Example 7.3 the mass and stiffness matrices are
       
M + m −mL 7 −3 k 0 8 0
[m] = 4 2 = [k] = = [a]
−mL 3 mL −3 4 0 −mgL 0 −30

Considering a solution {x (t)} = {u}eλt , the characteristic equation becomes

8 + 7λ2 −3λ2
 
det λ2 [m] + [k] = det

= 0 [b]
−3λ2 −30 + 4λ2

which, when expanded, leads to the characteristic equation

19λ4 − 178λ2 − 240 = 0 [c]

Solution of the characteristic equation is



2 −178 ± 1782 + 4 × 19 × 240
λ1,2 = = −10.56, 1.196 (rad/sec)2 [d]
2 × 19
One of the characteristic values is positive, indicating an exponentially growing, or unstable,
solution. Hence, as expected, the equilibrium point θe = π is unstable.

7.4.2 Modal Vectors


Associated with each natural frequency ωj , there is a real-valued amplitude vector {uj }
that is a solution of the eigenvalue problem and hence satisfies the equation

−ωj2 [m] + [k] {uj } = {0} j = 1, 2, . . . , n



(7.15)

The amplitude vector is known as the modal vector or eigenvector. The modal vector
{uj } gives the amplitudes of the motion variables when the system is vibrating with the
j-th natural frequency. The natural frequency ωj and associated modal vector {uj } make
316 Applied Dynamics

up the j-th mode of a system. An n-degrees-of-freedom system has n modes. The modal
vectors are denoted by
 
u1j
 u2j 
{uj } = 
 . . .  j = 1, 2, . . . , n
 (7.16)
unj
Note that, because the right side of Equation (7.15) is zero, the modal vectors can only
be calculated to within a multiplicative constant. This does not pose any problems with
the uniqueness of the solution, as the ratio between any two elements of a modal vector is
constant.
The signs of the elements of the modal vectors have an interesting property. These signs
indicate the direction in which the motion variables move in a particular mode. For the first
mode, that is, the modal vector associated with the lowest natural frequency, the elements
of the modal vector have the same signs.5 This means that all components are moving in
the same direction in the first mode.
There is one sign change in the second modal vector, indicating that all elements except
one are moving in one direction and one element is moving in the other direction. There
are two sign changes in the third modal vector. It follows that there are j − 1 sign changes
in the j-th modal vector. We can plot the modal vectors, which are also referred to as the
mode shapes. Viewing mode shapes is a good way of visualizing the amplitude ratios of the
components of a system.
The process of solving for the characteristic values and modal vectors is known as solving
the eigenvalue problem.

Example 7.9
Consider the cart-rod system in the previous examples. Calculate and plot the modal vectors
for motion about the stable equilibrium position (xe = 0, θe = 0).
The mass and stiffness matrices are
   
7 3 8 0
[m] = [k] = [a]
3 4 0 30
and the natural frequencies are
√ √
ω1 = 1.084 = 1.041 ω2 = 11.653 = 3.414 rad/sec [b]
 
u11
Denoting the first modal vector as {u1 } = , Equation (7.15) becomes
u21
 2 
−ω1 [m] + [k] {u1 } = {0}

    
8 − 7 × 1.084 −3 × 1.084 u11 0
=⇒ = [c]
−3 × 1.084 30 − 4 × 1.084 u21 0
We can verify that the matrix above is indeed singular. Either the top or the bottom of
the equations can be used to solve for the first modal vector. The top equation gives
(8 − 7 × 1.084) u11 − (3 × 1.084) u21 = 0 [d]
5 An exception can occur when the motion variables have different units, as in having a translational

coordinate and a rotational coordinate. In Example 15.4, which analyzes vehicle suspensions, there is one
sign change in the first mode and no sign changes in the second mode. The number of sign changes in each
mode is different, but the sign changes are not in order.
Response of Multi-Degrees-of-Freedom Systems 317

which is solved for u21 as


8 − 7 × 1.084
u21 = u11 = 0.1267u11 [e]
3 × 1.084
The first modal vector can be expressed as
 
1
{u1 } = d1 [f ]
0.1267

where d1 is an arbitrary constant. When solving an eigenvalue problem by hand, it is always


wise to check the accuracy of the calculations using another row of the matrix equation, in
this example, the second row of Equation [c]. The second row of Equation [c] leads to

(−3 × 1.084) u11 + (30 − 4 × 1.084) u21 = 0 [g]

which can be shown to give the same result as in Equation [e].


For the second mode, the corresponding eigenvector equation is
 2 
−ω2 [m] + [k] {u2 } = {0}

    
8 − 7 × 11.653 −3 × 11.653 u12 0
=⇒ = [h]
−3 × 11.653 30 − 4 × 11.653 u22 0
We can use either the top or bottom row of the above equation to solve for the elements
of {u2 }. Choosing the bottom row gives

(−3 × 11.65) 3u12 + (30 − 4 × 11.653) u22 = 0 [i]

which leads to the second modal vector as


 
1
{u2 } = d2 [j]
−2.104

{u1} {u2}

1 1

1 2 1 2

–2

FIGURE 7.6
Plots of the modal vectors. Notice the sign change in the second mode.

The modal vectors are plotted in Figure 7.6 for d1 = d2 = 1. In the first modal vector,
the amplitudes have the same sign, meaning that when the mass moves forward so does
318 Applied Dynamics

the rod. By contrast, in the second mode, the amplitudes have opposite signs, indicating
that the mass and rod move in opposite directions. The total motion can be viewed as a
combination of these two motions: 1) the mass and rod move together and 2) the mass
and rod move in opposite directions. Figure 7.7 shows the displacements of the cart and
pendulum for the two modes (not drawn to scale).

a) b)
u11 u12
u22

u21

Original Original
position position

FIGURE 7.7
Shapes assumed by system for a) first mode, b) second mode.

7.4.3 Normalization of Modal Vectors


As discussed earlier, the modal vectors are unique but their amplitudes can only be deter-
mined to within a multiplicative constant. It is customary to assign a value to the multi-
plicative constants. The process is called normalization and there are several conventions,
such as

• To make the top element of {uj } equal to 1. Expressing the j-th modal vector as
{uj } = [uj1 uj2 . . . ujn ]T , this convention sets uj1 = 1.
T
• To have the scalar product {uj } {uj } = 1, (j = 1, 2, . . . , n).
• To have the scalar product
T
{uj } [m] {uj } = 1 j = 1, 2, . . . , n (7.17)

This last convention is the most widely used. We refer to it as the second convenience.
Left multiplying Equation (7.15) with {uj }T and using Equation (7.17) gives
T
{uj } [k] {uj } = ωj2 (7.18)

It should be reiterated that normalization is done for convenience and is a process that
does not change the mode shapes or the response. The amplitude ratios among the elements
of modal vectors remain the same.

Example 7.10
Return to the cart-rod problem about stable equilibrium and normalize the modal vectors
using the second convenience.
Response of Multi-Degrees-of-Freedom Systems 319

For the first mode, the normalization procedure results in


 T   
T 1 7 3 1
{u1 } [m] {u1 } = d21
0.1267 3 4 0.1267
 T  
1 7 + 3 × 0.1267
= d21
0.1267 3 + 4 × 0.1267

= d21 7 + 3 × 0.1267 + 3 × 0.1267 + 4 × 0.12672 = 7.8244d21 = 1



[a]
Solving for d1 gives
1
d1 = ± √ = ±0.3575 [b]
7.8244
We have the choice of selecting the positive or negative value, again pointing to the
arbitrariness of the normalization factor. Let us select the positive value, so d1 = 0.3575.
For the second mode
 T   
T 1 7 3 1
{u2 } [m] {u2 } = d22
−2.104 3 4 −2.104
 T  
1 7 + 3 × (−2.104)
= d22
−2.104 3 + 4 × (−2.104)

= d22 7 + 3 × (−2.104) + 3 × (−2.104) + 4 × 2.1042 = 12.083d22 = 1



[c]
Solving for d2 gives
1
d2 = ± √ = ±0.2876 [d]
12.083
and we (arbitrarily) select the positive value for d2 . The normalized modal vectors are
   
1 0.3575
{u1 } = 0.3575 =
0.1267 0.0453
   
1 0.2876
{u2 } = 0.2876 = [e]
−2.104 −0.6053

Example 7.11
Consider the three-degrees-of-freedom mass-spring system in Figure 7.4 and obtain the
natural frequencies and modal vectors for the following parameters: W1 = W2 = 80 lb,
W3 = 92 lb, c1 = c2 = c3 = c4 = 0, k1 = k2 = 23 lb/in and k3 = k4 = 15 lb/in.
The damping matrix is zero. In order to have dimensional homogeneity, let us write the
masses in terms of slugs and spring constants in terms of lb/ft. We hence have
80 92
m1 = m2 = , m3 = slugs k1 = k2 = 276, k3 = k4 = 180 lb/ft [a]
32.17 32.17
The mass and stiffness matrices become
   
80 0 0 552 −276 0
1 
[m] = 0 80 0  slugs [k] =  −276 456 −180  lb/ft [b]
32.17
0 0 92 0 −180 360
320 Applied Dynamics

It is hard to solve the characteristic equation by hand, so use of a digital computer and
eigenvalue solver software is warranted.6
Solving the eigenvalue problem, the squares of the natural frequencies are obtained as
ω12 = 50.82 ω22 = 155.2 ω32 = 325.7 (rad/sec)2 [c]
and the corresponding normalized modal vectors are
     
0.2756 0.3492 0.4524
{u1 } =  0.4251  {u2 } =  0.2104  {u3 } =  −0.4214  [d]
0.3562 −0.4533 0.1330
The modal vectors are plotted in Figure 7.8. As expected, there are no sign changes in
{u1 }, one sign change in {u2 } and two in {u3 }.

{u1 } {u2 } {u3 }

0.4 0.4 0.4

1 2 3 1 2 3 1 2 3
– 0.4 – 0.4 – 0.4

FIGURE 7.8
Plots of normalized modal vectors. Notice the sign changes.

7.4.4 General Form of the Free Response


An n-degrees-of-freedom stable and conservative system has n natural frequencies. Hence,
the free response consists of superposition of n sinusoidal functions. Each sinusoidal has the
same form as the response expression that was discussed in the previous chapter. The free
response of an undamped multi-degrees-of-freedom system can be written as
n
X
{uj } C1j eiωj t + C2j e−iωj t

{x (t)} = (7.19)
j=1

Using the same arguments as in the previous chapter, we can show that C1j and C2j
are complex conjugates and the free response can be expressed as
{x (t)} = {u1 } A1 cos(ω1 t − φ1 ) + {u2 } A2 cos(ω2 t − φ2 )

+ . . . + {un } An cos (ωn t − φn ) (7.20)


where the amplitudes Aj and phase angles φj (j = 1, 2, . . . , n) depend on the initial condi-
tions. There are 2n initial conditions, expressed as two vectors of order n each, {x (0)} and
{ẋ (0)}. Evaluating the above equation and its derivative at t = 0 gives
{x (0)} = {u1 } A1 cos φ1 + {u2 } A2 cos φ2 + . . . + {un } An cos φn (7.21)
6 For example, the MATLABr command [U,V] = eig(K,M) gives the solution for the matrix eigenvalue

problem λ[M ]{u} = [K]{u}. The eigenvalues are returned as the diagonal elements of the matrix [V ] and
the eigenvectors are the columns of [U ], so that [U ] = [{u1 } {u2 } . . . {un }]. When [M ] is positive definite
and [K] is symmetric, the eigenvectors are returned in normalized form, in accordance with the second
convenience.
Response of Multi-Degrees-of-Freedom Systems 321

{ẋ (0)} = {u1 } A1 ω1 sin φ1 + {u2 } A2 ω2 sin φ2 + . . . + {un } An ωn sin φn (7.22)


The above column vector equations represent 2n scalar equations in terms of the 2n
unknowns Aj and φj . Because of the transcendental terms that are involved, Equations
(7.21) and (7.22) are not easy to solve for the amplitudes Aj and phase angles φj , especially
for problems with large degrees of freedom. The above equations are not commonly used
for the general free response, except for special cases with two degrees of freedom. Rather,
it is preferable to employ techniques based on modal analysis to obtain the response.

7.5 Solving for the Natural Frequencies and Modal Vectors


There are different approaches and points of concern when analyzing and calculating the
natural frequencies and modal vectors. We identify two below:
1. Parametric analysis, where we analyze the problem in terms of its parameters, such
as the elements of the mass and stiffness matrices or the ratios of these elements. The
characteristic equation is solved in terms of these parameters. The natural frequencies
or modal vectors are expressed in terms of these parameters. Conversely, we can solve
for the system parameters in terms of known (or desired) frequency values or in terms
of desired motion amplitudes.
2. Modal analysis, where we develop a general procedure to solve for the roots of the
characteristic equation, the modal vectors and the response. Modal analysis uses con-
cepts from linear algebra and eigenvalue analysis. Modal analysis readily lends itself
to computer implementation and thus is suitable for large-degrees-of-freedom systems.
There exist a vast number of computational techniques to solve matrix eigenvalue prob-
lems.
Some argue that with the proliferation of computers and advanced computing tech-
niques we should forgo hand calculations, even for two-degrees-of-freedom systems. The
reason parametric studies and hand calculations are still a significant part of analysis is the
tremendous amount of insight they bring into a problem. Simple back of the envelope calcu-
lations can enhance our understanding of a dynamical system, and they provide a starting
point for more detailed analysis.
We will study parametric analysis within the context of examples and special cases. The
reader is also urged to study vehicle suspension systems in Chapter 15 for some interesting
applications of parametric analysis, such as calculating spring constants in terms of desired
suspension frequencies.

Example 7.12
Given the mass-spring system in Figure 7.9 with m1 = m, m2 = m, calculate the value of
ω2
k2 in terms of k1 such that the relationship ω1
= 3 is satisfied.
The free-body diagrams are given in Figure 7.10. Using the approach for mass-spring
systems in series, we can write the mass and stiffness matrices directly as
   
m 0 k1 + k2 −k2
[m] = [k] = [a]
0 m −k2 k2
so that the characteristic equation becomes
−ω 2 m + k1 + k2
 
−k2
det −ω 2 [m] + [k] = det

= 0 [b]
−k2 −ω 2 m + k2
322 Applied Dynamics

"! ""
!! !"

#! #"

FIGURE 7.9
Mass-spring system.

k2(x2 – x1 )
k1x1 m1 m2

FIGURE 7.10
Free-body diagrams.

Dividing each row of the above equation by k1 and introducing the ratio k2 = ek1 and
2
the term β = ωk1m , the characteristic equation can be expressed as
 
2 −β + 1 + e −e
= β 2 − (2e + 1) β + e = 0

det −ω [m] + [k] = det [c]
−e −β + e

Using the quadratic formula, solution of the characteristic equation becomes

ω2 m 1 p 
β = = (2e + 1) ± 4e2 + 1 [d]
k 2
and the ratio of the two frequencies thus is

ω22 β2 (2e + 1) + 4e2 + 1
= = √ [e]
ω12 β1 (2e + 1) − 4e2 + 1

Equation [e] can be solved for the desired frequency ratio. Figure 7.11 plots the frequency
ratio ω ω2
ω1 as a function of the ratio e. The desired frequency ratio of ω1 = 3 is obtained when
2

e = 1.6382.

7.6 Beat Phenomenon


Beat phenomenon occurs when two of the natural frequencies of a system are very close
to each other.7 A two-degrees-of-freedom system that exhibits beat phenomenon can be
constructed by coupling two similar single-degree-of-freedom systems with a weak spring.
7 Beat phenomenon is also observed in one-degree-of-freedom problems subjected to harmonic excitation

where the natural frequency and the excitation frequency are close to each other.
Response of Multi-Degrees-of-Freedom Systems 323

Natural frequency ratio


3.5

2.5

2
0 0.5 1 1.5 2 2.5 3
Ratio of spring constants e

FIGURE 7.11
Natural frequency ratio ω2 /ω1 as a function of the stiffness ratio e = k2 /k1 .

O1 O2 g

m, L
1 2

m, L
k

FIGURE 7.12
Two pendulums connected by a weak spring.

Consider the two identical rod pendulums in Figure 7.12 connected to each other by a
spring. Each rod is of mass m and length L. The motion of the pendulums is assumed to be
small, so that the spring stretches only in the horizontal direction. Figure 7.13 shows the
free-body diagrams. Summing moments about the pivot point of each pendulum gives

X L
Left: MO1 = IO θ̈1 = −mg sin θ1 + k (L sin θ2 − L sin θ1 ) L cos θ1
2

X L
Right: MO2 = IO θ̈2 = −mg sin θ2 − k (L sin θ2 − L sin θ1 ) L cos θ2 (7.23)
2
1
where IO = 3 mL2 . The equations of motion are

L
IO θ̈1 + mg sin θ1 + kL2 cos θ1 (sin θ1 − sin θ2 ) = 0
2

L
IO θ̈2 + mg sin θ2 + kL2 cos θ2 (sin θ2 − sin θ1 ) = 0 (7.24)
2
324 Applied Dynamics

O1v O2v
O1h O2h

G1 G2
1 2

mg mg

k(Lsin 2 – Lsin 1 )

FIGURE 7.13
Free-body diagrams.

For small motions about the vertical positions of the pendulums, using a small angles
assumption, the linearized equations of motion become
 
1 L
mL2 θ̈1 + mg + kL2 θ1 − kL2 θ2 = 0
3 2
 
1 L
mL2 θ̈2 − kL2 θ1 + mg + kL2 θ2 = 0 (7.25)
3 2

Introducing the position vector {x} = [θ1 θ2 ]T the equations of motion can be written
in matrix form, where the coefficient matrices are
 1 2
mg L2 + kL2 −kL2
  
3 mL 0
[m] = 1 [k] = (7.26)
0 3 mL
2
−kL2 mg L2 + kL2
The next step is to solve for the natural frequencies and mode shapes. The characteristic
equation becomes

det −ω 2 [m] + [k] = 0




− 13 ω 2 mL2 + mg L2 + kL2 −kL2


 
= det = 0 (7.27)
−kL2 − 3 ω mL + mg L2 + kL2
1 2 2

It is convenient to divide both rows of the above matrix by mgL/2 and to introduce the
quantities
1 2 2
3 mL ω 2L 2 kL2 2kL
β = −1 = ω −1  = = (7.28)
mgL/2 3g mgL/2 mg
where  is recognized as the ratio of the moment exerted by the spring force versus the
moment generated by the gravitational force. A small value for  indicates that the con-
tribution of the spring force is less than that of the gravitational force. The determinant
equation can now be written as
−β +  −
 
2
det = (−β + ) − 2 = β 2 − 2β = 0 (7.29)
− −β + 
Response of Multi-Degrees-of-Freedom Systems 325

whose solution is

β = 0, 2 (7.30)

so that the natural frequencies are


3g 3g
β1 = 0 =⇒ ω12 = β2 = 2 =⇒ ω22 = (1 + 2) (7.31)
2L 2L
Note that ω12 has the same value as the square of the natural frequency of each rod when
the spring between the rods is removed. We will see why shortly.
To find the modal vectors, it is preferable to use the form of the coefficient matrix in
Equation (7.29), which gives
    
−βj +  − u1j 0
= j = 1, 2 (7.32)
− −βj +  u2j 0

For j = 1, β1 = 0 and
        
 − u11 0 u11 1
= =⇒ {u1 } = = d1 (7.33)
−  u21 0 u21 1

The rods move in unison for the first mode with each rod having the same amplitude.
Hence, the spring between them is not stretched. This explains why the first natural fre-
quency of the two-degrees-of-freedom system is the same as the natural frequency of each
individual rod.
For the second mode, β2 = 2, resulting in
        
− − u12 0 u12 1
= =⇒ {u2 } = = d2 (7.34)
− − u22 0 u22 −1

In the second mode, the rods again move with the same amplitude, but in different direc-
tions. The modal vectors of the system are depicted in Figure 7.14.

a) b)
O1 O2 O1 O2

1 1 2 2

FIGURE 7.14
Mode shapes of coupled pendulums: a) {u1 }, b) {u2 }.

The free response can be written using Equation (7.22) as


     
θ1 (t) 1 1
{x (t)} = = A1 cos (ω1 t − φ1 ) + A2 cos (ω2 t − φ2 ) (7.35)
θ2 (t) 1 −1

where the constants d1 and d2 have been absorbed into A1 and A2 , indicating once more
326 Applied Dynamics

that the normalization constants do not affect the final result. Without loss of generality,
we can select the initial conditions as θ1 (0) = θ0 , θ2 (0) = 0, θ̇1 (0) = θ̇2 (0) = 0, which
result in
θ0
A1 = A2 = φ1 = φ2 = 0 (7.36)
2
so the free response can be written as
1 1 1 1
θ1 (t) = θ0 cos ω1 t + θ0 cos ω2 t θ2 (t) = θ0 cos ω1 t − θ0 cos ω2 t (7.37)
2 2 2 2
Up to this point, the analysis was for the general case and it did not consider any
special values of the spring constant. Now, consider that the moment generated by the
spring is much smaller than the gravitational moment. This implies that the parameter
 = 2kL
mg << 1.
The first natural frequency is not affected by the spring. The second natural frequency,
however, is affected by the spring and for a weak spring, it can be approximated as
q
1/2
ω2 = ω12 (1 + 2) = ω1 (1 + 2) ≈ ω1 (1 + ) (7.38)

We next define the average frequency ωa and beat frequency ωb as


ω2 + ω1   ω2 − ω1 1
ωa = = ω1 1 + ≈ ω1 ωb = = ω1  (7.39)
2 2 2 2
ωb
The average frequency is much higher than the beat frequency as ωa ≈ 2 . Also,

ωa + ωb = ω2 ωa − ωb = ω1 (7.40)

Using the trigonometric identity cos (c ± d) = cos c cos d∓sin c sin d, Equation (7.37) can
be rewritten as

θ1 (t) = θ0 cos ωb t cos ωa t θ2 (t) = θ0 sin ωb t sin ωa t (7.41)

or

θ1 (t) = θ0 cos ω1 t cos ω1 t θ2 (t) = θ0 sin ω1 t sin ω1 t (7.42)

The response is the product of two sinusoids. One of the sinusoids has a frequency
of ωa ≈ ω1 , and the second has a frequency of ωb , which is much smaller than ωa . The
slowly varying sinusoidal creates an amplitude envelope for the faster varying sinusoidal.8
The response, plotted in Figures 7.15 and 7.16, shows the modulation for the parameters
θ0 = 1, ωa = 1 rad/s, ωb = 0.1 rad/s. The period of beating is half the period of the
envelope, as
2π 2π
Tbeating = ≈ (7.43)
ωb ω1

and, for the parameters considered above, the beating period becomes 0.1 = 62.83 seconds,
while the period of the envelope is twice that, at 125.7 seconds.
The phase angle between θ1 (t) and θ2 (t) is 90◦ , so when θ1 (t) has its maximum value,
θ2 (t) = 0 and vice versa. Hence, energy gets transferred from one pendulum to the other
through the connecting spring every half period.
8 This is analogous to the description of damped free motion, where an exponential decay curve envelopes

a sinusoid.
Response of Multi-Degrees-of-Freedom Systems 327

0.5
θ (t)
0
1

−0.5

−1
0 20 40 60 80 100 120 140 160 180 200
Time (sec)

FIGURE 7.15
Response of θ1 (t), with θ0 = 1, ωa = 1 rad/s, ωb = 0.1 rad/s.

0.5
θ (t)

0
2

−0.5

−1
0 20 40 60 80 100 120 140 160 180 200
Time (sec)

FIGURE 7.16
Response of θ2 (t), with θ0 = 1, ωa = 1 rad/s, ωb = 0.1 rad/s.

7.7 Unrestrained Motion and Rigid Body Modes


The examples in this chapter have so far dealt with systems that have at least one component
connected to a fixed support by a spring or damper. This section considers systems that
are not attached to a fixed point. Such systems do not oscillate about a fixed point and are
free to translate or to rotate. Figure 7.17 shows examples of unrestrained systems.
Consider, for example, the mass-spring system in Figure 7.9. Removing the spring con-
necting the first mass (m1 ) to the fixed wall, the resulting configuration is shown in Figure
7.17a. Similarly, removing the spring in the cart-rod system in Figure 7.1 results in Fig-
ure 7.17b. If a horizontal force is applied to these systems, in the absence of friction or
other energy dissipation mechanism, the masses will keep on moving. There are no forces
constraining the masses to come back to their original positions. Such systems are called
unrestrained.
At least one of the natural frequencies of an unrestrained system is zero, and the corre-
sponding modal vector indicates that the components do not move relative to each other.
This type of motion is called a rigid body mode. The system moves as one piece, as if it
were a rigid body.
328 Applied Dynamics
#$ %$
'
$"
% %
#! #" & "

!! !" ! #&'"(

FIGURE 7.17
Examples of unrestrained systems.

A zero eigenvalue is indication of a stiffness matrix that is positive semi-definite. In


all examples up to now, the stiffness matrix was positive definite. A positive semi-definite
stiffness matrix is always associated with a rigid body mode with zero natural frequency.
It is possible for a system to have a rigid body mode with nonzero natural frequency.
In this mode, the system moves and oscillates as a whole, but the system is restrained.
For example, in the previous section on beat phenomenon, the two pendulums have a rigid
body mode with nonzero frequency. As given in Equation (7.33), the first modal vector is
{u1 } = d1 [1 1]T and it describes a rigid body mode, where the two pendulums swing with
the same amplitude. Table 7.1 summarizes the relationship between rigid body modes and
zero natural frequencies.

TABLE 7.1
Laws governing the existence of zero frequencies and rigid body modes

If there is there
a zero frequency, also is an associated a rigid body mode.
no zero frequency, may be a rigid body mode.
a rigid body mode, may be an associated zero frequency.

For an unrestrained n-degrees-of-freedom system, with ω1 = 0, we can reduce the fre-


quency equation to one in terms of n − 1 frequencies and solve for the natural frequencies.
This approach is useful for three-degrees-of-freedom systems, as it results in frequency equa-
tion of order two, which can be solved by hand.

Example 7.13
Consider the mass-spring system in Figure 7.9 (or Figure 7.17a) with m1 = 2m, m2 =
m, k1 = 0, k2 = k and find its eigensolution.
The mass and stiffness matrices are
   
2m 0 k −k
[m] = [k] = [a]
0 m −k k

The stiffness matrix is positive semi-definite because its determinant is zero. The char-
Response of Multi-Degrees-of-Freedom Systems 329

acteristic determinant becomes


k − 2mω 2
 
−k
det −ω 2 [m] + [k]

= 0 =⇒ det = 0 [b]
−k k − mω 2
2
Introducing the quantity β = mωk and dividing both rows by k gives
 
1 − 2β −1
det = 2β 2 − 3β = 0 [c]
−1 1−β
whose solution is
r
3k
β1 = 0, β2 = 1.5 =⇒ ω1 = 0, ω2 = [d]
2m
The modal vector associated with the rigid body mode satisfies the relationship
    
 2  k −k u11 0
−ω1 [m] + [k] {u1 } = {0} =⇒ = [e]
−k k u21 0
As expected, the coefficient matrix is singular. The top equation gives ku11 − ku21 = 0,
from which we conclude that the associated modal vector is
 
1
{u1 } = d1 [f ]
1
verifying that the two masses move with the same amplitude. Neither mass moves relative
to the other. The second mode is found from
    
 2  −2k −k u12 0
−ω2 [m] + [k] {u2 } = {0} =⇒ = [g]
−k − k2 u22 0

which, when solved, gives the second modal vector as


 
1
{u2 } = d2 [h]
−2
The second mode has a sign change, indicating that the masses move against each other.
The entire motion can be described as the summation of two modes: in the first (rigid body)
mode, the masses move together with the same amplitude and no oscillation; and in the
second mode, the masses oscillate and move against each other.

Example 7.14
Consider the mass-pendulum system in Figure 7.1 and the case when the spring and damper
connecting the cart to a fixed wall are removed. Find the natural frequencies and the modal
vector associated with the rigid body mode.
The system is shown in Figure 7.17b. The mass matrix is the same as before. Setting
k = 0, the coefficient matrices become
     
7 3 k 0 0 0
[m] = [k] = = [a]
3 4 0 mgL 0 30
As expected, the stiffness matrix is positive semi-definite, evidenced by the zero diagonal
element. The characteristic equation becomes
−7ω 2 −3ω 2
 
det −ω 2 [m] + [k] = det

= 0 [b]
−3ω 2 30 − 4ω 2
330 Applied Dynamics

which, when expanded, gives the characteristic equation

19ω 4 − 210ω 2 = 0 [c]

The solution of the characteristic equation is


210
ω12 = 0 ω22 = [d]
19
The modal vector associated with the rigid body mode is calculated using
    
 2  0 0 u11 0
−ω1 [m] + [k] {u1 } = {0} =⇒ = [e]
0 30 u21 0
The top equation is inconclusive. The bottom equation gives 30u21 = 0, whose solution
is u21 = 0. Hence, the modal vector associated with the rigid body mode is
 
1
{u1 } = d1 [f ]
0

Examining the modal vector {u1 }, the element u21 , which describes the amplitude of
the rod with respect to the mass, is zero. In the rigid body mode, the cart moves but the
rod does not move with respect to the cart. Consequently, the entire the system moves as
one.

Example 7.15
Consider the two systems in Figure 7.18 and write their rigid body modal vectors.

a) b)
k1 k2 g
k
m1 m2 m3 m1 m2

x1 x2 x3 m 3, L 3 m 4, L 4
x1 x2
1 2

FIGURE 7.18
Two systems exhibiting rigid body motion.

For the three mass-spring system, the position vector in terms of the displacements of
the masses is {x} = [x1 x2 x3 ]T . In the rigid body mode, all masses move with the same
amplitude, so the modal vector associated with rigid body mode is
 
{u1 } = d1 1 1 1 [a]

For the two carts about which the pendulums pivot, the position vector can be written as
{x} = [x1 θ1 x2 θ2 ]T . In the rigid body mode, the two carts move with the same amplitude,
and the rods do not move with respect to the carts. Hence, the modal vector associated
with the rigid body mode is
 
{u1 } = d1 1 0 1 0 [b]
Response of Multi-Degrees-of-Freedom Systems 331

7.8 Orthogonality of the Modal Vectors


The modal vectors possess a very important property called orthogonality. Similar to the
same concept in Fourier series or in Cartesian vectors, orthogonality provides physical in-
sight, and it allows us to break down the system differential equations into a set of inde-
pendent equations.
Consider the case where all the eigenvalues are distinct, that is, ωj 6= ωs (j, s =
1, 2, . . . , n). The orthogonality property of the eigenvectors states that
T T
{uj } [m] {us } = 0 {uj } [k] {us } = 0 when j 6= s j, s = 1, 2, . . . , n (7.44)

To show the existence of orthogonality, write the eigenvalue problem for the j-th and
s-th modes as

[k] {uj } = ωj2 [m] {uj } [k] {us } = ωs2 [m] {us } (7.45)

Next, left multiply the first of the above equations by {us }T and the second by {uj }T and
subtract the first equation from the second. Because the matrices [m] and [k] are symmetric,
the relationships
T T T T
{uj } [m] {us } = {us } [m] {uj } {uj } [k] {us } = {us } [k] {uj } (7.46)

hold. Hence, the subtraction yields


T T
{uj } [k] {us } = ωs2 {uj } [m] {us }
T T (7.47)
− {us } [k] {uj } = − ωj2 {us } [m] {uj }

−−−−−−−−−−−−−−−−−−−−−−−−−
 T
0 = ωs2 − ωj2 {uj } [m] {us }
Because the natural frequencies are distinct, the only way the above relationship can
hold is if {uj }T [m] {us } = 0 when j 6= s. In a similar fashion, we can prove the orthogonality
relationship with respect to the stiffness matrix. Proof of orthogonality when the eigenvalues
are not distinct can be found in more advanced texts.
The next step is to combine the normalization process and orthogonality. The term δjs
is referred to as the Kronecker delta and is defined as
(
1 when j = s
δjs = (7.48)
0 when j 6= s

Recall that by using the second convenience, the modal vectors are normalized by

{uj }T [m] {uj } = 1, j = 1, 2, . . . , n (7.49)

Combining this relationship with Equation (7.44), we can write

{uj }T [m] {us } = δjs {uj }T [k] {us } = ωj2 δjs j, s = 1, 2, . . . , n (7.50)

which are known as the orthonormality relationships. The normalized eigenvectors are re-
ferred to as orthonormal.
The orthogonality relationships can also be written in matrix form. To this end, introduce
332 Applied Dynamics

the modal matrix [U ] of order n × n, which is obtained by placing all of the eigenvectors
side by side

[U ] = [{u1 } {u2 } . . . {un }] (7.51)

We can then write the orthogonality relationships as

[U ]T [m] [U ] = [1] [U ]T [k] [U ] =


 2
ω (7.52)
 
in which ω 2 is a diagonal matrix whose elements are ωj2 (j = 1, 2, . . . , n).

Example 7.16
Consider the mass-pendulum problem, motion about stable equilibrium, and show that the
modal vectors are orthogonal.
The mass and stiffness matrices are
   
7 3 8 0
[m] = [k] = [a]
3 4 0 30

and the modal vectors were shown to be


   
1 1
{u1 } = d1 {u1 } = d2 [b]
0.1267 −2.104

For this example, we do not need to write the modal vectors in normalized form and we
arbitrarily set d1 = d2 = 1. The orthogonality condition with respect to the mass matrix
becomes
 T   
T T 1 7 3 1
{u1 } [m] {u2 } = {u2 } [m] {u1 } =
0.1267 3 4 −2.104
 T  
1 7 − 3 × 2.104
= = 7 − 3 × 2.104 + 0.1267 × (3 − 4 × 2.104) = 0 [c]
0.1267 3 − 4 × 2.104
The orthogonality condition with respect to the stiffness matrix is
 T   
1 8 0 1
{u1 }T [k] {u2 } = {u2 }T [k] {u1 } =
0.1267 0 30 −2.104
 T  
1 8
= = 8 − 0.1267 × 30 × 2.104 = 0 [d]
0.1267 −30 × 2.104

7.9 Expansion Theorem


The expansion theorem, also known as the decomposition theorem, is a useful mathematical
tool that enables us to expand a vector (or function) in terms of a series of known vectors
(functions). The theorem makes use of the property that the modal vectors constitute an
independent and orthogonal set.
This section begins by considering sets of functions. The distinction between complete,
independent, and orthogonal sets is highlighted next.
Response of Multi-Degrees-of-Freedom Systems 333

7.9.1 Expansion by Sets of Functions


Consider a set of one-dimensional functions f1 (x) , f2 (x) , . . . in a domain a ≤ x ≤ b. This
set is said to be complete if any one-dimensional piecewise-continuous function g (x) in the
same domain can be represented as a linear combination of the set of functions in the form

X
g (x) = aj fj (x) (7.53)
j=1

in which aj (j = 1, 2, . . . ) are a set of coefficients to be determined. Often, the number of


terms is truncated to a finite value n for an approximate solution
n
X
g (x) ≈ aj fj (x) (7.54)
j=1

A set of functions fj (x) is called an independent set if any of the individual functions
in the set fk (x) (k = 1, 2, . . . ) cannot be expressed in the form

X
fk (x) 6= aj fj (x) k = 1, 2, . . . (7.55)
j=1
j6=k

A set of functions fj (x) is called orthogonal with respect to the weighting function
W (x) in the interval a ≤ x ≤ b if the following relationships hold:
Z b
W (x) fj (x) fk (x) dx = 0 j, k = 1, 2, . . . ; j 6= k (7.56)
a

It follows that an infinite set of orthogonal functions is independent and complete.


When expanding a function by Equation (7.53), it is important to be able to determine
the coefficients of expansion aj . For example, the set of simple polynomials 1, x, x2 , x3 , . . .
is a complete and independent set but it is not an orthogonal set. When expanding a function
by simple polynomials, we need to create a set of rules to solve for the coefficients aj . There
are several rules and combinations available, such as Legendre polynomials.
By contrast, when dealing with orthogonal functions, such as Fourier series, the coeffi-
cients of the expansion can be obtained readily, by making use of the orthogonality property.
Given a periodic function g (x) with period T , so that

g (x + T ) = g (x) (7.57)

we can expand that function by the Fourier series


∞     
1 X 2πj 2πj
g (x) = a0 + aj cos x + bj sin x (7.58)
2 j=1
T T

The coefficients aj and bj are ascertained by invoking the orthogonality properties of


the sine and cosine functions, which are
Z T /2    
2πj 2πk
sin x cos x dx = 0 j, k = 0, 1, 2, . . . (7.59)
−T /2 T T

Z T /2    
2 2πj 2πk
sin x sin x dx
T −T /2 T T
334 Applied Dynamics
Z T /2    
2 2πj 2πk
= cos x cos x dx = δjk (7.60)
T −T /2 T T
2πk

For example, multiplying the expansion in Equation (7.58) by cos T x , (k = 1, 2, . . . )
and integrating over the period T gives
Z T /2  
2πk
g (x) cos x dx
−T /2 T
 
Z T /2 ∞       
1 X 2πj 2πj 2πk
=  a0 + aj cos x + bj sin x  cos x dx (7.61)
−T /2 2 j=1
T T T

By using the orthogonality relationships above, we can show that only one term survives
from the above expansion, which is
Z T /2  
2πk T
g (x) cos x dx = ak k = 1, 2, . . . (7.62)
−T /2 T 2
so the coefficients of the cosine terms become
2 T /2
Z  
2πj
aj = g (x) cos x dx j = 1, 2, . . . (7.63)
T −T /2 T
In a similar way, we can show the other coefficients to be
2 T /2 2 T /2
Z Z  
2πj
a0 = g (x) dx bj = g (x) sin x dx (7.64)
T −T /2 T −T /2 T
with a0 denoting the average value of the function g (x).

7.9.2 Expansion by Geometric Vectors


The expansion of a vector in terms of a different set of vectors was discussed in Chapter 2.
For example, given a geometric vector, say, r = 3I − 4J in an XY Z coordinate system, let
us express it in terms of an xyz coordinate system, which is obtained by rotating the XY Z
coordinates by an angle 30◦ counterclockwise about the Z axis, as shown in Figure 7.19.
From Figure 7.19, the unit vectors associated with the rotated coordinate system are
related to the initial coordinates XY by
i = cos 30◦ I + sin 30◦ J j = − sin 30◦ I + cos 30◦ J (7.65)
The components of r in terms of the xyz frame can be written as r = rx i + ry j, where
rx = r · i ry = r · j (7.66)
Introduction of the values for the unit vectors from Equation (7.65) gives
rx = r · i = (3I − 4J) · (cos 30◦ I + sin 30◦ J) = 3 cos 30◦ − 4 sin 30◦ = 0.5981

ry = r · j = (3I − 4J) · (− sin 30◦ I + cos 30◦ J) = −4.9641 (7.67)


We can, of course, use the inverse relationship between the unit vectors
I = cos 30◦ i − sin 30◦ j J = sin 30◦ i + cos 30◦ j (7.68)
and substitute these values into the expression for r, r = 3I − 4J, with the same result.
Response of Multi-Degrees-of-Freedom Systems 335

Y
y
30° x

30° X
3

–4
(3, – 4)

FIGURE 7.19
Vector r in two coordinate systems.

7.9.3 Expansion by Modal Vectors


This section extends the previous discussion to modal vectors by noting that the modal
vectors constitute an orthogonal and complete set. Hence, any vector {v} of dimension n
can be expressed (decomposed) as a uniformly convergent series of the modal vectors in the
form

{v} = α1 {u1 } + α2 {u2 } + . . . + αn {un } (7.69)

The coefficients αj (j = 1, 2, . . . , n) are calculated by making use of the orthogonality


property with respect to the mass matrix. Let us left multiply the above equation with
{us }T [m] (s = 1, 2, . . . , n). For example, for s = 3, left multiplying with {u3 }T [m] yields
T T T
{u3 } [m] {v} = α1 {u3 } [m] {u1 } + α2 {u3 } [m] {u2 }

T T
+ α3 {u3 } [m] {u3 } + . . . + αn {u3 } [m] {un } (7.70)

Of all the terms on the right side, only one, {u3 }T [m] {u3 }, is equal to unity and the rest
T
vanish due to the orthogonality of the modal vectors. Hence, {u3 } [m] {v} = α3 . Extending
T
this to every modal vector, we obtain αj = {uj } [m] {v} (j = 1, 2, . . . , n). The expansion
theorem for modal vectors becomes
n
X
{v} = αj {uj } αj = {uj }T [m] {v} (7.71)
j=1

We can express the expansion theorem in matrix form by making use of the modal
matrix [U ] (Equation (7.51)) as
T
{v} = [U ]{α} {α} = [U ]−1 {v} = [U ] [m] {v} (7.72)
T
in which the vector {α} = [α1 α2 . . . αn ] contains the coefficients of the expansion.
The expansion theorem has several applications in mathematics, physics, and engineer-
ing. The interest here is obtaining the response of the dynamical system. To this end, the
next section will seek an expansion of the position vector within the context of the equations
of motion.
336 Applied Dynamics

Example 7.17
Return to the cart-rod problem, consider the linearized equations of motion about stable
equilibrium and expand the vectors {v} = [1 1]T and {w} = [1 −2]T in terms of the modal
vectors.
The expansion theorem states
{v} = α1 {u1 } + α2 {u2 } [a]
The coefficients α1 and α2 are calculated using the second part of the expansion theorem
as
T T
α1 = {u1 } [m] {v} α2 = {u2 } [m] {v} [b]
From Example 7.10, the mass matrix and normalized modal vectors are
     
7 3 0.3575 0.2876
[m] = {u1 } = {u2 } = [c]
3 4 0.0453 −0.6053
so the coefficients of the expansion of {v} become
 T   
T 0.3575 7 3 1
α1 = {u1 } [m] {v} = = 3.892 [d]
0.0453 3 4 1
 T   
T 0.2876 7 3 1
α2 = {u2 } [m] {v} = = −1.361 [e]
−0.6053 3 4 1
One can express the vector {v} as
     
1 0.3575 0.2876
{v} = = 3.892 − 1.361 [f ]
1 0.0453 −0.6053
or      
1 1 1
{v} = = 1.3916 − 0.3916 [g]
1 0.1267 −2.104
The coefficient of the first modal vector, α1 , is higher than its counterpart for the second
modal vector, α2 . This is because the vector [1 1]T is closer to the first modal vector than the
second. Hence, the coefficient of the first term in the expansion is higher than the coefficient
of the second term.
Next, consider the vector {w} = [1 −2]T . This vector resembles the second modal vector
closely, so the coefficient of the second mode should be higher than the first. Indeed, the
counterparts of Equations [d] and [e] become
 T   
T 0.3575 7 3 1
α1 = {u1 } [m] {w} = = 0.1310 [h]
0.0453 3 4 −2
 T   
T 0.2876 7 3 1
α2 = {u2 } [m] {w} = = 3.314 [i]
−0.6053 3 4 −2

7.10 Modal Equations of Motion and Response


This section applies the expansion theorem to the equations of motion and demonstrates
that by doing so, we can decouple the equations of motion and put them in a form whose
solution can be obtained using methods described in Chapter 6.
Response of Multi-Degrees-of-Freedom Systems 337

Consider the application of the expansion theorem to the displacement vector {x (t)}.
From Equation (7.69), because the eigenvectors are not functions of time, the coefficients
of the expansion must be functions of time. Denoting these coefficients by qj (t) (j =
1, 2, . . . , n), and calling them modal coordinates, the first part of the expansion theorem
becomes
n
X
{x (t)} = {uj } qj (t) (7.73)
j=1

From the second part of the expansion theorem, the modal coordinates are expressed as

qj (t) = {uj }T [m] {x (t)} (7.74)

Introduction of Equation (7.73) to the equations of motion, [m] {ẍ (t)} + [k] {x (t)} =
{F (t)}, gives
n
X n
X
[m] {uj } q̈j (t) + [k] {uj } qj (t) = {F (t)} (7.75)
j=1 j=1

The next step is to left multiply the above equation by {us }T (s = 1, 2, . . . , n) and rearrange
the terms by moving the summation to the left. For example, for s = 3, the resulting
expression is
n
X n
X
q̈j (t) {u3 }T [m] {uj } + qj (t) {u3 }T [k] {uj } = {u3 }T {F (t)} (7.76)
j=1 j=1

The term on the right side of Equation (7.76) is called modal force. Denoted by Qs (t),
the modal force is defined as

Qs (t) = {us }T {F (t)} s = 1, 2, . . . , n (7.77)

The terms on the left side of Equation (7.76) can be evaluated by invoking the orthog-
onality relations. In each summation one term survives: the one corresponding to j = 3.
Equation (7.76) thus becomes

q̈3 (t) + ω32 q3 (t) = Q3 (t) (7.78)

Generalizing this result to each modal coordinate, the equations of motion can be written
as n independent equations, known as modal equations or modal equations of motion, as

q̈j (t) + ωj2 qj (t) = Qj (t) j = 1, 2, . . . , n (7.79)

7.10.1 Modal Response


The original equations of motion, which were in terms of the coordinates x1 , x2 , . . . , xn
and in the form of n coupled second-order differential equations, have been converted into
n uncoupled, independent equations. These independent equations can be solved separately
from each other.9 Recall that in the previous chapter we developed expressions for the
response of single-degree-of-freedom systems.
Each modal equation can be treated as a single-degree-of-freedom problem independent
9 This, of course, is possible as long as the j-th modal force Q (t) is not a function of the modal coordinates
j
and modal velocities other than qj (t) and q̇j (t), that is, Qj (t) = Qj (qj , q̇j , t) only.
338 Applied Dynamics

of the other modal equations. From Equation (6.146), the solution for each modal coordinate
can be expressed as
Z t
q̇j (0) 1
qj (t) = qj (0) cos ωj t + sin ωj t + Qj (t − τ ) sin ωj τ dτ j = 1, 2, . . . , n (7.80)
ωj ωj 0

with qj (0) and q̇j (0) denoting the initial conditions associated with the modal coordinates.
These initial conditions can be obtained from the initial displacement and velocity by using
the expansion theorem, Equation (7.74), as
T T
qj (0) = {uj } [m] {x (0)} q̇j (0) = {uj } [m] {ẋ (0)} (7.81)

The total response is obtained by adding (or superposing) the individual responses by
n
X
{x (t)} = {uj } qj (t) (7.82)
j=1

and by introducing the modal responses in Equation (7.80) into the above equation.
The modal equations of motion can also be derived in matrix form. The expansion
theorem in matrix form is

{x (t)} = [U ] {q (t)} (7.83)

where {q (t)} is called the modal vector and has the form {q (t)} = [q1 (t) q2 (t) . . . qn (t)].T
In essence, the above equation is a coordinate transformation, such as the ones we saw in
Chapter 2 when studying kinematics. Introducing this expansion into the equation of motion
and left multiplying by [U ]T gives

[U ]T [m] [U ] {q̈ (t)} + [U ]T [k] [U ] {q (t)} = [U ]T {F (t)} (7.84)

Defining the modal force vector by

{Q (t)} = [Q1 (t) Q2 (t) . . . Qn (t)]T = [U ]T {F (t)} (7.85)

and using the matrix orthogonality relationships in Equation (7.52) leads to the modal
equations in matrix form as

{q̈ (t)} + ω 2 {q (t)} = {Q (t)}


 
(7.86)

where [ω 2 ] is a diagonal matrix whose elements are the squares of the natural frequencies,
ωj2 (j = 1, 2, . . . , n).
The modal matrix makes it possible to decouple the equations of motion. If we use an
arbitrary transformation matrix, say, [A], instead of the modal matrix [U ], the resulting
matrix products [A]T [m] [A] and [A]T [k] [A] in Equation (7.84) will be symmetric but the
equations of motion will not be uncoupled.
There are two distinct advantages of the modal analysis approach for finding the re-
sponse. First, modal analysis makes it possible to take a series of coupled differential equa-
tions and rewrite them as independent equations that can be solved individually. Second,
modal analysis readily lends itself to computer implementation, which makes it possible to
obtain the response of a system with several degrees of freedom. Indeed, as discussed earlier,
powerful software exists to solve the eigenvalue problem, to normalize modal vectors, and
to obtain the response.
Following is the procedure for obtaining the response of a multi-degrees-of-freedom un-
damped system:
Response of Multi-Degrees-of-Freedom Systems 339

1. Obtain the equations of motion, find the equilibrium positions, and (if necessary) lin-
earize around equilibrium or another operating position. This yields the matrix equations
of motion [m] {ẍ (t)} + [k] {x (t)} = {F (t)}.
2. Solve the characteristic equation, obtain the natural frequencies ωj and modal vectors
{uj } (j = 1, 2, . . . , n), and normalize the modal vectors.
3. Calculate the initial modal coordinates and velocities, as well as modal forces, using
qj (0) = {uj }T [m] {x (0)}, q̇j (0) = {uj }T [m] {ẋ (0)}, Qj (t) = {uj }T {F (t)}.
4. Solve the modal equations of motion q̈j (t) + ωj2 qj (t) = Qj (t) , (j = 1, 2, . . . , n) for the
modal coordinates qj (t). This gives the response of each mode qj (t).
5. Using the first
Ppart of the expansion theorem, calculate the response of the entire system
n
as {x (t)} = j=1 {uj } qj (t).

7.10.2 Modal Response Using Homogeneous and Particular Solution


Approach
Chapter 6 discussed two approaches for obtaining the response of single-degree-of-freedom
systems:

• In the homogeneous plus particular solution approach, we first obtain the homogeneous
solution xH (t). This solution is in terms of undetermined coefficients. The particular
solution xP (t) is obtained next. The two solutions are added, the initial conditions are
invoked, and the undetermined coefficients are obtained in terms of the initial conditions.
• In the Laplace transform approach, terms that contain coefficients of the Laplace trans-
form X (s) of the motion variable x (t) are collected and solved for X (s). The inverse
Laplace transform of X (s) gives the response, as derived in Equation (6.146). The
Laplace transform solution is identical to the convolution integral approach and the
modal response in Equation (7.80) is based on it.

This section obtains the modal response using homogeneous and particular solution
approach. From the developments of Section 7.4.4, the homogeneous solution for the j-th
mode is

qjH (t) = Aj cos (ωj t − φj ) (7.87)

where the amplitudes Aj and phase angles φj (j = 1, 2, . . . , n) associated with each mode
are not yet known. Given the modal force Qj (t) = {uj }T {F (t)}, the particular solution
qjP (t) is obtained using the approaches discussed in Chapter 6, so that qjP (t) satisfies the
relationship

q̈jP (t) + ωj2 qjP (t) = Qj (t) j = 1, 2, . . . , n (7.88)

The homogeneous and particular solutions are combined to obtain the response as

qj (t) = qjH (t) + qjP (t) = Aj cos (ωj t − φj ) + qjP (t) (7.89)

The next step is to solve for the unknowns Aj and φj in terms of the initial conditions
qj (0) and q̇j (0). Note that we solve for the unknowns in each modal equation separately.
The modal displacement at time t = 0 is

qj (0) = Aj cos φj + qjP (0) (7.90)


340 Applied Dynamics

The modal velocity is obtained by differentiating the modal displacement:

q̇j (t) = −ωj Aj sin (ωj t − φj ) + q̇jP (t) (7.91)

Evaluating this expression at the initial time t = 0 gives

q̇j (0) = ωj Aj sin φj + q̇jP (0) (7.92)

Rewriting the initial condition expressions as


1
Aj cos φj = qj (0) − qjP (0) Aj sin φj = (q̇j (0) − q̇jP (0)) (7.93)
ωj

and squaring and adding the above two expressions and solving for Aj results in
s  2
2 q̇j (0) − q̇jP (0)
Aj = (qj (0) − qjP (0)) + (7.94)
ωj

Dividing the two equations in Equation (7.93) gives the modal phase angle as
 
−1 q̇j (0) − q̇jP (0)
φj = tan (7.95)
ωj (qj (0) − qjP (0))

Using the homogeneous plus particular solution is generally more cumbersome than the
Laplace transform solution for the modal response in Equation (7.80) because of the need
to solve for the modal amplitudes and modal phase angles. However, for certain types of
excitation, it may be easier to obtain a particular solution than to evaluate a convolution
integral.

7.10.3 Response of Modes with Zero Eigenvalues


As discussed in Section 7.7, unrestrained systems have zero frequencies associated with their
rigid body modes. The modal equation associated with the rigid body mode (say, q1 (t), so
that ω1 = 0) has the form

q̈1 (t) = Q1 (t) (7.96)

We can obtain the solution to the above equation by considering the general solution,
Equation (7.80), and by taking the limit as the natural frequency approaches zero. We can
rewrite Equation (7.80) as
Z t
sin ωj t sin ωj τ
qj (t) = qj (0) cos ωj t + q̇j (0) t + Qj (t − τ ) τ dτ (7.97)
ωj t 0 ωj τ

Noting that
sin ωj t
lim = 1 lim cos ωj t = 1 (7.98)
ωj t→0 ωj t ωj t→0

and taking the limit of the solution as ωj t → 0 for the general case gives the response for a
zero frequency mode (j = 1) as
Z t
q1 (t) = q1 (0) + q̇1 (0) t + Q1 (t − τ ) τ dτ (7.99)
0
Response of Multi-Degrees-of-Freedom Systems 341

7.10.4 Response to Impulsive Loading


From the developments of Chapter 6, an idealized impulsive force, that is, a force of very
large magnitude applied over a very short period of time, has the effect of causing a sudden
change in velocity, with no change in position. Here, we extend this concept to modal
equations.
Without loss of generality, consider a multi-degrees-of-freedom system where the force
vector has only one nonzero element, say, in the l-th location. The force vector has the form

{F (t)} = [0 0 . . . 0 F (t) 0 0 . . . 0]T (7.100)

Writing the j-th modal vector as {uj } = [u1j u2j . . . unj ]T , the modal force for the j-th
mode has the form

Qj (t) = {uj }T {F (t)} = ulj F (t) , j = 1, 2, . . . , n (7.101)

The applied force is impulsive, so F (t) = F̂ δ (t). It follows that all the modal forces are
also impulsive and they have the form

Qj (t) = ulj F̂ δ (t) = Q̂j δ (t) (7.102)

so the impulsive modal force is Q̂j = ulj F̂ .


For the modal response consider, again without loss of generality, zero initial conditions.
We saw in Chapter 6 that an impulsive force F̂ causes a sudden change in velocity of
F̂ /m. It follows that an impulsive modal force causes a sudden change in modal velocity
of q̇j (0+ ) = Q̂j . When the initial conditions are zero, the response of each mode can be
written as

q̇j (0+ ) Q̂j ulj F̂


qj (t) = sin ωj t = sin ωj t = sin ωj t (7.103)
ωj ωj ωj

As discussed in the previous chapter, the validity of the impulsive motion assumption
should be checked. Because the mode with the highest frequency, the n-th mode, has the
shortest period, the duration of the impulsive excitation ∆ should be much smaller than
the smallest period, that is, of the n-th mode. Using the guideline of one tenth the period,
for multi-degrees-of-freedom systems, we have
Tn π
∆ < = (7.104)
10 5ωn

Example 7.18
Given the cart-rod system moving about stable equilibrium, obtain the response to an
impulsive force F̂ . The system is at rest before the impulsive force is applied.
The modal forces are
Qj = {uj }T {F } [a]
and the natural frequencies and modal vectors are
   
0.3575 0.2876
ω1 = 1.041 ω2 = 3.414 {u1 } = {u2 } = [b]
0.0453 −0.6053

The force vector is [F 0]T . The modal forces have the form

Q1 = {u1 }T {F } = 0.3575F Q2 = {u2 }T {F } = 0.2876F [c]


342 Applied Dynamics

The applied impulsive force F̂ leads to impulsive modal forces of the form

Q̂1 = 0.3575F̂ Q̂2 = 0.2876F̂ [d]

The impulsive modal force Q̂j causes a sudden jump in the j-th modal velocity. Because
the system is at rest before the impulse is applied, the modal velocities immediately after
the impulse become

q̇1 0+ = Q̂1 = 0.3575F̂ q̇2 0+ = Q̂2 = 0.2876F̂


 
[e]

From Equation (7.103), the response is

q̇j (0+ )
qj (t) = sin ωj t [f ]
ωj
so that the modal responses become

0.3575F̂ 0.2876F̂
q1 (t) = sin 1.041t q2 (t) = sin 3.414t [g]
1.041 3.414
The total response is obtained by adding the modal responses as {x (t)} = {u1 }q1 (t) +
{u2 }q2 (t), which gives
   
0.3575 0.3575F̂ 0.2876 0.2876F̂
{x (t)} = sin 1.041t + sin 3.414t
0.0453 1.041 −0.6053 3.414
   
0.1228 0.0242
= F̂ sin 1.041t + F̂ sin 3.414t [h]
0.0156 −0.0510

Example 7.19
For the mass-spring system in Figure 7.20, obtain the response for the following parameters
and initial conditions: m1 = m, m2 = 2m, k1 = k2 = k, k3 = 2k, F1 (t) = F0 , F2 (t) = 0,
x1 (0) = 1.2, x2 (0) = 0, ẋ1 (0) = 0, ẋ2 (0) = 0.0.

#! ##
"! "# ""
!! !#

$! $#

FIGURE 7.20
Two-degrees-of-freedom mass-spring system.

Step 1. The free-body diagrams are given in Figure 7.21. Summing forces in the horizontal
direction, we obtain the equations of motion. For the first mass, the force balance gives
+
X X
→ F = m1 a1 a1 = ẍ1 F = −k1 x1 + k2 (x2 − x1 ) + F0

=⇒ m1 ẍ1 + (k1 + k2 ) x1 − k2 x2 = F0 [a]


Response of Multi-Degrees-of-Freedom Systems 343

F1 F2
k1x1
m1 m2
k2(x2 – x1) k3x2

FIGURE 7.21
Free-body diagrams.

and for the second mass


+
X X
→ F = m2 a2 a2 = ẍ2 F = −k3 x2 − k2 (x2 − x1 )

=⇒ m2 ẍ2 − k2 x1 + (k2 + k3 ) x2 = 0 [b]


T
Introducing the position vector {x} = [x1 x2 ] , the equations of motion are written in
matrix form, where the mass and stiffness matrices are
       
m1 0 m 0 k1 + k2 −k2 2k −k
[m] = = [k] = = [c]
0 m2 0 2m −k2 k2 + k3 −k 3k
Step 2. The natural frequencies and modal vectors are calculated next. The frequency
equation is
−ω 2 m + 2k
 
 2  −k
−ω [m] + [k] {u} = {u} = {0} [d]
−k −2ω 2 m + 3k
ω2 m
Dividing both rows of the above equation by k and introducing the variable β = k ,
the characteristic equation becomes
 
−β + 2 −1
det = 2β 2 − 7β + 5 = 0 [e]
−1 −2β + 3
The solution of the characteristic equation is
1 p  7±3
β = 7 ± 72 − 4 × 2 × 5 = = 1, 2.5 [f ]
4 4
so that the natural frequencies are
k k
β1 = 1 =⇒ ω12 = β2 = 2.5 =⇒ ω22 = 2.5 [g]
m m
To find the modal vectors, we can use either ωj2 or βj (j = 1, 2). The simpler notation
is to use βj . For the first mode, (β1 = 1),
      
−β1 + 2 −1 1 −1 u11 0
{u1 } = {0} =⇒ = [h]
−1 −2β1 + 3 −1 1 u21 0
The coefficient matrix is indeed singular. We can use either the top or bottom expression
in Equation [h] to solve for the elements of the modal vector, which gives u11 − u21 = 0, so
that the first modal vector can be written as
 
1
{u1 } = d1 [i]
1
344 Applied Dynamics

Based on the discussion in Section 7.7, {u1 } is a rigid body mode, where both masses
move with the same amplitude. This rigid body mode has a nonzero natural frequency.
For the second modal vector, (β2 = 2.5), the modal vector equation becomes
      
−β2 + 2 −1 −0.5 −1 u12 0
{u2 } = {0} =⇒ = [j]
−1 −2β2 + 3 −1 −2 u22 0

Again, the coefficient matrix is singular. Solving for the elements of the modal vector
gives  
1
{u2 } = d2 [k]
−0.5
Note the sign change in the elements of the second modal vector. A check for orthogonal-
ity verifies that the modal vectors are indeed orthogonal. The modal vectors are normalized
using Equation (7.17), with the result
 T   
T 1 1 0 1
{u1 } [m] {u1 } = md21 = 3md21 = 1 [l]
1 0 2 1

which, when solved for d1 (and arbitrarily selecting the positive value) gives d1 = √1 .
3m
For the second mode,
 T   
T 1 1 0 1
{u2 } [m] {u2 } = md22 = 1.5md22 = 1 [m]
−0.5 0 2 −0.5

which, when solved for d2 (and arbitrarily selecting the positive value) gives d2 = √ 2 .
3m
Step 3. The response of the modal equations

q̈j (t) + ωj2 qj (t) = Qj (t) , with initial conditions qj (0) , q̇j (0) , j = 1, 2 [n]

is obtained next. The modal initial conditions and modal forces are
 T    r
T m 1 1 0 1.2 m
q1 (0) = {u1 } [m] {x (0)} = √ = 1.2 [o]
3m 1 0 2 0 3
√  T    r
T 2m 1 1 0 1.2 2m
q2 (0) = {u2 } [m] {x (0)} = √ = 1.2 [p]
3m −0.5 0 2 0 3

q̇1 (0) = q̇2 (0) = 0 as the initial velocity is 0 [q]

 T  
T 1 1 F0 F0
Q1 (t) = {u1 } {F (t)} = √ = √ [r]
3m 1 0 3m
√  T   √
T 2 1 F0 2F0
Q2 (t) = {u2 } {F (t)} = √ = √ [s]
3m −0.5 0 3m
Step 4. Both modal forces are constant inputs. Section 6.13 derives the response for a step
input. Repeating the procedure here, and noting that Q1 (t) and Q2 (t) are both constants,
gives the following expression for the convolution sum:
Z t Z t
sin ωj τ sin ωj τ 1 − cos ωj t
Qj (t − τ ) dτ = Qj dτ = Qj [t]
0 ω j 0 ωj ωj2
Response of Multi-Degrees-of-Freedom Systems 345

We are now ready to obtain the response. For the first mode, the response is
r
1 − cos ω1 t m F0 1 − cos ω1 t
q1 (t) = q1 (0) cos ω1 t + Q1 = 1.2 cos ω1 t + √ [u]
ω12 3 3m ω12

and for the second mode the response has the form
r √
1 − cos ω2 t 2m 2F0 1 − cos ω2 t
q2 (t) = q2 (0) cos ω2 t + Q2 = 1.2 cos ω2 t + √ [v]
ω22 3 3m ω22

Step 5. Using the expansion theorem, the total response is expressed as

{x (t)} = {u1 }q1 (t) + {u2 }q2 (t)

     
1 F0 1 − cos ω1 t 1 2F0 1 − cos ω2 t
= 0.4 cos ω1 t + + 0.8 cos ω2 t + [w]
1 3m ω12 −0.5 3m ω22
Let√us check the units of the expressions in the response. The units of the modal vectors
are
√ 1/ mass. From Equations [o] and [p] the unit√ of the modal coordinates √ is then distance×
mass. Hence, the unit of the response is (1/ mass)×distance×(mass/√ mass) = distance.
Looking at the forcing, the unit of the modal force Qj is force/ mass. Noting that the
units of the natural
√ frequencies are√1/time, the units of the convolution integral become
force × time2 / mass = distance × mass, the same as for the modal coordinates. These
observations are verified by examining Equation [w].

7.11 Mode Participation and Isolation


The previous section made use of the expansion theorem and a coordinate transformation
that made it possible to express the equations of motion as n independent equations in
terms of modal coordinates. We used the response of a single-degree-of-freedom system,
which was derived in Chapter 6, to write the response of each mode. This section examines
the magnitude of the terms in the modal response and discusses which modes contribute
more to the motion.

7.11.1 Mode Participation


Begin with an n-degrees-of-freedom system whose equation of motion is [m] {ẍ (t)} +
[k] {x (t)} = {F (t)}, and whose natural frequencies and modal vectors are ωj and {uj }
(j = 1, 2, . . . , n). Using the modal expansion
n
X
{x (t)} = {uj }qj (t) (7.105)
j=1

we arrive at the modal equations of motion, Equation (7.79), whose solution is obtained in
Equation (7.80).
Based on observations from previous examples, elements of the normalized modal vec-
tors are of comparable magnitude. This is indeed the case for most systems. Let us now
examine the magnitudes of the modal forces. Without loss of generality, consider zero initial
346 Applied Dynamics

conditions and that the force vector has only one nonzero element, say, in the l-th location.
The force vector has the form

{F (t)} = [0 0 . . . 0 F (t) 0 0 . . . 0]T (7.106)

Writing the j-th modal vector as {uj } = [u1j u2j . . . unj ]T , the modal force for the j-th
mode has the form

Qj (t) = {uj }T {F (t)} = ulj F (t) , j = 1, 2, . . . , n (7.107)

Because elements of the modal vectors have similar magnitudes, we conclude that, unless
the external forces are designed to have specific values, the modal forces Qj (t) also have
similar magnitudes.
Next, calculate and compare the modal response for a variety of forcing functions. By
using the first convenience, the natural frequencies are expressed in ascending order. Con-
sider first an impulsive force F (t) = F̂ δ (t). It follows that all the modal forces are also
impulsive and, as discussed in Section 7.10.4, they have the form

Qj (t) = ulj F̂ δ (t) = Q̂j δ (t) (7.108)

Following the argument that the magnitudes of ulj are similar, the magnitudes of Q̂j =
ulj F̂ also are similar. As discussed in Section 7.10.4, an impulsive modal force causes a
sudden change in modal velocity of q̇j (0+ ) = Q̂j . For zero initial conditions, the response
of each mode can be written as
q̇j (0+ ) Q̂j ulj F̂
qj (t) = sin ωj t = sin ωj t = sin ωj t (7.109)
ωj ωj ωj
Comparing the response of the j-th and (j + 1)-th modes, we get
qj (t) ulj ωj+1 sin ωj t
= (7.110)
qj+1 (t) ul(j+1) ωj sin ω(j+1) t
lju j sin ω t
Two of the three ratios in Equation (7.110), ul(j+1) and sin ω(j+1) t , are of the same order
of magnitude, so that the ratio of the responses can be expressed as
qj (t) ωj+1
≈ (7.111)
qj+1 (t) ωj
This equation indicates that amplitudes of the modal vectors decrease by 1/ωj when an
impulsive force is applied. The lowest mode has the highest amplitude and the amplitudes
of the responses of the modes decrease in proportion to the natural frequencies.
Next, consider the step response, that is, when F (t) = F0 u (t). It follows that the modal
forces are step functions with magnitudes Qj (t) = ulj F0 u (t) and that they too have similar
magnitudes. From Section 6.13, the response of a single-degree-of-freedom system to a step
F0
input has the form x (t) = mω 2 (1 − cos ωn t) u (t), so that the step response of each mode is
n

ujl F0
qj (t) = (1 − cos ωj t) u (t) (7.112)
ωj2

Following an argument similar to the one presented for the impulse response, the ratio
of the modal responses to a step input can be expressed as
 2
qj (t) ωj+1
≈ (7.113)
qj+1 (t) ωj
Response of Multi-Degrees-of-Freedom Systems 347

The amplitude of the step response for each mode decreases by the square of the natural
frequency. This is a much faster rate of decrease than in the impulse response. For a force
that increases linearly with time (ramp response) such as F (t) = F0 t, we can show that the
amplitudes of the modes decrease by 1/ωj3 . The following conclusions can be drawn from
the above analysis:
• An impulsive force excites the higher modes more than any other type of excitation.
• In general, if the signature of the time excitation is expressed as tκ , the participation of
each mode decreases by 1/ωjκ+2 .
• Except for special cases of external excitation, the lower modes have higher amplitudes
than the higher modes.
• The mode with the lowest natural frequency has the highest amplitude and hence con-
tributes more to the motion than the other modes. This is why when approximating
a multi-degrees-of-freedom problem by a single-degree-of-freedom model, we retain the
fundamental mode, that is, the mode associated with the lowest natural frequency, also
known as the fundamental frequency.

7.11.2 Mode Isolation—Mode Control


In certain situations, it may be desirable to provide a system with external inputs so that
only one of the modes is excited and the others are not. Or, we may be interested in observing
the amplitude of a particular mode. This concept is called mode isolation. We need to extract
that particular mode’s response from the measured response data to accomplish this.
Measurement of the amplitude of a particular mode can be accomplished by means of
the expansion theorem. Indeed, writing the expansion theorem for the s-th mode
n
X
{x (t)} = {uj }qj (t) qj (t) = {uj }T [m] {x (t)} (7.114)
j=1

If we measure all of the displacement coordinates {x (t)}, we can then use measurements
of {x (t)} and the above equation to extract qk (t) from it. The concept is known as modal
space measurement and it requires that there be as many measurements as there are degrees
of freedom. When this requirement is physically not feasible and fewer sensors are available,
we use estimation techniques.
The mode isolation problem for the s-th mode is posed mathematically as follows: Given
the modal equations of motion in Equation (7.79), design the forcing function {F (t)} such
that the resulting modal forces have the form

Qj (t) = {uj }T {F (t)} = Q (t) δjs j = 1, 2, . . . , n (7.115)

In other words, it is desired that all modal excitations be zero, except for Qs (t). To
accomplish this, we need to invert the relationship between the modal force vector and
external input vector, Equation (7.85), so that

{F (t)} = [U ]−T {Q (t)} (7.116)

The above equation calculates the actual force inputs. For the mode isolation procedure
to work, the input matrix {F (t)} must not have any zero elements. That is, for an n-
degrees-of-freedom mass-spring system, there must be n forces, one on each mass.
The developments above can be extended to the mode control process, where we design
348 Applied Dynamics

the external inputs so that the system will behave in a prescribed manner. For example,
we may want to impart damping forces to the first mode which may not have adequate
damping without affecting the other modes. We design the modal forces as

Q1 (t) = −cq̇1 (t) Q2 (t) = Q3 (t) = . . . = Qn (t) = 0 (7.117)

where c is an energy dissipation coefficient. The resulting modal equations become

q̈1 (t) + cq̇1 (t) + ω12 q1 (t) = 0, q̈j (t) + ωj2 qj (t) = 0 j = 2, 3, . . . , n (7.118)

The first mode behaves like a damped oscillator, while the other modes are unaffected
by the control action.

7.12 Approximate Approach for Damped Systems


It is possible to conduct an exact modal analysis of a damped system. The procedure
is lengthy and many times not of great interest, because of the added complexity over un-
damped systems and also because of the inaccuracies associated with modeling the damping
matrix. As discussed in the previous chapter, mass and stiffness properties are known more
accurately than damping properties.
Including a damper in the mathematical model is an idealization of the energy dissipation
process. Except for cases where dampers whose properties are known are used, such as in
automotive applications, the energy dissipation characteristics of most dynamical systems
are not accurately known. Several energy dissipation mechanisms may simultaneously act,
such as friction, drag and hysteresis, in addition to viscous damping. This section describes
an approximate method for obtaining the eigensolution of a viscously damped system.
The equations of motion of a damped multi-degrees-of-freedom system are

[m] {ẍ (t)} + [c] {ẋ (t)} + [k] {x (t)} = {F (t)} (7.119)

Using the same approach of synchronous motion as previously, the displacement is written
as {x (t)} = {u} eλt and introduced to the equation of motion. The resulting characteristic
equation is

det λ2 [m] + λ [c] + [k] = 0


 
(7.120)

The characteristic equation is a polynomial of order 2n in λ, with 2n roots and can


be solved using standard techniques. However, we need to go to a state-space formulation
consisting of matrices of order 2n to find the associated modal vectors. State form of the
equations of motion is discussed in Section 7.16. The procedure is cumbersome. Furthermore,
there usually are uncertainties with the damping model.
The approximate procedure begins by obtaining the eigensolution of the corresponding
undamped system in the form of the natural frequencies ωj and modal vectors {uj }, (j =
1, 2, . . . , n). As in Section 7.10, the expansion
n
X
{x (t)} = {uj } qj (t) (7.121)
j=1

is introduced to the equation of motion. Left multiplying by {us }T , rearranging terms, and
Response of Multi-Degrees-of-Freedom Systems 349

making use of orthogonality of the modal vectors gives


n
X
q̈s (t) + c0sj q̇j (t) + ωs2 qs (t) = Qs (t) = {us }T {F (t)} (7.122)
j=1

in which

c0sj = {us }T [c] {uj } j, s = 1, 2, . . . , n (7.123)

The modal damping terms can be written in matrix form as

[c0 ] = [U ]T [c] [U ] (7.124)

where [U ] = [{u1 } {u2 } . . . {un }]T is the modal matrix and the entries of [c0 ] are c0sj .
The equations of motion in Equation (7.122), which are in terms of modal coordinates
of the undamped system, are still coupled, so they cannot be solved independently of each
other. For a complete decoupling of the equations of motion, we need to conduct an exact
analysis. Under certain circumstances, it is possible to neglect the off-diagonal terms in
[c0 ] = [U ]T [c] [U ], or set c0sj = 0 for j 6= s. This assumption leads to a set of decoupled
equations. Introducing the modal damping factor ζs , such that
c0ss
ζs = (7.125)
2ωs
we can rewrite Equation (7.122) as

q̈s (t) + 2ζs ωs q̇j (t) + ωs2 qs (t) = Qs (t) s = 1, 2, . . . , n (7.126)

which are now decoupled and independent of each other. Hence, they can be solved in-
dependently of each other. Chapter 6 derived the response of a single-degree-of-freedom
underdamped system. Using Equation (6.145), the response of each mode has the form
 
−ζs ωs t q̇s (0) + qs (0) ζs ωs
qs (t) = e qs (0) cos ωsd t + sin ωsd t
ωsd

Z t
1
+ Qs (t − τ ) e−ζs ωs τ sin ωsd τ dτ (7.127)
ωsd 0
p
in which ωsd = ωs 1 − ζs2 is the damped natural frequency associated with the s-th mode.
The assumption of ignoring the off-diagonal elements of [c0 ] becomes reasonable under
the following circumstances:
• When damping is small, the contribution to the response of the c0sj terms when s 6= j
is less significant than the diagonal terms c0ss , even though the magnitudes of the off-
diagonal elements of [c0 ] may be similar in magnitude to the diagonal terms c0ss .
• When assuming proportional damping (also called Rayleigh damping). Under this as-
sumption, the damping matrix can be written as a combination of the mass and stiffness
matrices as

[c] = α [m] + β [k] (7.128)

in which α and β are coefficients. As a result c0sj = 0 for s 6= j and c0ss = α + βωs2 . While
this model is more of a mathematical convenience, it occasionally represents a realistic
model.
350 Applied Dynamics

• When the energy dissipation characteristics are not accurately known, the error associ-
ated with using an incorrect damping model is usually more significant than ignoring the
c0sj terms. While by doing this we add another error to an already incorrect model, the
cumulative effect of the errors usually is not large, especially for low levels of damping.
When the approximation is no longer valid, we need to use the exact solution for damped
systems. This solution can be found in texts on vibration analysis.
An interesting property of damped systems is that, in general, the lower modes have
lower damping factors than the higher modes. Adding this property to the discussion on
mode participation, we conclude that the contribution of the lower modes to the overall
motion becomes even more significant, as the higher modes damp out faster than the lower
modes.

Example 7.20

#! #"

!! !"

"! ""

$! $"

FIGURE 7.22
Two-degrees-of-freedom damped system.

This example examines the range of validity of the damping assumption in the previous
section. Consider the system in Figure 7.22. The mass, stiffness, and damping matrices are
     
m1 0 k1 + k2 −k2 c1 + c2 −c2
[m] = [k] = [c] = [a]
0 m2 −k2 k2 −c2 c2

with m1 = 3 kg, m2 = 6 kg, k1 = 4 N/m, k2 = 2 N/m, c1 = 0, c2 = 1 N·s/m. The natural


frequencies and normalized modal vectors of the undamped system are
   
0.1470 0.5583
ω1 = 0.4574, ω2 = 1.4574 rad/s {u1 } = {u2 } = [b]
0.3948 −0.1039

Defining the modal matrix as [U ] = [{u1 } {u2 }], we obtain for [c0 ]
 
0.06142 −0.1641
[c0 ] = [U ]T [c] [U ] = [c]
−0.1641 0.4384

As discussed in the previous section, the off-diagonal elements of [c0 ] are larger than c011 .
The approximate values of the damping factors are
c011 0.06142 c022 0.4384
ζ1 = = = 0.06713 ζ2 = = = 0.1505 [d]
2ω1 2 × 0.4574 2ω2 2 × 1.4574
We next compare this approximate solution with the actual solution. Solving the char-
acteristic equation in Equation (7.120), we can show that the actual eigenvalues are

λ1,2 = −0.0308 ± i0.4597 λ3,4 = −0.2192 ± i1.4304 [e]


Response of Multi-Degrees-of-Freedom Systems 351

From Chapter 6, the eigenvalues of an underdamped system have the form


p
λ = −ζωn ± iωn 1 − ζ 2 [f ]

Writing a generic eigenvalue as λ = −A + iB and equating it to the above equation leads


to p
ζωn = A ωn 1 − ζ 2 = B [g]
Squaring each term above and adding gives A2 + B 2 = ωn2 , from which we conclude that
p A
ωn = A2 + B 2 ζ = [h]
ωn
Applying Equation [h], the exact damping factors and natural frequencies are

ζ1 = 0.06696 ζ2 = 0.1514 ω1 = 0.4607 ω2 = 1.4471 [i]

which are very close to the approximate values with errors less than 1% in the damping
factors.

TABLE 7.2
Exact and approximate eigensolution for c1 = 0, c2 = 5 N·s/m

Solution ζ1 ω1 ζ2 ω2
Exact 0.2529 0.5689 0.9440 1.1718
Approximate 0.3357 0.4574 0.7523 1.4574

Let us explore the limitations of the approximate approach. Table 7.2 gives the exact and
approximate values for the natural frequencies and damping factors for damping constants
of c1 = 0, c2 = 5 N·s/m. The value of c2 tests the limit as it leads to a damping factor
for the second mode that approaches unity. Considering the discussion in Chapter 6, such
high damping factors are rarely encountered. The exact and approximate values are quite
far apart. Table 7.3 shows the results when the value of the damping constant is reduced
to a smaller value, c2 = 3 N·s/m. Here, the error in the damping factors is less than 7%, a
more reasonable value.

TABLE 7.3
Exact and approximate eigensolution for c1 = 0, c2 = 3 N·s/m

Solution ζ1 ω1 ζ2 ω2
Exact 0.1938 0.4911 0.4824 1.3575
Approximate 0.2014 0.4574 0.4514 1.4574

It should be noted that the above example, while representative of a tuned mass damper
that will be discussed later in this chapter, is not a very accurate model of a general damped
system. This is because the example neglects the first damper, so c1 = 0. Consider a more
realistic example, where both c1 and c2 are not zero, say, c1 = 1, c2 = 2 N·s/m. The results
are given in Table 7.4.
352 Applied Dynamics

TABLE 7.4
Exact and approximate eigensolution for c1 = 1, c2 = 2 N·s/m

Solution ζ1 ω1 ζ2 ω2
Exact 0.1562 0.4650 0.4143 1.4337
Approximate 0.1579 0.4574 0.4079 1.4574

The exact and approximate results are very close, with less than 2% difference. This is
because the damping matrix for this case resembles the stiffness matrix more than when
c1 = 0. The conclusion is, with some caution, that the approximate approach for damped
systems presented in the previous section is acceptable for most cases of damping expected
in mechanical systems. For example, for vehicle suspensions, the damping factor is in the
range ζ = 0.2–0.4.

7.13 Response to Harmonic Excitation


As discussed in the previous chapter, harmonic or periodic excitations are frequently encoun-
tered in engineered systems, and they can have dangerous consequences. A multi-degrees-
of-freedom system of order n has n natural frequencies, so a harmonic excitation can create
resonance in n possible ways, as opposed to one possible way in a single-degree-of-freedom
system.
As discussed in Chapter 6, we analyze response to periodic excitations in a different way
from response to general excitation. With a system subjected to periodic excitation, the
interest is in the steady-state motion, that is, in the motion amplitudes and phase angles
that result after the effects of initial conditions and other transient effects (such as impulsive
or other short-term forces) have died out.
This section studies the response of multi-degrees-of-freedom systems to harmonic ex-
citation. The results can be extended to periodic excitations by representing a periodic
excitation in terms of a Fourier series and by making use of the principle of superposition.

7.13.1 General Formulation


Consider a multi-degrees-of-freedom system subjected to a harmonic excitation. It is con-
venient to express the harmonic excitation in complex form
[m] {ẍ (t)} + [c] {ẋ (t)} + [k] {x (t)} = {F (t)} = {F0 } eiωt (7.129)
in which the vector {F0 } is the amplitude of the harmonic excitation. The steady-state
response is, as described in the previous chapter, the particular solution. The steady-state
response has the same form as the excitation, and it can be written as
{xss (t)} = {X} eiωt (7.130)
where {X} is the steady-state response amplitude.
Substituting Equation (7.130) into the equations of motion, collecting like terms, and
eliminating eiωt from both sides gives
−ω 2 [m] + iω [c] + [k] {X} = [I (iω)] {X} = {F0 }

(7.131)
Response of Multi-Degrees-of-Freedom Systems 353

in which [I (iω)] = −ω 2 [m] + iω [c] + [k] is known as the impedance matrix, a term coined
in electrical theory. The impedance matrix has real and complex terms in it. The above
equation is known as the frequency response equation. The steady-state amplitudes are
obtained by inverting the impedance matrix
{X} = [I (iω)]−1 {F0 } (7.132)
The inverse of the impedance matrix is known as the transmissibility matrix, defined as
[I (iω)]−1 = [H (iω)]. It is of interest to analyze amplitudes of the steady-state response,
that is, the magnitude of {X}, as a function of the excitation frequency. Hand calculations
are restricted to two-degrees-of-freedom systems. For a system of order three or higher, we
need to use symbolic manipulation. Another way to calculate the inverse of the impedance
matrix is to create an array of excitation frequencies of interest, invert the impedance
matrix at each excitation frequency, and plot the amplitudes as a function of the excitation
frequency.

7.13.2 Special Case: Two-Degrees-of-Freedom Systems


A closed-form solution for the inverse of the impedance matrix is available for two-degrees-
of-freedom systems. The impedance matrix is can be written as
I11 I12
 
[I (iω)] = (7.133)
I21 I22
The impedance matrix is symmetric. Writing the response and force vectors as {X} =
[X1 X2 ]T and {F0 } = [F1 F2 ]T , we can solve for the response amplitudes by
  
1 I22 −I12 F1
{X} = [I (iω)]−1 {F0 } = (7.134)
2
I11 I22 − I12 −I12 I11 F2
with the result
I22 F1 − I12 F2 −I12 F1 + I11 F2
X1 = 2 X2 = 2 (7.135)
I11 I22 − I12 I11 I22 − I12
Consider the mass-spring-damper system in Figure 7.23. Using x1 and x2 as motion
variables, we write the equations of motion in matrix form, with the coefficient matrices
     
m1 0 c1 + c2 −c2 k1 + k2 −k2
[m] = [c] = [k] = (7.136)
0 m2 −c2 c2 + c3 −k2 k2 + k3
It follows that elements of the impedance matrix are
I11 = −m1 ω 2 + iω (c1 + c2 ) + k1 + k2 I12 = I21 = −iωc2 − k2

I22 = −m2 ω 2 + iω (c2 + c3 ) + k2 + k3 (7.137)


Introducing the above relationships into Equation (7.135) produces the general solution
for the damped case. The solution is lengthy and not given here.
For undamped systems (c1 = c2 = c3 = 0), and, without loss of generality, setting
F2 = 0, leads to the steady-state values X1 and X2 in the form

k2 + k3 − m2 ω 2 F1
X1 = (7.138)
(k1 + k2 − m1 ω 2 ) (k2 + k3 − m2 ω 2 ) − k22

−k2 F1
X2 = (7.139)
(k1 + k2 − m1 ω 2 ) (k2 + k3 − m2 ω 2 ) − k22
The amplitudes X1 and X2 are real for the undamped case.
354 Applied Dynamics

$! $"
#! #" ##

!! !"

"! "" "#

%! %"

FIGURE 7.23
Two-degrees-of-freedom system.

Example 7.21
Consider the two-degrees-of-freedom system in Figure 7.20. Obtain the frequency response
and plot X1 and X2 for the undamped case and when m1 = m, m2 = 2m and k1 = k2 =
k, k3 = 2k.
The coefficient matrices are
   
m 0 2k −k
[m] = [k] = [a]
0 2m −k 3k

and the characteristic equation is

det −ω 2 [m] + [k]



= 0 [b]

Dividing each row of the above matrix by k and introducing the quantity β = ω 2 m/k,
the characteristic equation becomes
 
β−2 1
det = 2β 2 − 7β + 5 = 0 [c]
1 2β − 3

whose solution is β1 = 1, β2 = 2.5, so the natural frequencies are


r r r
k 2.5k k
ω1 = ω2 = = 1.581 [d]
m m m
The next step is to calculate the steady-state amplitudes. Substitution of the mass and
stiffness values into the expressions for X1 and X2 yields

3k − 2mω 2 F1 −kF1
X1 = X2 = [e]
(3k − 2mω 2 ) (2k − mω 2 ) − k 2 (3k − 2mω 2 ) (2k − mω 2 ) − k 2
We can plot the amplitudes X1 and X2 using their forms above, but it is more instructive
to express them in terms of the natural frequencies. Introduce the frequency ratios
ω ω2 √
ω̄ = ω̄2 = = 2.5 [f ]
ω1 ω1
Let us divide both the top and bottom of the expressions for X1 and X2 by k. The
numerator in the left term in Equation [e] when divided by k becomes

3k − 2mω 2 F1
= 3 − 2ω̄ 2 F1

[g]
k
Response of Multi-Degrees-of-Freedom Systems 355

The denominator in Equation [e] can be expressed as


3k − 2mω 2 2k − mω 2 − k 2 = 2m2 ω14 1 − ω̄ 2 ω̄22 − ω̄ 2
   

= 2k 2 1 − ω̄ 2 ω̄22 − ω̄ 2
 
[h]
Substituting these expressions into Equation [e], the steady-state amplitudes become
F1 3 − 2ω̄ 2 F1 1
X1 = X2 = [i]
2k (ω̄22 − ω̄ 2 ) (1 − ω̄ 2 ) 2
2k (ω̄2 − ω̄ ) (1 − ω̄ 2 )
2

Figure 7.24 plots the steady-state amplitudes, X1 and X2 (scaled by F2k1 ), as a function of
the excitation frequency ratio ω̄. There are two peaks, when ω equals the natural frequencies
ω1 (ω̄ = 1) and ω2 (ω̄ = 1.581). The steady-state values of the amplitudes at the peaks
are infinity, indicating resonance. In the presence of damping, amplitudes of the peaks are
reduced, but the high amplitudes at resonance are still present.

a) b)

2kX1 6 2kX2 6
F1 F1
4 4
2 2
0 0
–2 –2
–4 –4
–6 –6
0 1 2 3 0 1 2 3
= /1 = /1

FIGURE 7.24
Steady-state amplitudes of two-degrees-of-freedom system. a) 2kX1 /F1 , b) 2kX2 /F1 .

In vibration testing, we generate plots like Figure 7.24 by taking a system whose param-
eters are not known accurately. Subjecting the system to a range of harmonic excitations
and plotting the response amplitudes, we can determine the natural frequencies and identify
system parameters. There are several types of vibration exciters in the marketplace for such
testing, such as shakers and vibration tables.

7.13.3 Modal Analysis


The previous analysis formulated the response to harmonic excitation for multi-degrees-of-
freedom systems in terms of the impedance matrix, which needs to be inverted to find the
steady-state amplitudes. This section considers the modal response.
Consider an undamped system subjected to a single harmonic excitation
[m] {ẍ (t)} + [k] {x (t)} = {F (t)} = {F0 } eiωt (7.140)
in which {F0 } is the vector of excitation amplitudes. The modal equations of motion are
q̈j (t) + ωj2 qj (t) = Qj (t) j = 1, 2, . . . , n (7.141)
356 Applied Dynamics

where the modal forces have the form


Qj (t) = {uj }T {F (t)} = {uj }T {F0 } eiωt (7.142)
Recalling that the steady-state solution is the particular solution and considering a
particular solution in the form qj (t) = qj0 eiωt , the steady-state response of each mode is
obtained as
{uj }T {F0 } 1
qjss (t) = 2 eiωt (7.143)
ωj 1 − ω̄j2
in which ω̄j = ωωj is the frequency ratio for each mode. The steady-state response of the
entire system can be written using the expansion theorem as
 
n n T
X X {u j } {F0 } 1  iωt
{xss (t)} = {uj } qj (t) =  2 2 {uj } e (7.144)
j=1 j=1
ω j 1 − ω̄j

Two observations can be made from the above equation. First, the system has n natural
frequencies so that the system can be brought into resonance n different ways. Second,
the steady-state amplitude of each mode contains the coefficients {uj }T {F0 } /ωj2 , (j =
1, 2, . . . , n), indicating that the mode participation decreases with the square of the natural
frequency. As in the case of aperiodic excitations, the lower modes have larger steady-state
amplitudes than the higher modes.

7.14 Vibration Reducing Devices


A very important and frequently used application of response to harmonic excitation is
the design of vibration reducing devices. A vibration reducing device consists of a mass-
spring-damper or similar system that is added to an existing vibrating system (referred to
as the primary system) that is subjected to harmonic or other type of excitation. We lower
the amplitudes of the primary system by properly selecting the properties of the vibration
reducer. This reduction is achieved in two different ways:
• By means of a vibration absorber, which transfers the energy of the primary system
to the added device. A vibration absorber is usually designed without a damper, as
the intent is to transfer vibration energy and not to dissipate it. Vibration absorbers
are frequently used with rotating machinery or other systems that are continuously
subjected to periodic excitation.
• By means of a tuned mass damper, which transfers and dissipates the energy of the
primary system. Tuned mass dampers are used in larger structures, such as tall buildings,
pipelines, and bridges, that encounter harmonic or aperiodic excitation in the form of a
wind gust, fluid flow, earthquakes, and stop and go traffic.
Tuned mass dampers are much larger than vibration absorbers. For example, two 300-
ton tuned mass dampers were added to the John Hancock Building in Boston after it
experienced vibrations that were causing discomfort to its occupants and damage in the
form of broken windows. The Taipei 101 building has a tuned mass damper in the form
of a pendulum that weighs 728 tons and is suspended over six floors of the building. The
Millennium Bridge in London experienced undesirable levels of vibration on its opening
day. A series of tuned mass dampers, as well as ordinary dampers, were added to it to
reduce vibration amplitudes.
Response of Multi-Degrees-of-Freedom Systems 357

mr
xr

a) b) kr cr

m m
xb xa
k k k k
2 c 2 2 c 2

y y

FIGURE 7.25
a) A single-degree-of-freedom vibrating system, b) added vibration reducing device.

The primary system is modeled as a mass-spring-damper system in Figure 7.25a. It is


subjected to base excitation in the form y (t) = Aeiωt . Figure 7.25b depicts the system
after the vibration reducer is added. The variables xb (t) and xa (t) denote the amplitudes
of the primary mass before and after the vibration reducing device, respectively, and xr (t)
describes the amplitude of the vibration reducer.

m
xb

k (xb – y) b

FIGURE 7.26
Free-body diagram of primary mass before reducer is added.

The free-body diagram of the primary mass before the reducer is added is shown in
Figure 7.26. The free-body diagrams of the primary mass and vibration reducer are shown
in Figure 7.27. The vibration amplitudes are being measured from static equilibrium so that
no gravity terms are involved.

7.14.1 Motion Amplitude before Vibration Reducer Is Added


From Figure 7.26, summing forces in the vertical direction and rearranging, the equation of
motion of the primary mass before the reducer is added becomes
mẍb (t) + cẋb (t) + kxb (t) = ky (t) + cẏ (t) (7.145)
Upon division by the mass m and introduction of the excitation as y (t) = Aeiωt , the
equation of motion takes the form
ẍb (t) + 2ζωn ẋb (t) + ωn2 xb (t) = A 2iζωωn + ωn2 eiωt

(7.146)
358 Applied Dynamics

kr (xr – xa ) cr r
– a
)

m mr
xa xr

k (xa – y) a
– kr (xr – xa ) cr ( r – a)

FIGURE 7.27
Free-body diagrams of primary mass and vibration reducer after reducer is added.

Using a solution in the form xb (t) = Xb eiωt and the approach in Chapter 6, the vibration
amplitude Xb before the vibration reducing device is added is obtained as
Xb = A |H (iω̄)| (7.147)
in which |H(iω̄)| is the transmissibility defined in Section 6.18. The steady-state response
is
xb (t) = A |H (iω̄)| cos (ωt − ψh ) (7.148)
where ψh is the phase angle. As the frequency ratio ω̄ = ω/ωn approaches 1, the motion
amplitudes increase and may reach undesirable or unsafe values.

7.14.2 Undamped Vibration Absorbers


A vibration absorber consists of a mass-spring system that is added to a vibrating system
subjected to harmonic excitation. A vibration absorber is usually designed as undamped,
and it transfers the vibration energy of the primary mass to itself.
Consider the case where both the primary mass and absorber are undamped. The
amplitude of the primary mass (recall that the subscript b denotes before absorber and
H (iω̄) = G (iω̄) for undamped systems) is
A
Xb (iω̄) = AH (iω̄) = AG (iω̄) = (7.149)
1 − ω̄ 2
where we recall that ω̄ = ω/ωn is the frequency ratio.
Next, the undamped vibration absorber, which consists of a mass-spring system, is added
to the primary mass, as shown in Figure 7.25. From the free-body diagrams in Figure 7.27,
the equations of motion of the resulting two-degrees-of-freedom system are
mẍa (t) + (k + kr )xa (t) − kr xr (t) = ky (t) = kAeiωt

mr ẍr (t) − kr xa (t) + kr xr (t) = 0 (7.150)


with the subscript a denoting after absorber is added and subscript r denoting the vibration
reducer amplitude. Considering a steady-state solution in the form xa (t) = Xa eiωt , xr (t) =
Xr eiωt , the response amplitudes can be expressed as
k + kr − ω 2 m −kr
    
Xa Ak
= (7.151)
−kr kr − ω 2 m r Xr 0
Response of Multi-Degrees-of-Freedom Systems 359

The steady-state amplitudes Xa and Xr are obtained by inverting the impedance matrix

Ak kr − ω 2 mr
Xa = (7.152)
(k + kr − ω 2 m) (kr − ω 2 mr ) − kr2

Akkr
Xr = 2
(7.153)
(k + kr − ω m) (kr − ω 2 mr ) − kr2

It is desirable to reduce the amplitude of the primary mass Xa as much as possible.


Setting Xa = 0 in Equation (7.152) gives the design condition for the absorber as

kr
kr − ω 2 mr = 0 =⇒ = ω2 (7.154)
mr
When the absorber stiffness and mass are selected so that their ratio is equal to the square of
the driving frequency, the steady-state amplitude of the primary mass becomes zero. Using
this condition, we can show that the steady-state amplitude of the absorber (vibration
reducer) becomes

Akkr k k
Xr = = −A = −A (7.155)
−kr2 kr mr ω 2

In general,pthe case of concern is when the primary mass is in resonance, that is, ω = ωn ,
where ωn = k/m is the natural frequency of the primary mass before the absorber is
added. In this case, the design condition and resulting amplitude of the absorber become
kr k k m
= ωn2 = Xr = −A 2
= −A (7.156)
mr m m r ωn mr

The two equations above are used to select the mass and stiffness of the absorber. Two
parameters need to be selected to satisfy one condition. In general, we select the absorber
mass mr to be much smaller than the primary mass m, for example, m m < 0.2. Lowering the
r

absorber mass raises its vibration amplitude, as mr is in the denominator of the expression
for Xr in Equation (7.156).
It is of interest to compare the magnitudes of the primary mass before the absorber is
added and the amplitude after the addition of the absorber. Absolute values of Xb (ω̄) and
Xa (ω̄) are plotted in Figure 7.28 as a function of the excitation frequency ratio (ω̄ = ω/ωn ),
for the case when the absorber is designed to eliminate high amplitudes at resonance, that
is, according to Equation (7.156), and for a mass ratio of mr /m = 0.2. The vibration
amplitude of the primary mass at resonance (ω̄ = 1) is zero.
From Figure 7.28, there is a small range in which the amplitude of the primary mass
after the absorber, Xa (ω̄), is smaller than its amplitude before the absorber is added,
Xb (ω̄). The mass ratio affects this range. Lowering the mass ratio reduces the range of
effectiveness of the absorber. To show this mathematically, let us calculate the points at
which |Xb (ω̄)| = |Xa (ω̄)|.
As a vibration absorber is usually designed for resonance, we design the absorber mass
and stiffness as kr /mr = ωn2 . Denoting the mass ratio by e = mr /m, it follows that kr = ek.
The mass and stiffness matrices of the system with added absorber become
   
m 0 k + ek −ek
[m] = [k] = (7.157)
0 em −ek ek
 
The characteristic determinant is det −ω 2 [m] + [k] = 0. Divide each row of
360 Applied Dynamics

10
| Xb/A| 9 Before
absorber (Xb)
| Xa/A| 8
7
After
6 absorber (Xa)
5
4
3
2
1
0
0 0.5 d1 1 d2 1.5 2

Amplitude reduction range

FIGURE 7.28
mr
Amplitudes of primary mass before and after the absorber are added for m = 0.2.

 2 
−ω [m] + [k] by k, and introduce the variable β = mω 2 /k. Because ωn = k/m, for
this case β becomes
 2
ω
β = = ω̄ 2 (7.158)
ωn
The characteristic determinant becomes
 
1+e−β −e
= e β 2 − (2 + e) β + 1 = 0

det (7.159)
−e e − eβ

The roots of the characteristic determinant are


r
e e2
β1,2 = 1 + ± e + (7.160)
2 4
q
2
We can show that e + e4 is always greater than 2e , so that β1 < 1 and β2 > 1, which
means that

ω1 < ω n ω2 > ωn (7.161)

The above result can be seen as a direct application of the enclosure theorem, also known
as the inclusion principle, which is used when obtaining approximate solutions to systems
modeled by partial differential equations, such as beams, strings, and plates.
Next, let us find the points where the amplitudes |Xb (ω̄)| and |Xa (ω̄)| coincide. There
are two possibilities:

Xb (ω̄) = Xa (ω̄) or Xb (ω̄) = −Xa (ω̄) (7.162)

The first possibility, Xb (ω̄) = Xa (ω̄), can be shown to lead to ω̄ = 0, so it is not


Response of Multi-Degrees-of-Freedom Systems 361

significant. To calculate the values of ω̄ that satisfy the second possibility, we note that
kr = ek and mr = em and write Equation (7.152) as
 
Ak kr − ω 2 mr Ake k − ω 2 m
= (7.163)
(k + kr − ω 2 m) (kr − ω 2 mr ) − kr2 (k + ke − ω 2 m) e (k − ω 2 m) − e2 k 2

Dividing both sides by k 2 , recalling that ω̄ 2 = ω 2 m/k, and equating with Equation
(7.149), we obtain

A Ae 1 − ω̄ 2
= − (7.164)
1 − ω̄ 2 (1 + e − ω̄ 2 ) e (1 − ω̄ 2 ) − e2

Let v = 1 − ω̄ 2 . Substituting to the above equation gives


1 v
= − (7.165)
v (v + e) v − e

which can be expressed as the quadratic equation

2v 2 + ev − e = 0 (7.166)

whose solution is

−e ± e2 + 8e
v = (7.167)
4
Recalling that v = 1 − ω̄ 2 and denoting the points at which the before and after curves
intersect by ω̄ = d1 and ω̄ = d2 , we obtain

di = 1 − vi i = 1, 2 (7.168)

The absorber is effective when d1 < ω̄ < d2 , in which case the primary mass has a
smaller amplitude after the absorber is added. By contrast, when ω̄ < d1 or when ω̄ > d2 ,
the primary mass attains a higher amplitude after the absorber is added.
It is of interest to investigate how β1,2 and the intersection points d1 and d2 change as
the absorber mass ratio e is varied. Figure 7.29 plots ω̄1 = ω1 /ωn , ω̄2 = ω2 /ωn , d1 , and d2
for values of the mass ratio e. The difference between the roots decreases as the mass ratio
e becomes smaller. This difference is related to the range of effectiveness of the absorber.
For a mass ratio of e = 0.5, ω̄1 = 0.707, ω̄2 = 1.41, which is quite a large spread.
The range of effectiveness of the absorber, d1 ≤ ω̄ ≤ d2 is from 0.78 to 1.28. However,
the absorber is half the size of the primary mass, which may not be a feasible design. By
contrast, for e = 0.05, the absorber is small, but so is the range of operation of the absorber.
When e = 0.05, the range for which the absorber is effective is between d1 = 0.93ωn and
d2 = 1.08ωn , less than a 10% difference from the resonant frequency.
Herein lies the problem associated with vibration absorbers. They are very sensitive to
changes in the operating frequency, they have a limited range of operation, and their range
of operation is dictated by the absorber mass. In the presence of errors or if the operating
frequency changes, the vibration absorption will not be as effective. In general, we deal with
such a situation by adjusting the absorber parameters.
The operating range of a vibration absorber does not increase by much when damping is
introduced to the primary mass-spring system or to the absorber, even though the vibration
amplitudes are reduced.
362 Applied Dynamics

1.5

1.4

1.3

1.2 ω1/ωn
ω/ωn

1.1 ω /ω
2 n

1 d1
d2
0.9

0.8

0.7
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
e

FIGURE 7.29
Frequency ratio ω̄ = ω/ωn and intersection points d1 and d2 as a function of absorber mass
ratio e = m
m .
r

7.14.3 Tuned Mass Dampers


The tuned mass damper (TMD) is modeled as a mass-spring-damper added to a primary
mass-spring system that has insufficient damping. The mass ratio for a tuned mass damper
and primary mass is still small. However, because tuned mass dampers are designed for
very large masses (such as tall buildings or bridges), they themselves end up being large.
Sometimes, instead having one very large TMD, engineers install several smaller TMDs.10
A tuned mass damper is designed to dissipate the energy of the primary mass (transient
motion), as well as reduce steady-state amplitudes.
Using the notation of the previous section and assuming an undamped primary mass,
the equations of motion become

mẍa (t) + cr ẋa (t) − cr ẋr (t) + (k + kr ) xa (t) − kr xr (t) = ky (t) = kAeiωt

mr ẍr (t) − cr ẋa (t) + cr ẋr (t) − kr xa (t) + kr xr (t) = 0 (7.169)

For a steady-state solution xa (t) = Xa eiωt , xr (t) = Xr eiωt , we can write the steady-
state equations in matrix form

k + kr + iωcr − ω 2 m −kr − iωcr


    
Xa Ak
= (7.170)
−kr − iωcr kr + iωcr − ω 2 mr Xr 0

Solution of the equation above gives the steady-state amplitude of the primary mass as

kr + iωcr − ω 2 mr
Xa (iω) = (7.171)

10 For example, 52 TMDs were installed in the Millennium Bridge in London, England.
Response of Multi-Degrees-of-Freedom Systems 363

where ∆ is the determinant of the impedance matrix. Dividing both sides by the mass of
the TMD gives
ωr2 + 2iζr ωωr − ω 2
Xa (iω) = (7.172)
mr ∆
p
in which ωr = kr /mr is the undamped natural frequency of the TMD if it were a stan-
dalone mass-spring-damper, and ζr is the corresponding damping factor. Selecting the TMD
parameters so that ωr = ω results in the smallest value of the denominator. The same design
condition that holds for vibration absorbers also holds for tuned mass dampers.
The difference between a TMD and vibration absorber is that with a TMD we are
concerned with the steady-state response as well as the transient response. With a vibration
absorber, the focus is on the steady-state response only. The primary mass may acquire an
undesired amplitude after periodic or aperiodic excitation. The TMD shortens the amount
of time it takes for the system to return to rest.

Example 7.22
Consider a primary mass with ωn = 1 and a TMD designed for resonance with mr /m = 0.05.
The damping factor of the TMD, if it were vibrating by itself, is ζ = 0.1. Obtain the
eigensolution of the primary mass with the added TMD. Analyze the damping properties
of the resulting system by means of the approximate solution in Section 7.12 and estimate
the resulting damping factors.
The following relationships and design criteria are needed for the primary mass and
TMD:
k kr cr mr
= ωn2 = ωn2 = 2ζωn = e = 0.05 [a]
m mr mr m
so that
cr
= 2eζωn [b]
m
Dividing the equations of motion given by Equation (7.169) by the mass m and intro-
ducing the expressions in Equations [a] and [b] gives
ẍa (t) + 2eζωn ẋa (t) − 2eζωn ẋr (t) + ωn2 (1 + e) xa (t) − ωn2 exr (t) = ky (t) = kAeiωt [c]
and
eẍr (t) − 2eζωn ẋa (t) + 2eζωn ẋr (t) − ωn2 exa (t) + ωn2 exr (t) = 0 [d]
Introducing the vector {x (t)} = [xa (t) xr (t)]T , we write the equations of motion in
matrix form, with the coefficient matrices
     
1 0 1 −1 2 1 + e −e
[m] = [c] = 2eζωn [k] = ωn [e]
0 e −1 1 −e e
We use the approximate method in Section 7.12 to solve for the eigensolution. To this
end, first the natural frequencies and modal vectors of the corresponding undamped case
are obtained by solving the eigenvalue problem
ω 2 [m] {u} = [k] {u} [f ]
Introducing the values ωn = 1 and e = 0.05, the eigensolution of the undamped system
can be shown to be
   
0.6667 0.7454
ω12 = 0.8 ω22 = 1.25 {u1 } = {u2 } = [g]
3.3333 −2.9814
364 Applied Dynamics

Let us examine the nature of the eigensolution. Recall that the natural frequencies of
the primary mass by itself and the TMD by itself are both 1. The natural frequencies of
the combined system are
√ √
ω1 = 0.8 = 0.8944 rad/sec ω2 = 1.25 = 1.118 rad/sec [h]

which are close to the previous natural frequencies. Examination of the modal vectors
indicates that the second element in each modal vector, which corresponds to the amplitude
of the TMD, is larger than the first. This is expected, as the mass of the TMD is much
smaller than the primary mass so that the TMD vibrates more.
Using the results of Section 7.12, the modal damping matrix, given by Equation (7.124),
becomes  
0 T 0.0711 −0.0994
[c ] = [u] [c] [u] = [i]
−0.0994 0.1389
Neglecting the off-diagonal terms leads to the approximate damping factors

2ω1 ζ1 = c011 = 0.0711 2ω1 ζ1 = c022 = 0.1389 [j]

Solving for the damping factors gives

c011 0.0711 c022 0.1389


ζ1 = = = 0.0397 ζ2 = = = 0.0621 [j]
2ω1 2 × 0.8944 2ω2 2 × 1.118
As a result of the coupling between the primary mass and TMD, the damping factor
ζ = 0.1 of the (standalone) TMD results in two modal damping factors, both smaller than
0.1. The sum of the two damping factors is close to 0.1. In essence, the damping introduced
by the TMD is split between the two modes. The resulting modal damping factors determine
the decay rates of the combined system. Also, as discussed in Section 7.12, the first mode
has a smaller damping factor than the second.

7.15 First-Order Systems


First-order differential equations that describe dynamical systems are mainly encountered
under the following conditions:

• A set of n second order differential equations is transformed into 2n first-order differ-


ential equations. This is done primarily for the purpose of numerical integration, as
most software for integrating ordinary differential equations is for first-order differential
equations. Numerical integration is discussed in the next section.
• In certain dynamical systems, such as Example 4.6, and in the study of vehicle dynamics
in Chapter 14, velocity ratios are used as variables. First-order models are also commonly
encountered in electrical engineering.
Consider a first-order system described by m first-order differential equations in the form

{ẋ (t)} + [A] {x (t)} = {f (t)} (7.173)

in which {x (t)} is a vector of order m containing the system variables, [A] is a matrix
of order m × m, and {f (t)} is a vector of external excitations or inputs. A solution will
Response of Multi-Degrees-of-Freedom Systems 365

be sought by considering the free response, that is, by setting {f (t)} = {0}. Assuming a
solution in the form

{x (t)} = {w} eλt (7.174)

and introducing the above equation into the describing equation with {f (t)} = {0} gives

(λ[1] + [A]) {w} = {0} (7.175)

where [1] is the identity matrix. Following the discussion earlier in this chapter regarding
nontrivial solutions, the coefficient matrix in the above equation must be singular, which
leads to the relationship

det (λ[1] + [A]) = 0 (7.176)

The above equation is recognized as the characteristic equation, a polynomial of order


m in λ. Solution of the characteristic equation gives the m eigenvalues λj and associated
eigenvectors {wj } (j = 1, 2, . . . , m). The eigenvalues can be real or imaginary. When imag-
inary, they occur in complex conjugate pairs. If any of the eigenvalues are real positive or
complex with positive real parts, the system is unstable.
The associated eigenvectors are real or imaginary, depending on the eigenvalues. The
eigenvector corresponding to a real eigenvalue is real and the eigenvector corresponding to
a complex eigenvalue is complex. The response can be expressed as
m
X
{x (t)} = {wj }eλj t (7.177)
j=1

and the same approach that was used earlier in this chapter can be used to deal with the
complex eigenvalues.
Because matrix [A] is not symmetric, we must also solve the left eigenvalue problem in
order to show orthogonality. What was solved above is the right eigenvalue problem. The
left eigenvalue problem is for the eigenvalues and eigenvectors of [A]T in the form

λ[1] + [A]T {z} = {0}



(7.178)

We can show that the eigenvalues of the right and left eigenvalue problem are the same.
But the left eigenvectors, denoted by {z1 }, {z2 }, . . . , {zm } are different from the right
eigenvectors {wj } (j = 1, 2, . . . , m). The orthogonality conditions are

{zj }T {wk } = 0, {zj }T [A] {wk } = 0 when j 6= k (7.179)

and a widely used normalization method is

{zj }T {wj } = 1 j = 1, 2, . . . , m (7.180)

which results in the relationship {zj }T [A]{wj } = λj . First-order systems will be analyzed
in more detail when studying vehicle dynamics and vehicle stability in Chapter 14.

7.16 Numerical Integration


An important tool for obtaining the response of a system is numerical integration. De-
velopments in computer hardware and software have led to the proliferation of numerical
366 Applied Dynamics

techniques. This section discusses numerical integration of the equations of motion of multi-
degrees-of-freedom systems.
We will consider two approaches. One is to directly integrate the original equations of
motion, and the other is to integrate the modal equations. Direct integration is simpler to
implement while integrating modal equations of motion is computationally more efficient.
Another advantage of integrating the original equations of motion is that it can be used to
integrate nonlinear equations, without the need for linearization.
Numerical integration software requires that the differential equations be written in state
form. As discussed in Section 1.9, in state form, the differential equations are of order one,
there is a single derivative on the left side of the equations, and there are no derivatives on
the right side.
Let us cast the equations of motion [m] {ẍ (t)} + [c] {ẋ (t)} + [k] {x (t)} = {F (t)} in state
form. We introduce the vector variables of order n
{y1 (t)} = {x (t)} {y2 (t)} = {ẋ (t)} (7.181)
and write their derivatives as
{ẏ1 (t)} = {ẋ (t)} = {y2 (t)} (7.182)
−1 −1 −1
{ẏ2 (t)} = {ẍ (t)} = − [m] [c] {ẋ (t)} − [m] [k] {x (t)} + [m] {F (t)}
−1 −1 −1
= − [m] [k] {y1 (t)} − [m] [c] {y2 (t)} + [m]
{F (t)} (7.183)
h i
T T
Introducing the vector {y (t)} of order 2n in the form {y (t)}T = {y1 (t)} | {y2 (t)} ,
h i
T T T
and the vector {F (t)} = {F (t)} | {0 (t)} , we write the first-order differential equations
as
{ẏ (t)} = [A] {y (t)} + [B] {F 0 (t)} (7.184)
where [A] and [B] are partitioned matrices of order 2n × 2n and 2n × n, respectively, in the
form
" # " #
[0] [1] [0]
[A] = −1 −1 [B] = −1 (7.185)
− [m] [k] − [m] [c] − [m]
The initial conditions can be expressed as
{x (0)} {y1 (0)}
   
{y (0)} = = (7.186)
{ẋ (0)} {y2 (0)}
When numerically integrating modal equations, we need to
1. Find eigenvalues and eigenvectors and normalize the eigenvectors;
2. Find modal initial conditions qj (0) , q̇j (0) as well as modal forces Qj (t) (j =
1, 2, . . . , n);
3. Write each modal equation in state form. Introducing the variables yj (t) = qj (t) and
yn+j (t) = q̇j (t), the modal equations in state form are
ẏj (t) = q̇j (t) = yn+j (t)

ẏn+j (t) = q̈j (t) = −2ζj ωj yn+j (t) − ωj2 yj (t) + Qj (t) (7.187)

4. Numerically integrate these n differential equations of order two;


Pn
5. Use the expansion theorem {x (t)} = j=1 qj (t) {uj } to obtain the response.
Response of Multi-Degrees-of-Freedom Systems 367

Example 7.23
Consider the cart-rod system in Figure 7.1 and write the nonlinear equations of motion in
state form.
This system has two degrees of freedom (n = 2). From Example 7.2, the equations of
motion are given in Equations [l] and [n] and are repeated here as

(M + m) ẍ + mLθ̈ cos θ − mLθ̇2 sin θ + cẋ + kx = F [a]

4
mL2 θ̈ + mLẍ cos θ + mgL sin θ = 0 [b]
3
In state form, 2n = 4 state variables are needed. A convenient way to select them is

y1 = x y2 = ẋ y3 = θ y4 = θ̇ [c]

Two of the state equations are easy to ascertain:

ẏ1 = ẋ = y2 ẏ3 = θ̇ = y4 [d]

To find the other two state equations, note that both Equations [a] and [b] have second
derivatives in two variables, so we cannot directly write these equations in state form.
Introducing the vector [y2 y4 ]T = [ẋ θ̇]T , we can write Equations [a] and [b] as

mLy42 sin y3 − cy2 − ky1 + F


    
M + m mL cos y3 ẏ2
4 2 = [e]
mL cos y3 3 mL ẏ4 −mgL sin y3

The expressions for the derivatives of y2 and y4 are obtained by inverting Equation [e].
Because the coefficient matrix involved is 2 by 2, the inversion can be done in closed form
−1 4 2
−mL cos y3
  
M +m mL cos y3 1 3 mL
4 2 = [f ]
mL cos y3 3 mL
∆ −mL cos y3 M +m

where
4
∆ = (M + m) mL2 − m2 L2 cos2 y3 [g]
3
is the determinant of the matrix. The other two state equations become
4 2
−mL cos y3 mLy42 sin y3 − cy2 − ky1 + F
    
ẏ2 1 3 mL
= [h]
ẏ4 ∆ −mL cos y3 M +m −mgL sin y3

The two state equations in Equation [d] and the two state equations in Equation [h]
constitute the four state equations for the system, cast in a form that is ready for numerical
integration.

7.17 Bibliography
Benaroya, H., and Nagurka, M., Mechanical Vibration, 3rd Edition, CRC Press, 2009.
Bottega, W.J., Engineering Vibrations, CRC Press, 2006.
Greenberg, M.D., Advanced Engineering Mathematics, 2nd Edition, Prentice-Hall, 1998.
Meirovitch, L., Fundamentals of Vibrations, Waveland Press, 2010.
368 Applied Dynamics

7.18 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 7.2—Modeling of Multi-Degrees-of-Freedom Systems


7.1 (M) Consider the schematic of a payload on a truck in Figure 6.48. The payload is
relatively heavy and the motion of the truck is affected by the payload. Using the position
of the truck y and the position of the payload on the truck x as variables, obtain the
equations of motion and cast them in matrix form. A force F acts on the truck in the
horizontal direction.

$! $#
"#
"! ""
!! !#

%! %#

FIGURE 7.30
Figure for Problems 7.2, 7.35, 7.37, and 7.38.

7.2 (M) For the mass-spring system in Figure 7.30, obtain the equations of motion in terms
of m1 , m2 , k1 , k2 , k3 , and c. Write the equations of motion in matrix form.

$% &%
" " )./+
!
#
ș ș
' ,
% $
)./+ +"# & &- &
&
0
! !
' )*!

(&

FIGURE 7.31
Figures for a) Problem 7.3, b) Problems 7.4 and 7.34.

7.3 (M) Obtain the equations of motion of the system in Figure 7.31a, using the coordinates
x and θ. Both coordinates have the value of zero when the springs are unstretched. Then,
Response of Multi-Degrees-of-Freedom Systems 369

linearize the equations of motion about x = 0, θ = 0. Write the linearized equations in


matrix form and check sign definiteness of the coefficient matrices.
7.4 (M) Obtain the equations of motion of the inverted pendulum in Figure 7.31b, using
the coordinates x and θ. Both coordinates have the value of zero when the springs are
unstretched. Then, linearize the equations of motion about x = 0, θ = 0. Write the linearized
equations in matrix form and check sign definiteness of the coefficient matrices.

g
k 2k Undeformed
x

G B
A  2m, L
k

s m

FIGURE 7.32
Figure for Problem 7.5.

7.5 (M) Obtain the equations of motion of the system in Figure 7.32. The springs are
unstretched when the coordinates x, θ, and s all have the value of zero; s is the deflection of
the lower spring. The mass m moves only vertically and the springs deflect only vertically.
Then, linearize the equations of motion about x = 0, θ = 0, and s = 0. Write the linearized
equations in matrix form and check sign definiteness of the coefficient matrices.
7.6 (E) Determine if the matrices below are positive definite.
 
6 2 −3  
4 −5
[A] =  2 3 −1  [B] =
−5 6
−3 −1 4

7.7 (M) For what values of x will the matrix [A] below be positive definite?
 
4 x −3
[A] =  x x −2 
−3 −2 5

7.8 (E) The stiffness matrix of a three-degrees-of-freedom mass-spring system is


 
k1 + k2 + k4 −k2 −k4
[k] =  −k2 k2 + k3 + k5 −k3 
−k4 −k3 k3 + k4

Sketch the mass-spring system.


7.9 (E) Write the stiffness matrix for the mass-spring-damper system in Figure 7.33 directly
from the figure, without drawing a free-body diagram.
7.10 (E) Write the stiffness matrix for the mass-spring system in Figure 7.34 directly from
the figure, without drawing a free-body diagram.
370 Applied Dynamics
"% "&

"! "# "" "$


!! !# !" !$

"'

FIGURE 7.33
Figure for Problem 7.9.

k6 k7

k1 k2 k3 k4 k5
m1 m2 m3 m4

x1 x2 x3 x4

FIGURE 7.34
Figure for Problems 7.10, 7.15, and 7.30.

Section 7.4—Free Motion of Undamped Multi-Degrees-of-Freedom Systems


7.11 (M) Obtain the natural frequencies and modal vectors of the truck and payload system
in Problem 7.1 for M = 5500 kg, m = 800 kg, k = 100 N/cm, c = 0. Plot the modal vectors.
7.12 (M) For the mass-spring system in Problem 7.2, obtain the natural frequencies and
modal vectors for m1 = m2 = m, k1 = k2 = k3 = k, c = 0, and plot the modal vectors.
7.13 (M) Obtain the natural frequencies and modal vectors of the system in Problem 7.3
for m = 2 kg, c = 0, k = 40 N/m, and L = 0.8 m. Plot the modal vectors.
7.14 (C) Obtain the natural frequencies and modal vectors of the system in Problem 7.9 for
mi = m, ci = 0 and ki = k in terms of m and k. Plot the modal vectors and observe the
sign changes.
7.15 (C) Write a computer program to calculate the natural frequencies and modal vectors
of the system in Figure 7.34 for ki = 15 − i lb/in. (i = 1, 2, 3, 4) and mi = 0.2 + 0.05i slugs.

Section 7.5—Solving for the Natural Frequencies and Modal Vectors


7.16 (M) Consider a two-degrees-of-freedom representation of a dynamical system in terms
of a mass and stiffness matrix of the form
   
m1 0 k11 k12
[m] = [k] =
0 m2 k12 k22
Obtain the squares of the natural frequencies, ωj2 , and modal vectors {uj } in terms of
m1 , m2 , and kij (i, j = 1, 2).
7.17 (M) Consider a mass-spring system of the form shown in Figure 7.20. You are given
that m1 = 400 kg, m2 = 40 kg and k3 = 0. Design k1 and k2 so that the natural frequencies
of the system are ω1 = 6 rad/s and ω2 = 60 rad/s.
Response of Multi-Degrees-of-Freedom Systems 371

7.18 (C) Plot the natural frequencies of the truck and payload system in Problem 7.1 as
a function of the weight of the truck. Consider the range of 500–8,000 lb. Based on the
natural frequencies you obtain, can you identify a certain weight of the truck after which
the motion of the truck is not affected by the motion of the payload?
7.19 (M) Solve for the eigenvalues and modal vectors of the mass-spring system in Figure
7.4 for the following parameters: m1 = m2 = m, m3 = 2m, k1 = k2 = k, k3 = k4 = 2k,
ci = 0 i = 1, 2, 3, 4. Solve the eigenvalue problem first by hand (the numbers are such that
you can do so without difficulty) and then by computer.
7.20 (C) Consider the two-degrees-of-freedom quarter-car model in Figure 15.11. Given the
values of mS = 400 kg, mU = 40 kg, kS = 17 N/mm, calculate the natural frequencies
in terms of the tire stiffness kt . Consider the range of 150–200 N/mm for kt . Calculate
∂ω1 /∂kt . Plot the natural frequencies as a function of kt .

Section 7.6—Beat Phenomenon


7.21 (M) Consider the two mass-spring system in Figure 7.30. The masses are the same
(m1 = m2 = m) and so are the outer springs k1 = k3 = k connecting the masses to the
fixed supports. The middle spring (k2 ), which connects the two masses is much weaker than
the two outer springs. Denoting the middle spring’s constant by k2 = k, where  is small,
calculate the natural frequencies and modal vectors and show that the system exhibits beat
phenomenon (show similarities of eigenvectors and closeness of natural frequencies).
7.22 (C) Consider Problem 7.21. Plot the natural frequencies as a function of the ratio of
the springs e = k2 /k and ascertain when the beat phenomenon can no longer be observed
easily. Assume that beat phenomenon cannot be clearly identified if the ratio of the two
natural frequencies exceeds 1.4.
7.23 (M) Derive the expression for energy for the beat phenomenon problem in Section 7.6
and write it in terms of the initial angle θ0 .

Section 7.7—Unrestrained Motion and Rigid Body Modes

k1 k2 k3
F1 F2
m1 m2 m3 m4

x1 x2 x3 x4

FIGURE 7.35
Figure for Problem 7.24.

7.24 (E) For the four-degrees-of-freedom mass-spring system in Figure 7.35, verify that the
modal vector corresponding to the rigid body mode is {u1 } = d1 [1 1 1 1]T , where d1 is an
arbitrary constant. Do this by calculating {u1 } from the eigenvalue problem.
7.25 (M) Consider the three mass-spring-damper system in Figure 7.4. Write the equations
of motion. Then, obtain the natural frequencies and modal vectors for m1 = m2 = m3 = m,
k1 = k4 = 0, k2 = k3 = k, c1 = c2 = c3 = c4 = 0. Note: For this problem the eigensolution
can be obtained by hand.
372 Applied Dynamics

Section 7.8—Orthogonality of the Modal Vectors


7.26 (M) Calculate the modal vectors in Problem 7.11 for c = 0 and verify that they are
orthogonal.
7.27 (M) Verify that the modal vectors are orthogonal in Problem 7.2 using the values
m1 = m2 = m, k1 = k2 = k3 = k, c = 0. Normalize the modal vectors using the second
convenience.
7.28 (C) Write a computer program to obtain the natural frequencies and modal vectors of
the four mass-spring system in Problem 7.9 for ki = 18 − i N/m and mi = 3i kg. Verify
that the modal vectors are orthogonal and normalize them using the second convenience.

Section 7.9—Expansion Theorem


7.29 (M) Given the three-degrees-of-freedom system in Example 7.11, expand the following
vectors in terms of the modal vectors:
     
1 1 1
{v1 } =  2  {v2 } =  −3  {v3 } =  1 
0 1 −2

7.30 (C) Given the four-degrees-of-freedom system in Problem 7.15, expand the following
vectors in terms of the modal vectors:
     
1 1 1
2  −3   −1 
{v1 } = 
3
 {v2 } = 
 1
 {v3 } = 
 2

4 2 −2

Section 7.10—Modal Equations of Motion and Response


7.31 (E) Consider the mass-spring system in Figure 7.23, with m1 = m2 = 1 slug, k1 = 30
lb/ft, k2 = 20 lb/ft, k3 = 0, c1 = c2 = c3 = 0, F1 = F2 = 0. The initial conditions are
x1 (0) = x2 (0) = 0, ẋ1 (0) = 3 in./sec, ẋ2 (0) = 0. Obtain the response of each mass.

k F
W1 W2

x1 x2

FIGURE 7.36
Figure for Problem 7.32.

7.32 (M) Consider the unrestrained two mass-spring system in Figure 7.36. Obtain the
response for the case W1 = W2 = 10 lb, k = 12 lb/in, F = 0.2t lb, x1 (0) = x2 (0) = 0,
ẋ1 (0) = 0, ẋ2 (0) = −0.2 ft/sec.
7.33 (M) Obtain the response of the mass-pendulum problem about stable equilibrium
(Example 7.7) to a horizontal impulsive force F̂ applied to the cart. Before the impulsive
force is applied, the system is at rest with x = 0, θ = 0.
7.34 (M) The inverted pendulum in Figure 7.31b is used to simulate an individual driving
a vehicle, with the cart as the body of the driver and the rod as the head. During whiplash,
the vehicle is rear-ended by another vehicle and the driver’s head moves backward (this is
Response of Multi-Degrees-of-Freedom Systems 373

why headrests are necessary equipment in vehicles). Modeling the accident as an impulsive
force F̂ = 20 N·s acting on the mass M = 70 kg, and the driver’s head as m = 5 kg and
L = 25 cm, calculate the angular velocity of the head immediately after the impulse.
7.35 (M) Consider the mass-spring system in Figure 7.30, with m1 = m2 = m, k1 = k2 =
k3 = k, c1 = c2 = 0. The initial conditions are x1 (0) = −2, x2 (0) = 0, ẋ1 (0) = ẋ2 (0) = 0.
The forcing is F1 (t) = 0, F2 (t) = 0.1t. Find the response of each mass.

Section 7.11—Mode Participation and Isolation


7.36 (M) Consider the mass-spring system in Problem 7.12. Neglecting the damping, it is
desired to impart forces F1 and F2 such that the second mode is not excited. Obtain the
relationship between F1 and F2 that will make this possible.
7.37 (M) Consider the mass-spring system in Figure 7.30, with m1 = m2 = m = 2, k1 =
k2 = k3 = k = 8, c = 0. System measurements show that the time histories of the two masses
are x1 (t) = 3 sin 2t + 1.2 sin 6t, x2 (t) = −1.9 sin 2t + 0.4 sin 6t. Find the time histories of
the modes.
7.38 (D) Consider the mass-spring system in Figure 7.30, with m1 = m2 = m = 2, k1 =
k2 = k3 = k = 8, c = 0. It is desired to impart a modal force to the second mode in the
form Q2 (t) = −0.2q̇2 (t), while Q1 = 0. Calculate the forces Fi (ẋ1 (t) , ẋ2 (t)) (i = 1, 2) that
will make this control action possible.

Section 7.12—Approximate Approach for Damped Systems


7.39 (E) Given a two-degrees-of-freedom system whose coefficient matrices are
     
1 0 1 2 1.2 0.8
[m] = [k] = [c] =
0 2 2 5 0.8 3.6

determine if the damping matrix represents proportional damping.


7.40 (M) Consider the mass-spring system in Figure 7.23, with m1 = m2 = 1 slug, k1 = 25
lb/ft, k2 = 20 lb/ft, c1 = 2 lb·sec/ft, c2 = 0. Calculate the damping factors for the two
modes using the approximate approach.
7.41 (M) Consider the mass-spring system in Figure 7.23, with m1 = m2 = 11 kg, k1 = 250
N/m, k2 = 200 N/m, c1 = 28 N·s/m, c2 = 20 N·s/m. Calculate the damping factors for the
two modes using the approximate approach.
7.42 (M) Consider the mass-spring system in Example 7.20. It is desired to give the first
mode a damping factor of ζ1 = 0.1 by means of a damper c2 . What should be the value of
c2 be so that the approximate value for ζ1 = 0.1? Take c1 = 0. Calculate the value of ζ2
that results.
7.43 (C) Solve, using the approximate approach, for the natural frequencies and damping
factors of the mass-spring-damper system in Figure 7.4 for the following parameters: mi = 5
kg, ki = 20 N/m, ci = 2 + 2i N·s/m.
7.44 (D) Consider the mass-spring system in Figure 7.23, with m1 = m2 = 1 slug, k1 = 30
lb/ft, k2 = 20 lb/ft. It is desired to give the first mode a damping factor of ζ1 = 0.1 and
the second mode a damping factor of ζ2 = 0.15 by means of two dampers c1 and c2 and by
using proportional damping. Calculate the damping constants c1 and c2 that will lead to
the desired damping factors.

Section 7.13—Response to Harmonic Excitation


7.45 (M) The cart-rod system about stable equilibrium in Example 7.7 is subjected to a
harmonic excitation of F (t) = 3 cos 2t. Find the steady-state amplitudes of the cart and
rod using the impedance matrix.
374 Applied Dynamics

7.46 (M) The cart-rod system about stable equilibrium in Example 7.7 is subjected to a
harmonic excitation of F (t) = 3 cos 2t. Find the steady-state amplitudes of the cart and
rod using modal analysis.
7.47 (C) Consider the mass-spring system in Example 7.20. The second mass is now acted
upon by a harmonic force F = 100 sin ωt, where F0 = 104 N. You are asked to develop
frequency response charts, plotting X1 and X2 , where xj (t) = Xj eiωt , (j = 1, 2), vs ω.
Using the impedance matrix, plot X1 and X2 for 0 ≤ ω ≤ 4 rad/s and for the following
damping constants: a) c1 = 0, c2 = 1 N·s/m; b) c1 = c2 = 2 N·s/m; c) c1 = 3, c2 = 0 N·s/m.
Create two charts, one for X1 and one for X2 and include all three cases on the same chart.

Section 7.14—Vibration Reducing Devices


7.48 (M) Given a mass-spring system with mass 3 kg and stiffness 12 N/m, resting on a base
that undergoes harmonic motion Aeiωt , design a vibration absorber so that the amplitude
of the primary mass will be zero at resonance. Use a mass ratio for the absorber that is
less than 0.04. Compare the amplitude ratio of the primary mass before and after absorber
is added (Xa /A and Xb /A) when the excitation frequency is ω = 0.95ωn , where ωn is the
natural frequency of the primary mass before the absorber.
7.49 (M) Consider the cart-rod system about stable equilibrium. The cart has mass M = 4
kg and spring k = 8 N/m. It is being subjected to a harmonic force F (t) = F0 eiωt . We
want to design the rod as a vibration absorber pendulum. Develop the design criteria for
selecting the mass m and length L of the pendulum so that the cart’s motion amplitude
will be zero when it is excited by the resonance frequency of the cart-spring system.
7.50 (M) Design a vibration absorber for a system of primary weight W = 400 lb and natural
frequency ωn = 1.8 rad/sec. The absorber should be effective in the range 1.7 ≤ ω ≤ 1.9
rad/sec, where ω is the excitation frequency. What is the minimum weight of the absorber
that can achieve the desired range of effectiveness?
7.51 (D) Given mass-spring system with mass 3, 000, 000 kg and stiffness 300, 000 N/m,
resting an a base that undergoes harmonic motion, design a tuned mass damper. Following
are requirements: a) Use a mass ratio of 0.04. b) Minimum damping factors for the resulting
two-degrees-of-freedom system are ζ1 > 0.09, ζ2 > 0.11. c) TMD is designed for resonance.
Your answer should be the mass, stiffness, and damping constant of the TMD. You can use
the approach in Section 7.12 to calculate the damping factors for the two-degrees-of-freedom
model that results.

Section 7.15—First-Order Systems


7.52 (C) Write the equations of motion of the mass-spring-damper system in Figure 7.4.
Then, write these equations as a set of six first-order equations and solve for the eigenvalues.
Use the following parameters: mi = 5 kg, ki = 20 N/m, ci = 2 + 2i N·s/m. Compare your
results with the results of Problem 7.43.

Section 7.16—Numerical Integration


7.53 (C) Write the equations of motion of the mass-spring-damper system in Figure 7.4.
Then, cast these equations into state form and numerically integrate for the initial condition
x1 (0) = 0.2 m, with all other initial displacements and initial velocities being zero. Use the
following parameters: mi = 5 kg, ki = 20 N/m, ci = 2 + 2i N·s/m. Plot the positions of
the masses over time. Then, carry out the integration when the damping values are halved,
and ci = 1 + i N·s/m, and compare the responses of the masses.
8
Analytical Mechanics

8.1 Introduction
This chapter develops analytical techniques for obtaining the describing equations of dynam-
ical systems. The analytical approach differs from the Newtonian approach in that the dy-
namical system is considered as a whole, rather than breaking the system into components.
Analytical techniques incorporate scalar quantities such as energy and work. Constraint
forces and moments are treated differently from those in Newtonian mechanics. Constraint
forces that do no work do not appear in the formulation, as they are accounted for by an
appropriate selection of motion variables. Because of this property, analytical mechanics is
more advantageous than using Newtonian mechanics for systems in which the number of
components is larger than the degrees of freedom. On the other hand, the designer of a
device or a machine needs to know the magnitudes of the reaction and constraint forces. It
is best if an engineer or scientist can comfortably use both analytical as well as Newtonian
techniques.
Analytical mechanics makes use of concepts from variational calculus. Generalized co-
ordinates, which are combinations of physical coordinates, are used as motion variables,
making the analytical approach more flexible than Newtonian mechanics.
The chapter begins with a discussion of generalized coordinates, generalized velocities,
and generalized speeds. Kinematic variables and virtual displacements are introduced. The
principle of virtual work for static equilibrium is developed. D’Alembert’s principle is derived
as the extension of the principle of virtual work to the dynamic case. The Hamilton’s
principle, Lagrange’s equations, and Kane’s equations follow from D’Alembert’s principle.
A discussion of natural and nonnatural systems and equilibrium is presented. Damped
systems and gyroscopic motion is analyzed.
The equations of analytical mechanics are valid for particles, rigid bodies, or deformable
bodies. Illustrative examples in this chapter involve particles or plane motion of rigid bod-
ies. The kinetics of rigid bodies undergoing three-dimensional motion will be discussed in
Chapter 11.

8.2 Generalized Coordinates and Constraints


The concept of degree of freedom was introduced in Chapter 1. The number of degrees of
freedom (d.o.f.) of a system is the minimum number of independent variables required to
describe the motion of the system completely. In general, the number of degrees of freedom
is denoted by n. This section discusses the variables that describe motion.

375
376 Applied Dynamics

8.2.1 Generalized Coordinates


Once the number of degrees of freedom n is ascertained, the next step is to select the motion
variables. A set of variables that can completely describe the position of a system is called
generalized coordinates. The space spanned by the generalized coordinates is called the
configuration space. In general, the number of generalized coordinates will be denoted by
m, (m ≥ n) and expressed as q1 , q2 , . . . , qm . For example, the position vector of a point
is written as

r = r (q1 , q2 , . . . , qm , t) (8.1)

 

 

FIGURE 8.1
A spherical pendulum.

As an illustration, consider the spherical pendulum in Figure 8.1, whose length can
change. The pendulum has three degrees of freedom and its motion can be described by the
Cartesian coordinates x, y, and z, or spherical coordinates q1 , q2 , and q3 , where q1 = L
describes the length of the pendulum and q2 = θ and q3 = φ describe the polar and
azimuthal angles, respectively. The two sets of coordinates are related by

x = L cos θ sin φ y = L sin θ sin φ z = −L cos φ (8.2)

Deciding which set of generalized coordinates, x, y, z, or L, θ, φ, to use is the same


issue as selecting a coordinate system, a topic discussed in Chapter 2. In general, we select
the generalized coordinates so that the description of the motion is more meaningful and
the resulting equations of motion are simpler.
While for the pendulum above it is a simple choice to select for the number of generalized
coordinates the same as the number of d.o.f., m = n, the choice may not be as simple in
other cases. Consider the four-bar linkage in Figure 8.2, consisting of four links (as discussed
in Chapter 3; the fixed body to which the linkage is attached is referred to as the grounded
link, link 1). This linkage has one degree of freedom. The three angles, θ2 , θ3 , and θ4 , are
related to each other by two constraints. These two constraints can be written by considering
the coordinates of the joint that connects links 3 and 4, point P , as

In horizontal direction: L2 cos θ2 + L3 cos θ3 = L1 + L4 cos θ4

In vertical direction: L2 sin θ2 + L3 sin θ3 = L4 sin θ4 (8.3)


Analytical Mechanics 377

P
m3 , L3
Q
3

m2 , L2 m4 , L4

O 2 O4 4

L1

FIGURE 8.2
A four-bar linkage.

Writing equations similar to Equation (8.3) for the coordinates of the joint that connects
links 2 and 3, point Q, does not describe any additional relationships, as these equations
can be derived by rearranging the terms in Equation (8.3). We may think that it is easier to
obtain the describing equation of such a linkage by using one motion variable, say, θ2 , but
it turns out that the algebra involved becomes quite complex when a single motion variable
is used.
It is instructive to distinguish between two sets of generalized coordinates:

• Independent generalized coordinates, where there as many generalized coordinates


as there are degrees of freedom, that is, m = n.
• Constrained generalized coordinates, or dependent generalized coordinates, where
the number of generalized coordinates is larger than the number of degrees of freedom,
that is, m > n.
We cannot have a proper set of generalized coordinates when m < n. For the most time, we
deal with independent generalized coordinates. There needs to be a compelling reason (such
as in the four-bar linkage above) to prefer a set of constrained generalized coordinates.
When selecting the generalized coordinates, avoid ambiguous coordinates, which intro-
duce ambiguity to the description of motion. For example, the double link in Figure 8.3
has two degrees of freedom. The coordinates xP and yP of the endpoint P of the double
link constitute a set of ambiguous coordinates, as a given position of the endpoint can be
reached by two different configurations of the links.

8.2.2 Constraints and Constraint Forces


Chapter 1 discussed the basic concept of a constraint and distinguished between holonomic
and nonholonomic constraints. A more formal definition is introduced here of a constraint
and the associated constraint force.
A constraint acting on a system is, in essence, a restriction on the motion. To formally
describe a constraint, we need two things: aconstraint equation, which is the geometric (or
kinematic) description of the constraint, and a constraint force that enforces the constraint.
The constraint equation and constraint force are complementary quantities, as discussed
within the context of the cause-and-effect principle in Chapter 1.
378 Applied Dynamics

xP
First configuration
O L1 x
2
1
1 L2
yP
L1 y

L2
P
2 Second configuration

FIGURE 8.3
xP and yP are ambiguous generalized coordinates. Point P can be reached two ways for a
given value of xP and yP .

This section discusses only equality constraints. Such constraints are usually written in
two forms:

• Constraint in configuration form. The general form of a constraint in configuration


form, also known as a configuration constraint, is

f (q1 , q2 , . . . , qm , t) = 0 (8.4)

and it describes a surface on which the motion takes place. If the constraint is an explicit
function of time, then the surface itself is moving. If not, the surface is fixed.
• Constraint in velocity form. A constraint in velocity form, also known as a velocity
constraint, is written as

a1 q̇1 + a2 q̇2 + . . . + am q̇m + a0 = 0 (8.5)

where q̇1 , q̇2 , . . . , q̇m are the derivatives of the generalized coordinates, referred to as
generalized velocities. A constraint equation in velocity form does not define a physical
surface.

Consider holonomic constraints first. A constraint is called holonomic if it can be ex-


pressed in the form of Equation (8.4) or, if expressed in the velocity form Equation (8.5),
it can be integrated to the form of Equation (8.4).
The motion of the body is always tangent to the surface defined by the holonomic
constraint (that is, of course, as long as the constraint is enforced). The constraint force is
normal to this surface as the constraint force prevents the body from piercing the constraint
surface. Recall from the definition of contact in Chapter 3 that two contacting bodies have
a common normal and a common tangent. The contact force is along the common normal,
which is the source of the expressions normal direction and normal force.
We can show that the unit vector in the normal direction n is related to the constraint
equation by
∇f
n = (8.6)
|∇f |
Analytical Mechanics 379

where ∇f is the gradient of f . For example, using the xyz coordinates, ∇f = ∂f ∂f ∂f


∂x i+ ∂y j+ ∂z k.
Also, because the velocity is tangent to the surface, we can write v · n = 0.
Figure 8.4 describes the geometry when the constraint is not an explicit function of
time (so the surface defined by the constraint is fixed). The constraint force has the form
F0 = F 0 n. The work done by the constraint force F0 , denoted by W 0 , is

t
t
F'
FIGURE 8.4
Constraint force is along the normal.

Z
W0 = F0 · v dt (8.7)

When the constraint is not an explicit function of time, the constraint surface is fixed,
and the constraint force and velocity are always perpendicular to each other. Hence, their
dot product vanishes
F0 · v = F 0 n · vt = 0 (8.8)
leading to the conclusion that the work done by a holonomic constraint that is independent
of time in any possible displacement is zero. Such constraints are referred to as workless
constraints.
When a holonomic constraint is an explicit function of time, the work performed by the
corresponding constraint force is not zero. The path followed by the body can no longer be
completely described by the path variables associated with the surface. The normal vector
n still describes the normal to the surface, but because the surface is moving, the total
velocity is the velocity along the path plus the velocity of the surface. Hence, the total
velocity is not normal to the surface and F0 · v 6= 0.
A nonholonomic constraint cannot be expressed in the form of Equation (8.4), so it
does not define a surface in the configuration space. Hence, we cannot derive a general
relationship between the constraint force and velocity, even though for most nonholonomic
equality constraints we can visually observe such a relationship.
Consider the bicycle model of a rear-wheel vehicle discussed in Chapters 1 and 3. The
vehicle and forces that act on it are shown in Figure 8.5. The vehicle has two velocity degrees
of freedom, speed and angular velocity. Four position variables are needed to describe the
geometry. They can be selected as XA , YA , θ, and δ, with δ being the steer angle.1 The
rolling constraints state that the velocities of the wheels are along the heading of the wheels.
Mathematically,
vA · j = 0 vB · j0 = 0 (8.9)

1 Other suitable choices of generalized coordinates are associated with the center of mass G or with point

B, and the associated sets would be XG , YG , θ, δ, or XB , YB , θ, δ.


380 Applied Dynamics
y
y'

y x'
L B  x
b
c Y
G y
A y' 
F'B 
F 
x'
Y  x
F'A YA 
XA X
X

FIGURE 8.5
Constraint forces for bicycle model.

To show that these two constraints are nonholonomic, we begin by writing the velocities
of points A and B. The unit vectors of the reference frames involved are related by

i = cos θI + sin θJ j = cos θJ − sin θI (8.10)

and the inverse relationship is

I = cos θi − sin θj J = cos θj + sin θi (8.11)

The velocity of point A can be written as


   
vA = ẊA I + ẎA J = ẊA cos θ + ẎA sin θ i + −ẊA sin θ + ẎA cos θ j (8.12)

Imposing the constraint that the velocity of A has a component only in the x direction, or
vA · j = 0, the constraint equation is obtained as

vA · j = −ẊA sin θ + ẎA cos θ = 0 (8.13)

This equation cannot be integrated to one in terms of the generalized coordinates, and
hence, is nonholonomic. We can find the constraint relationship for the front wheel, point
B, in a similar form. The derivation is left as an exercise.
The force F is the propulsive force, and the forces F0A and F0B are constraint forces. In
vector form, F0A = FA0 j, F0B = FB0 j0 . These two constraint forces (which are associated with
the nonholonomic constraints) are perpendicular to the velocities at all times, and hence,
they do no work, as long as the no-sliding conditions at A and at B are maintained.
The following general form is a convenient way to represent constraints. Consider n
degrees of freedom and m generalized coordinates, so that there are p = m − n constraints.
The most general form of equality constraints is in velocity form
m
X
ajk q̇k + aj0 = 0 j = 1, 2, . . . , p (8.14)
k=1

where ajk and aj0 (j = 1, 2, . . . , p; k = 1, 2, . . . , m) are functions of the generalized coordi-


nates and time. If a constraint equation in velocity form can be integrated to a constraint
Analytical Mechanics 381

equation in the form of Equation (8.4), then the constraint is holonomic.2 Otherwise, it is
nonholonomic. Unfortunately, a straightforward way to determine if a velocity constraint
is holonomic or not does not exist. An algebraically cumbersome procedure, involving inte-
grating factors, exists for determining if a velocity constraint is holonomic. The description
can be found in the texts by Baruh or Ginsberg.

Example 8.1

z
R

FIGURE 8.6
Bead going up a spiral.

A bead slides up a spiral of constant radius R and height h, as shown in Figure 8.6. It
takes the bead six full turns to reach the top. Express the characteristics of the path as a
constraint relationship.
In each revolution (2π radians) the bead goes up by a height of h/6. Denoting the vertical
direction by z and the angle traversed by θ, the relationship between z and θ becomes
h θ h
z (θ) = = θ [a]
6 2π 12π

8.3 Velocity Representation


The previous section discussed representation of the position of a point or of a body in terms
of the generalized coordinates and in the presence of constraints. This section discusses the
velocity vector and incorporation of nonholonomic constraints.

8.3.1 Representation by Generalized Velocities


The position vector, in terms of n generalized coordinates, is

r = r (q1 , q2 , . . . , qn , t) (8.15)

where an explicit dependence of r on time is allowed. Differentiating the above equation


with respect to time gives the velocity as
∂r ∂r ∂r ∂r
v = ṙ = q̇1 + q̇2 + . . . + q̇n + (8.16)
∂q1 ∂q2 ∂qn ∂t
2 Of, course, we can take a holonomic constraint equation that is in configuration form, differentiate it,

and cast it in velocity form.


382 Applied Dynamics
∂r
where the quantities q̇k (k = 1, 2, . . . , n) are generalized velocities and ∂q k
are known as
kinematic variables. Taking the partial derivative of the above equation with respect to
q̇k (k = 1, 2, . . . , n) yields the result
∂v ∂r
= k = 1, 2, . . . , n (8.17)
∂ q̇k ∂qk
which is an important relationship and one that we will use extensively in this chapter.

8.3.2 Representation by Generalized Speeds (Quasi-Velocities)


Equation (8.17) is in terms of the generalized velocities. The question arises whether it is
a restriction to use only generalized velocities and if it is possible to use another set of
variables to describe velocities and angular velocities. A hint to this appears in Chapter 2,
when discussing angular velocity in three dimensions. For general three-dimensional motion,
the angular velocity vector is a defined quantity and not the derivative of an expression.
Yet another hint appears in the previous section of this chapter, within the context of
nonholonomic constraints. For example, the vehicle in Figure 8.5 has two nonholonomic
constraints acting on it: the velocity of point A is along the x direction and the velocity of
point B is in the x0 direction. Using vA and/or vB as motion variables is a more meaningful
choice than using ẊA or ẎA .
We introduce a set of variables called quasi-velocities 3 or generalized speeds 4 as a set of
variables that are linear combinations of the generalized velocities and that can describe the
velocity of a system completely. Denoted by u1 , u2 , . . . , un , quasi-velocities are not neces-
sarily derivatives of any coordinates and thus cannot always be integrated to be expressed
in terms of generalized coordinates.
Consider a system described by m independent generalized coordinates that are related
to each other by p nonholonomic constraints. It follows that n = m − p independent velocity
variables are needed to describe the velocities. How many degrees of freedom are there in
a system acted upon by nonholonomic constraints? We can better answer this question by
revising the definition of degree of freedom to the minimum number of independent velocity
variables needed to describe motion completely.
For unconstrained systems that have n degrees of freedom, we need n independent
generalized coordinates and n generalized velocities or n generalized speeds. When a system
with m coordinates is acted upon by p nonholonomic constraints, then m independent
generalized coordinates and n = m−p independent generalized speeds are needed to describe
the motion.

8.3.3 Relationship between Generalized Velocities and Generalized


Speeds
Two issues are of interest: a) selection of independent generalized speeds, and b) the re-
lationship between the generalized velocities and generalized speeds. Beginning with the
latter, we can write
m
X
uk = Ykj q̇j + Zk k = 1, 2, . . . , n (8.18)
j=1

3 The source of the expression quasi-velocity is that we can think of it as the derivative of a fictitious

quantity, a quasi-coordinate.
4 There has been controversy in the literature regarding the names of these variables. This text uses the

terms generalized speed and quasi-velocity interchangeably.


Analytical Mechanics 383

where Ykj = Ykj (q1 , q2 , . . . , qm , t) and Zk = Zk (q1 , q2 , . . . , qm , t) (k = 1, 2, . . . , n; j =


1, 2, . . . , m) are functions of the generalized coordinates and time. In column vector form

{u} = [Y ] {q̇} + {Z} (8.19)

where [Y ] is a matrix of order n × m and {Z} is a vector of order n, {q̇} = [q̇1 q̇2 . . . q̇m ]T ,
and {u} = [u1 u2 . . . un ]T .
In order for the set of quasi-velocities to completely describe a system, the above equa-
tion must be invertible. For holonomic systems, that is, with n = m, [Y ] must be nonsin-
gular, and in the presence of nonholonomic constraints, [Y ] must have rank n. In addition,
for constrained systems, the generalized speeds must be selected such that the constraint
relation

[a] {q̇} + {b} = {0} (8.20)

which is the matrix representation of Equation (8.14), is satisfied, where [a] is the constraint
matrix of order p × m.
Assuming that the generalized speeds are selected appropriately, we can express the
generalized velocities in terms of the generalized speeds by

{q̇} = [W ] {u} + {X} (8.21)

where [W ] is a matrix of order m×n and {X} is a vector of order m. The above relationship
is also known as the set of kinematic differential equations that were discussed in Section
3.6. Introducing Equation (8.21) into Equations (8.19) and (8.20) leads to the result

{u} = [Y ] ([W ] {u} + {X}) + {Z} = [Y ] [W ] {u} + [Y ] {X} + {Z} (8.22)

[a] [W ] {u} + [a] {X} + {b} = {0} (8.23)

The above two equations describe the requirements that a properly selected set of inde-
pendent generalized speeds must satisfy. To satisfy Equations (8.22) and (8.23), the following
relationships must hold:

[Y ] [W ] = [1] [Y ] {X} + {Z} = {0} (8.24)

[a] [W ] = [0] [a] {X} + {b} = {0} (8.25)

As with generalized coordinates, there are no general guidelines for selecting the general-
ized speeds. For systems acted upon by nonholonomic constraints, velocities perpendicular
to the constraint forces are suitable choices. Components of angular velocity are usually
good choices, as well. Experience also helps.
Consider next representation of velocity in terms of generalized speeds and associated
kinematic variables. To this end, the kinematic variable ∂∂v
q̇j (j = 1, 2, . . . , m) can be written
as
∂v ∂v ∂u1 ∂v ∂u2 ∂v ∂un
= + + ... + j = 1, 2, . . . , m (8.26)
∂ q̇j ∂u1 ∂ q̇j ∂u2 ∂ q̇j ∂un ∂ q̇j

and, from Equation (8.18),

∂uk
= Ykj k = 1, 2, . . . , n, j = 1, 2, . . . , m (8.27)
∂ q̇j
384 Applied Dynamics

Using column vector notation, Equation (8.26) can be expressed as


∂v
= [vu ] {Yj } (8.28)
∂ q̇j

in which
 
u ∂v ∂v ∂v
[v ] = ... (8.29)
∂u1 ∂u2 ∂un
∂v
is an array of dimension 1 × n whose elements are algebraic vectors. Their entries ∂uk (k =
1, 2, . . . , n) are referred to as partial velocities with the notation
∂v
vk = (8.30)
∂uk
In addition, the vector

{Yj } = [Y1j Y2j . . . Ynj ]T (8.31)

can be recognized as the j-th column of [Y ], where [Y ] = [{Y1 } {Y2 } . . . {Ym }]. The partial
velocities are the counterpart of the kinematic variables when generalized speeds are the
motion variables. It follows that we can write the partial derivatives of v with respect to
the generalized velocities in the form
 
q ∂v ∂v ∂v
[v ] = ... = [vu ][Y ] (8.32)
∂ q̇1 ∂ q̇2 ∂ q̇m

Now, rewrite Equation (8.16) as


∂r
v = [vq ] {q̇} + (8.33)
∂t
and introduce Eqs. (8.21) and Equation (8.32) into Equation (8.33), with the result
∂r
v = [vu ] [Y ] ([W ] {u} + {X}) + = [vu ] {u} + vt (8.34)
∂t
in which vt = [vu ] [Y ] {X} + ∂r
∂t is known as the time partial velocity. From the above
equation, the velocity can be written in terms of the quasi-velocities as
∂v ∂v ∂v
v = u1 + u2 + . . . + un + vt = v1 u1 + v2 u2 + . . . + vn un + vt (8.35)
∂u1 ∂u2 ∂un
In a similar fashion, we can obtain the kinematic parameters associated with angular ve-
locity. The angular velocity of a body can be expressed in terms of the generalized velocities
as5
∂ωω ∂ωω ∂ωω
ω = q̇1 + q̇2 + . . . + q̇m + ω ∗ (8.36)
∂ q̇1 ∂ q̇2 ∂ q̇m
where ω ∗ describes the explicit time dependence of the angular velocity. In matrix form,
the angular velocity can be written as

ω q ] {q̇} + ω ∗
ω = [ω (8.37)
5 See Section 8.5 for details.
Analytical Mechanics 385

where the kinematic parameters are


 
q ∂ωω ∂ω ω ∂ωω
ω ] =
[ω ... (8.38)
∂ q̇1 ∂ q̇2 ∂ q̇m
ω
∂ω
The kinematic parameters in terms of the generalized speeds are ω k = ∂u k
, where
ω k (k = 1, 2, . . . , n) is the k-th partial angular velocity and ω t is the time partial angular
velocity. The partial angular velocities can be expressed in matrix form by
 
∂ωω ∂ωω ω
∂ω
ωu] =
[ω ... (8.39)
∂u1 ∂u2 ∂un
and the two kinematic parameter matrices are related by

ω q ] = [ω
[ω ω u ] [Y ] (8.40)

Introducing Equation (8.21) into Equation (8.37) and making use of the above equation
gives

ω q ] ([W ] {u} + {X}) + ω ∗


ω = [ω

= [ω ω q ] {X} + ω ∗ = [ω
ω u ] [Y ] [W ]{u} + [ω ω u ] {u} + ω t (8.41)

where the expression for the time partial angular velocities is

ω q ] {X} + ω ∗ = [ω
ω t = [ω ω u ] [Y ] {X} + ω ∗ (8.42)

It follows that we can express the angular velocity in terms of the partial angular veloc-
ities as
ω
∂ω ω
∂ω ω
∂ω
ω = u1 + u2 + . . . + un + ω t = ω 1 u1 + ω 2 u2 + . . . + ω n un + ω t (8.43)
∂u1 ∂u2 ∂un
We have thus expressed a translational velocity as well as an angular velocity in terms
of the kinematic parameters associated with generalized speeds. Use will be made of these
relationships in kinematic analysis, such as in the example that follows, as well as when
deriving Kane’s equations.

Example 8.2
Consider the vehicle in Figure 8.7 and select a set of generalized speeds to describe it. Derive
the kinematic differential equations and the relationships between the generalized speeds
and generalized coordinates XG and YG of the center of mass, which are measured from a
fixed reference point.
This vehicle is slightly different from the one in Figure 8.5 in that it is guided by two
forces in the back and point B has no constraints (or friction forces) acting on it.6 In essence,
the vehicle is like a chariot or a supermarket cart. The velocity of the center of mass is

vG = ẊG I + ẎG J [a]

The unit vectors I, J and i, j are related by

I = cos θi − sin θj J = sin θi + cos θj [b]


6 As if there is a wheel with a caster at point B.
386 Applied Dynamics

y L

C x

FC B X Y
G t y
A 
d
x
 
D X X
FD FA

FIGURE 8.7
Vehicle resembling a chariot or supermarket cart.

so we can write the velocity of the center of mass in terms of the moving coordinates as
   
vG = ẊG cos θ + ẎG sin θ i + −ẊG sin θ + ẎG cos θ j [c]

and the velocity of point A becomes

vA = vG + ω × rA/G = vG + θ̇k × −di

   
= ẊG cos θ + ẎG sin θ i + −ẊG sin θ + ẎG cos θ − dθ̇ j [d]
The nonholonomic constraint is

vA · j = −ẊG sin θ + ẎG cos θ − dθ̇ = 0 [e]

A suitable choice for the generalized speeds is u1 = vA and u2 = θ̇. This selection is
intuitive, as it directly follows from the nonholonomic constraint. The velocity of point A
becomes vA = vA i, and we can express the velocity of the center of mass as

vG = vA + ω × rG/A = vA i + θ̇k × di = vA i + dθ̇j = u1 i + du2 j [f ]

Equating the x and y components of vG from Equations [c] and [f] gives

ẊG ẎG
u1 = ẊG cos θ + ẎG sin θ u2 = − sin θ + cos θ [g]
d d
which can be solved for ẊG and ẎG , with the result

ẊG = u1 cos θ − du2 sin θ ẎG = u1 sin θ + du2 cos θ [h]

The expression for θ̇ is straightforward:

θ̇ = u2 [i]

The three equations in Equations [h] and [i] constitute the kinematic differential equations
for the vehicle.
Analytical Mechanics 387

The next step is to confirm that the choice of generalized speeds is indeed appropriate.
To this end, first calculate the coefficient matrices [Y ] and [W ]. From Equations [g] we can
write  
 Ẋ
sin θ 0  G 
  
u1 cos θ
= ẎG = [Y ] {q̇} [j]
u2 − d1 sin θ d1 cos θ 0
θ̇
and from Equations [h]–[i]
   
ẊG cos θ −d sin θ  
u1
 ẎG  =  sin θ d cos θ  = [W ] {u} [k]
u2
θ̇ 0 1

Note that because u2 = θ̇, we can express the matrix [Y ] also as


 
cos θ sin θ 0
[Y ] = [l]
0 0 1

The conclusion is that the matrix [Y ] is not unique because two generalized speeds are
being expressed in terms of three generalized velocities. On the other hand, the matrix [W ]
is unique.
It is necessary to check that the relationship Equation (8.24) holds. The vectors {X}
and {Z} are identically zero; {X} = {0} and {Z} = {0}. Multiplying [Y ] and [W ] gives
(for both forms of [Y ])
 
  cos θ −d sin θ  
cos θ sin θ 0  1 0
Form 1: [Y ] [W ] = sin θ d cos θ =
− d1 sin θ d1 cos θ 0

0 1
0 1
 
  cos θ −d sin θ  
cos θ sin θ 0  sin θ 1 0
Form 2: [Y ] [W ] = d cos θ  = [m]
0 0 1 0 1
0 1
Next, use Equation (8.25) to check that the constraint relationships are satisfied. The
constraint equation, Equation [e], can be written in the form [a]{q̇} = {0}, where [a] =
T
[− sin θ cos θ − d ], {q} = [XG YG θ ] . The first expression in Equation (8.25) becomes
 
cos θ −d sin θ
[a] [W ] = [ − sin θ cos θ − d]  sin θ d cos θ  = [ 0 0 ] [n]
0 1

The second expression in Equation (8.25) is identically zero as {X} = {0} and {b} = {0}.
Therefore, the choice of generalized speeds is a proper one.
Let us next consider the partial velocities. Writing the angular velocity vector as ω =
u2 k, the partial velocities for vG and ω = u2 k become

1 ∂vG 2 ∂vG ω
∂ω
vG = = i vG = = dj ω1 = = 0
∂u1 ∂u2 ∂u1

ω
∂ω
ω2 = = k t
vG = 0 ωt = 0 [o]
∂u2
388 Applied Dynamics

8.4 Virtual Displacements and Virtual Work


The study of analytical mechanics requires knowledge of concepts from variational calculus.
This section introduces the basic ideas of the calculus of variations and applies these concepts
to dynamics. The discussion begins with types of variables and the differential of a function
versus its variation.
Consider a function of the form f (q1 , q2 , . . . , qn , t). The variables qk (k = 1, 2, . . . , n) are
independent of each other. Even though the variables qk are themselves functions of time,
time may explicitly enter the formulation. For example, f = q1 sin q2 + t cos q3 + 3t2 . It is
clear that qk and time are different types of variables. We refer to time t as the independent
variable and to qk as the dependent variable. The term dependent is used here to denote
the dependence of the variables on time, rather than on each other.
The differential of the function f is denoted by df and is defined as
∂f ∂f ∂f ∂f
df = dq1 + dq2 + . . . + dqn + dt (8.44)
∂q1 ∂q2 ∂qn ∂t
where the terms dqk are differentials of the variables. When obtaining the differential of a
function, we differentiate all the terms in the expression for f , including time. A common
function is the position vector r (q1 , q2 , . . . , qn , t), as defined in Equation (8.15). It follows
that the differential of r is
∂r ∂r ∂r ∂r
dr = dq1 + dq2 + . . . + dqn + dt (8.45)
∂q1 ∂q2 ∂qn ∂t
Now, consider another type of differential element, that of virtual displacements. De-
noted by δqk , virtual displacements are similar to the differential of a displacement (or of a
variable). Virtual displacements are
• Infinitesimal quantities;
• Consistent with the system constraints, but are arbitrary otherwise (as opposed to a
differential, which is exact); and
• Obtained by holding time (that is, the independent variable) fixed; therefore virtual
displacements occur instantaneously, and time is not involved in their applications.
Dealing with virtual displacements is like imagining the system in a different position
that is physically realizable, while freezing time. It is as if a different set of forces was applied
and as a result the system moved to another location via one of the admissible paths it can
follow.
The rules of differentiation and variation are similar. We can interchange the time differ-
entiation and variation operators. For example, δ q̇k = δ (dqk /dt) = d (δqk ) /dt. The variation
of the function f is denoted by δf , where
∂f ∂f ∂f
δf = δq1 + δq2 + . . . + δqn (8.46)
∂q1 ∂q2 ∂qn
There is no time expression in the variation, even though the function f may be an
explicit function of time. The variation of the position vector in Equation (8.15) is referred
to as virtual displacement and has the form
∂r ∂r ∂r
δr = δq1 + δq2 + . . . + δqn (8.47)
∂q1 ∂q2 ∂qn
Analytical Mechanics 389

The variation of a position vector can be obtained in two different ways. One way
is by differentiating the analytical expression for the position vector with respect to the
generalized coordinates. This is known as the analytical approach. While straightforward,
this approach may lead to complex expressions, especially when we express r in terms of a
moving reference frame. In such a case, we must also take the variations of the unit vectors
of the moving frame (except when the motion of the reference frame is considered to be a
known quantity and is not treated as a variable).
The second way to obtain the variation of a position vector is known as the kinematical
approach, in which we exploit the similarity between Equations (8.16) and (8.47). Elimina-
tion of the partial derivative with respect to time in Equation (8.16) and replacement of the
time derivative with the variation gives Equation (8.47). This implies that if the velocity
is known, the associated virtual displacements can be obtained directly from the velocity.
Comparing the differential in Equation (8.45) and variation in Equation (8.47) leads to the
same result.
Next, consider the work done by a force over a virtual displacement. Recall that the
expression for infinitesimal work has the form dW = F · dr, where r is the position of
the point to which the force F is applied. The work done by a force F over the virtual
displacement δr is defined as the virtual work or variation of work and is denoted by δW .
The virtual work has the form

δW = F · δr (8.48)

The general case of virtual work associated with a force will be discussed in the next
section. For now, consider a holonomic constraint, say, f (x, y, z, t) = 0, and the associated
constraint force F0 . It was shown earlier that the constraint force is perpendicular to the
surface defined by the constraint equation and that the incremental work done by a holo-
nomic constraint force that is not an explicit function of time is zero. We can extend this
idea to virtual work, and denote the virtual work of a constraint force by δW 0 = F0 · δr.
Because virtual displacements do not involve time (as they are obtained by holding time
fixed), we can state that

δW 0 = F0 · δr = 0 (8.49)

so that the work performed by a holonomic constraint force on any virtual displacement
is zero. This result will be used to derive the principle of virtual work as well as other
variational principles, such as D’Alembert’s principle and Hamilton’s principle.

Example 8.3
Consider the rotating pendulum in Figure 8.8, whose length changes as the spring is
stretched or compressed. The angular velocity of the column is Ω = Ω0 sin 3t. Find the
virtual displacement of point P .
The xyz coordinate system rotates with the pendulum. The position of the pendulum
is
rP = di + L (sin θi − cos θk) [a]
Let us find the virtual displacement of point P using the velocity approach. This system
has two degrees of freedom and the variables θ and L describe the position of point P . Note
that, even though it is not constant, the angular velocity of the column is a known quantity,
and hence, its variation is zero.
The velocity of point P is written as

vP = vB + vP/B [b]
390 Applied Dynamics

Z
z y
d
O x
B
L

FIGURE 8.8
Rotating pendulum.

where

vB = Ωk × di = dΩj vP/B = Ωk × L (sin θi − cos θk) + vP/Brel [c]

in which
vP/Brel = L̇ (sin θi − cos θk) + Lθ̇ (cos θi + sin θk) [d]
Collecting terms, the velocity of point P becomes
   
vP = L̇ sin θ + Lθ̇ cos θ i + (d + L sin θ) Ωj + −L̇ cos θ + Lθ̇ sin θ k [e]

To obtain the virtual displacement, replace L̇ with δL and θ̇ with δθ in the above
equation. Further, remove all the terms involving Ω, as Ω(t) is a known quantity and it is
not the derivative of a variable. The virtual displacement of P , δrP , then becomes

δrP = (sin θδL + L cos θδθ) i + (− cos θδL + L sin θδθ) k [f ]

which can also be written as

δrP = (sin θi − cos θk) δL + (L cos θi + L sin θk) δθ [g]

8.5 Virtual Displacements and Virtual Work for Rigid Bodies


The virtual displacement and virtual work expressions for rigid bodies have the same form
as they do for particles, as long as we consider individually each force that contributes to
the virtual work and we calculate the displacement of each point acted upon by external
forces. Sometimes, it is more advantageous to deal with the resultants of the applied forces.
This section extends the principle of virtual work to rigid bodies and considers resultant
forces and moments.
Consider a system (of particles, rigid bodies, or both) that has n degrees of freedom and
to which N forces Fi (i = 1, 2, . . . , N ) are applied. Denote the locations of the points at
Analytical Mechanics 391

M1

dm
1 F
2 MG
F1
G G
F2
r1 rG rG
M2
r2

FIGURE 8.9
Forces and moments applied to a rigid body and resultants.

which these forces are applied by ri , as shown in Figure 8.9. The virtual work of each force
is denoted by δWi and defined as

δWi = Fi · δri (8.50)

The total virtual work is the sum of the contributions of the individual forces
N
X N
X
δW = δWi = Fi · δri (8.51)
i=1 i=1

We saw in Chapter 4 that the resultant


P of all forces and moments acting on a body can
be viewed as a resultant force F = Fi acting through the center of mass and a resultant
moment about the center of mass,
0
N
X N
X
MG = ρ i × Fi + Mj (8.52)
i=1 j=1

in which ρ i is the vector from the center of mass G to the point at which force Fi is applied.
Also, Mj are the applied torques (concentrated moments).
Return to the rigid body in Figure 8.9 undergoing plane motion, and apply to it external
forces Fi that go through points described by ri , as well as a number of torques Mj .
When dealing with a rigid body, the point of application of the concentrated moments is
immaterial.7 The position vector of the points through which the forces are applied are

ri = rG + ρ i (8.53)

and the associated virtual work is


0
N
X N
X
δW = Fi · δri + Mj δθ (8.54)
i=1 j=1

For plane motion, the velocity of point i is

vi = vG + θ̇k × ρ i (8.55)
7 This is not the case with deformable bodies, where the location of the applied moment changes the

internal forces and the stress distribution.


392 Applied Dynamics

so that the virtual displacement of point i becomes

δri = δrG + δθk × ρi (8.56)

Introducing this expression into Equation (8.54) gives


0
N
X N
X
δW = Fi · (δrG + δθk × ρ i ) + Mj δθ
i=1 j=1

0
N
X N
X N
X
= Fi · δrG + Fi · δθk × ρ i + Mj δθ (8.57)
i=1 i=1 j=1

The last term in the above equation can be written in vector form by noting that
Mj = Mj k and the variation of the rotation vector is δθk. We can use this notation only
when dealing
P with planeP motion. As a result, the contribution of the concentrated moments
becomes Mj · δθk = Mj δθ. P
Let us evaluate the individual terms in Equation (8.57). Noting that F = Fi , the first
term becomes F · δrG . To evaluate the second term, the relationship

a · (b × c) = b · (c × a) (8.58)

is applied to the second term on the right side of Equation (8.57), with the result

Fi · (δθk × ρ i ) = δθk · (ρρi × Fi ) (8.59)

so that
N
X N
X
Fi · (δθk × ρ i ) = δθk · (ρρi × Fi ) (8.60)
i=1 i=1

Now, introduce the developments above to the expression for the virtual work in Equa-
tion (8.57), which gives
0
N
X N
X
δW = F · δrG + δθk · (ρρi × Fi ) + Mj · δθk (8.61)
i=1 j=1

Considering the definition of resultant moment from Equation (8.52), the second and third
terms in the above equation are recognized as
0
N
X N
X
δθk · (ρρi × Fi ) + Mj · δθk
i=1 j=1

 
XN N0
X
=  ρ i × Fi + Mj  · δθk = MG · δθk = MG δθ (8.62)
i=1 j=1

so that the virtual work expression for a rigid body in plane motion can be expressed in
terms of the resultant force and resultant moment as

δW = F · δrG + MG δθ (8.63)
Analytical Mechanics 393

For three-dimensional motion, as discussed in Chapter 2, the angular velocity vector is


not the derivative of a rotation vector but is a defined quantity. Angular velocity components
are nonholonomic variables, similar to the generalized speeds or quasi-velocities. It follows
that we cannot obtain the variation of angular velocity if angular velocity is expressed
in terms of the angular velocity components ωi (i = 1, 2, 3). However, expressing angular
velocity in terms of the angles of a rotation sequence makes it possible to use the variational
rotation.

a) Z b) Z'
Z' z
z 

 .

Y', y
X' 
 .
Y x 
. 

Y', y
X, X'
x

FIGURE 8.10
a) Two rotations: θ about X and ψ about Y 0 . b) The second rotation (by ψ).

As an illustration, consider the coordinate transformation in Figure 8.10a. Begin with


an inertial frame XY Z. First, rotate it about the X axis by θ to yield an X 0 Y 0 Z 0 frame and
then rotate the X 0 Y 0 Z 0 frame about Y 0 by ψ to get the xyz coordinates. The second rotation
is highlighted in Figure 8.10b. The angular velocity of the xyz frame can be expressed as

ω = θ̇I + ψ̇J0 = θ̇I0 + ψ̇j (8.64)

The unit vector J0 in the X 0 Y 0 Z 0 frame can be expressed in terms of XY Z as

J0 = cos θJ + sin θK (8.65)

and, from Figure 8.10b, we can write I0 as I0 = cos ψi + sin ψk, so the angular velocity
vector in terms of the XY Z and xyz coordinates becomes

ω = θ̇I + ψ̇ cos θJ + ψ̇ sin θK = θ̇ cos ψi + ψ̇j + θ̇ sin ψk (8.66)

It then becomes possible to define a “virtual rotation” vector as8

δ θ = δθI + cos θ δψJ + sin θ δψK = cos ψ δθi + δψj + sin ψ δθk (8.67)

keeping in mind that the above expression cannot be obtained by differentiating or taking
the variation of a vector. For the general case, the virtual rotation becomes
∂ωω ∂ωω ∂ω ω
δθ = δq1 + δq2 + . . . + δqn (8.68)
∂ q̇1 ∂ q̇2 ∂ q̇n
8 Note the difference in notation for δ θ (instead of δθ
θ ), indicating that the quantity δ θ is not a true
variation but a defined one.
394 Applied Dynamics

Hence, the expression for the virtual work for a rigid body in three-dimensional motion has
the form

δW = F · δrG + MG · δ θ (8.69)

We can write the virtual work of a body in terms of the individual forces by using
Equation (8.51) or in terms of the resultant force and moment, by using Equation (8.63)
for two-dimensional motion and Equation (8.69) for motion in three dimensions.

8.6 Generalized Forces


This section analyzes the virtual work expression for a variety of cases and derives the prin-
ciple of virtual work, which is a powerful method of deriving static equilibrium equations.
Consider a system that has n degrees of freedom and to which N forces Fi (i = 1, 2, . . . , N )
are applied. Denote the locations of the points at which these forces are applied by ri , as
shown in Figure 8.9. The virtual work of each force is δWi = Fi · δri , and the total virtual
work is given by Equation (8.51).
Next, express each virtual displacement in terms of the generalized coordinates as
∂ri ∂ri ∂ri
δri = δq1 + δq2 + . . . + δqn i = 1, 2, . . . , N (8.70)
∂q1 ∂q2 ∂qn
and introduce the above expression into the virtual work, with the result
N N X
n
X X ∂ri
δW = Fi · δri = Fi · δqk (8.71)
i=1 i=1 k=1
∂qk

Rearranging the summations, we can write


n
N X n N
!
X ∂ri X X ∂ri
δW = Fi · δqk = Fi · δqk (8.72)
i=1 k=1
∂qk i=1
∂qk
k=1

The term inside the parentheses in the above equation is referred to as the generalized force
Qk associated with the k-th generalized coordinate, so
N N
X ∂ri X ∂vi
Qk = Fi · = Fi · (8.73)
i=1
∂qk i=1
∂ q̇k

The virtual work can be expressed as


n
X
δW = Qk δqk (8.74)
k=1

Considering the cause-and-effect principle in Chapter 1, the relationship between a gen-


eralized coordinate and a generalized force is similar to the relationship between a displace-
ment and the force that causes that displacement.
Next, we classify forces that contribute to the virtual work and the associated generalized
forces. One classification methodology is to split forces into the three categories as follows:
Analytical Mechanics 395

• Conservative Forces. For a conservative force, the incremental work, dWc , can be
written as a perfect differential of the potential energy V ,

dWc = Fc · dr = −dV (8.75)

where the subscript c denotes that the force is conservative. It follows that the virtual
work of a conservative force can be written as

δWc = Fc · δr = −δV (8.76)

The variation of the potential energy is


∂V ∂V ∂V
δV = δq1 + δq2 + . . . + δqn (8.77)
∂q1 ∂q2 ∂qn

and, considering Equation (8.74), generalized forces associated with conservative forces
can be expressed in terms of the potential energy as
∂V
Qkc = − k = 1, 2, . . . , n (8.78)
∂qk

• Nonconservative Forces. The incremental work of nonconservative forces cannot be


written as a perfect differential. Hence, the only allowable expression for the incremental
work is

dWnc = Fnc · dr (8.79)

where the subscript nc denotes nonconservative. It follows that the associated general-
ized force, Qknc , has the form

∂r ∂v
Qknc = Fnc · = Fnc · (8.80)
∂qk ∂ q̇k

• Constraint and Reaction Forces. It was shown earlier that the virtual work asso-
ciated with a holonomic constraint force is zero. This result extends to interconnected
bodies and to forces that keep bodies interconnected, such as forces at joint B in the
slider in Figure 8.11. The forces at joint B are holonomic constraint forces, keeping the
two links from moving against each other, as will be discussed in the next example.

g B
1 2
m, L m, L

P k

O

FIGURE 8.11
Slider-crank mechanism.
396 Applied Dynamics

Consider a system consisting of N particles (or bodies) and a number of forces that act
on each body. Denoting the resultant of all forces acting on the i-th body by Ri we can
express it as

Ri = Fi + F0i = Fic + Finc + F0i (8.81)

where Fi is the resultant of all forces acting on the i-th body that are external to the
system and F0i are the forces exerted on the i-th body by the other bodies in the system.
Such forces are internal to the system. The external forces are separated as Fic , which is
the resultant of all conservative forces acting on body i and Finc , which is the resultant of
all nonconservative forces acting on body i.
Denoting the displacement of each body by ri , the virtual work for each body becomes

δWi = Ri · δri = (Fic + Finc + F0i ) · δri (8.82)

Summing over all forces gives


N
X N
X N
X N
X
δW = Ri · δri = Fic · δri + Finc · δri + F0i · δri (8.83)
i=1 i=1 i=1 i=1

The first term on the right describes contributions of the conservative forces and has
the form
N n n
X X ∂V X
Fic · δri = −δV = − δqk = Qkc δqk (8.84)
i=1
∂qk
k=1 k=1

The second term is the contribution of the nonconservative forces


N n N
! n
X X X ∂vi X
Finc · δri = Finc · δqk = Qknc δqk (8.85)
i=1 i=1
∂ q̇k
k=1 k=1

so that the generalized forces associated with the nonconservative forces are
N
X ∂vi
Qknc = Finc · (8.86)
i=1
∂ q̇k

The third term on the right side of Equation (8.83) vanishes, because when we add the
contribution of all constraint forces on the system, their effects cancel each other and
N
X
F0i · δri = 0 (8.87)
i=1

It follows that the virtual work can be calculated as


N
X N
X N
X
δW = Fi · δri = Fic · δri + Finc · δri (8.88)
i=1 i=1 i=1

and the generalized forces become


N
∂V X ∂vi
Qk = Qkc + Qknc = − + Fi · (8.89)
∂qk i=1 nc ∂ q̇k
Analytical Mechanics 397

When using resultant forces and moments, Equation (8.63) for two-dimensional motion
and Equation (8.69) for three-dimensional motion, the associated kinematic variables are
∂vG ∂ω ω
∂ q̇k and ∂ q̇k and the generalized forces become

∂vG ∂ω ω
Qk = F · + MG · (8.90)
∂ q̇k ∂ q̇k
Table 8.1 summarizes the different procedures for calculating the generalized forces.

TABLE 8.1
How to calculate generalized forces Qk (k = 1, 2, . . . , n)

Equation Procedure
(8.51) Consider each force individually, write associated virtual work,
collect coefficients of δqk (k = 1, 2, . . . , n)
∂ri
(8.73) Use individual forces and kinematic variables ∂q k
or ∂∂vq̇ki
(8.89) Use potential energy for contribution from conservative forces
(8.90) Use resultant forces and moments and associated kinematic variables

Example 8.4
For the slider-crank mechanism in Figure 8.11, calculate the generalized forces and show
that the reaction forces between the links at points O and B do not contribute to the virtual
work. The spring is not stretched when both bars are vertical.
The free-body diagrams of the individual components are shown in Figure 8.12. The
system has one degree of freedom and a suitable choice for the generalized coordinate is the
angle θ. The slider moves horizontally. The forces acting on the mechanism are the force
of gravity in each link, the spring force, the reaction forces at the joints O and B, and the
normal force NP . The reaction forces at O do not contribute to the virtual work because

F By
B y
1 F Bx
B 2
G1 F By G2
x
Oy mg mg
  P
O 2kLcos
Ox
NP

FIGURE 8.12
Free-body diagram.

point O does not move and δrO = 0. Considering the normal force at P , NP , we can express
in vector form the normal force, the position vector, and the virtual displacement as

NP = NP j rP = 2L cos θi δrP = −2L sin θ δθi [a]


398 Applied Dynamics

The virtual work associated with the normal force NP is zero, as the force and virtual
displacement vectors are perpendicular to each other

NP · δrP = −NP j · 2L sin θ δθi = 0 [b]

This result is expected, as the normal force is, in essence, a holonomic constraint force,
preventing the slider from moving in the y direction.
Next, consider the reaction forces at joint B. The force on link 1 is FB1 = FBx i + FBy j
and the position vector is rB = L cos θi + L sin θj, so the virtual displacement becomes
δrB = −L sin θδθi + L cos θδθj. It follows that the virtual work becomes

δW1 = FB1 · δrB = −FBx L sin θ + FBy L cos θ δθ [c]

For link 2, the reaction force at B is the same as the reaction force on the first link,
but in the opposite direction, FB2 = −FB1 = −FBx i − FBy j. The position vector rB is the
same as before, so the contribution to the virtual work of the force at B on the second link
becomes 
δW2 = FB2 · δrB = FBx L sin θ − FBy L cos θ δθ = −δW1 [d]
Adding the two virtual work components we obtain the expected result of

δW1 + δW2 = 0 [e]

The reaction forces at joints O and B, as well as the normal force at P , do not contribute
to the virtual work. Gravity and the spring are the only forces contributing to the virtual
work. Both these forces are conservative, so we can use potential energy to calculate the
virtual work and the generalized force. The potential energy can be written (taking point
O as the datum and recalling that the unstretched position of the spring is when both rods
are vertical) as
L 1 2
V = 2mg sin θ + k (2L cos θ) [f ]
2 2
The generalized force is (note that we can use a regular derivative and not partial derivative,
as there is one degree of freedom)
dV
Qθ = − = −mgL cos θ + 4kL2 cos θ sin θ [g]

Example 8.5
Consider the T-shaped bar in Figure 8.13a and find the virtual displacement of the center
of mass G and the generalized forces. Use as generalized coordinates the displacement of
point A along the incline, and the angle line AG makes with the horizontal.
We can show that the center of mass lies a distance of 3L 4 from point A. Consider the
xy and x0 y 0 coordinates for describing the motion. Both coordinate systems are fixed. The
positions of point A and of the center of mass G can be expressed as

rA = si0 = s cos 30◦ i + s sin 30◦ j


3 3
rG = rA + rG/A = si0 − L cos θj + L sin θi
4 4
3 3
= s cos 30◦ i + s sin 30◦ j − L cos θj + L sin θi [a]
4 4
Taking the variation of these displacement vectors results in
3
δrA = δsi0 δrG = δs (cos 30◦ i + sin 30◦ j) + Lδθ (cos θi + sin θj) [b]
4
Analytical Mechanics 399

y'
F x' F x'

A 15o A
s 15o
y N
L 30o L
30o  
G G
x
L mg L

FIGURE 8.13
a) T-shaped bar, b) free-body diagram.

The force of gravity and the external force can be expressed as

Fg = −mgj F = F (cos 15◦ i0 + sin 15◦ j0 ) [c]

The virtual work expression is

δW = Fg · δrG + N · δrA + F · δrA [d]

The second term, N · δrA , disappears because the constraint force N = N j0 and virtual
displacement of A, δrA = δsi0 , are perpendicular to each other. Carrying out the dot product
for the remaining terms results in
 
3
δW = Fg · δrG + F · δrA = −mg sin 30 δs + L sin θ δθ + F cos 15◦ δs

[e]
4

The virtual work, in terms of the generalized forces, is

δW = Qs δs + Qθ δθ [f ]

Comparing Equations [e] and [f] the generalized forces are


1 3
Qs = − mg + F cos 15◦ Qθ = − mgL sin θ [g]
2 4

8.7 Principle of Virtual Work for Static Equilibrium


The stage is set to state the principle of virtual work for static equilibrium. Consider a
system, composed of N components modeled as particles, that is at rest. This implies that
every component of the system is at rest. The resultant force on each component is zero, or

Ri = 0 i = 1, 2, . . . , N (8.91)
400 Applied Dynamics

Taking the dot product of the above expression with δri and summing gives
N
X
Ri · δri = 0 = δW (8.92)
i=1

so that the total virtual work of a system that is at rest is zero. But, from Equation (8.88),
only the external, or impressed forces contribute to the virtual work, so we can write
N
X
δW = Fi · ri = 0 (8.93)
i=1

The above equation, first formulated by Johann Bernoulli (1667–1748), scion of a large
Swiss family that also made contributions to mathematics, statistics, thermodynamics, and
financial systems, is known as the principle of virtual work for static equilibrium. It states
that, at static equilibrium, the work performed by the external, impressed forces through
virtual displacements compatible with the system constraints is zero. In terms of generalized
forces the principle of virtual work has the form
n
X
δW = Qk δqk = 0 (8.94)
k=1

When the system is represented in terms of a set of independent generalized coordinates,


because the generalized coordinates qk are independent of each other, their variations δqk
are also independent. Therefore, for the above equation to hold, each of the coefficients of
δqk , that is, Qk , must vanish individually
N N
X ∂ri X ∂vi
Qk = Fi · = Fi · = 0 k = 1, 2, . . . , n (8.95)
i=1
∂qk i=1
∂ q̇k

In the presence of conservative forces, we can take advantage of the potential energy
and write
∂V
− + Qknc = 0 (8.96)
∂qk
The above results can be interpreted as follows: Because independent generalized coor-
dinates represent the independent motion of each degree of freedom, their corresponding
generalized forces must vanish at equilibrium.
Derivation of the principle of virtual work can easily be extended to rigid bodies. We will
perform this extension in the next section, within the context of D’Alembert’s principle.

Example 8.6
Consider the slider in Example 8.4 and calculate the equilibrium positions. Recall that the
spring is not stretched when the bars are vertical.
Setting the expression for the generalized force to zero in Equation [g] of Example 8.4
gives
Qθ = −mgL cos θ + 4kL2 cos θ sin θ = 0 [a]
Dividing by mg and rearranging gives the equilibrium equation as
 
4kL
cos θ 1 − sin θ = 0 [b]
mg
Analytical Mechanics 401

The solutions are


π 3π mg  mg 
cos θ = 0 =⇒ θe = , sin θ = =⇒ θe = sin−1 [c]
2 2 4kL 4kL
The equilibrium position θe = π2 corresponds to when the rods are upright. The position
θe = 3π2 corresponds to when the rods are vertical and below the spring. The equilibrium
position θe = sin−1 4kL
mg mg
, which has two solutions, is only valid when 4kL < 1 or mg < 4kL,
that is, when the spring is strong enough to counter the weight of the rods. The three
equilibrium positions are shown in Figure 8.14.

B
a) b) c) B

e = /2 e = 3 /2 k
e
O O O P
P P
k k

FIGURE 8.14
π 3π
c) θe = sin−1 mg

Equilibrium positions: a) θe = 2, b) θe = 2 , 4kL (represents two solutions).

8.8 D’Alembert’s Principle


D’Alembert’s principle extends the principle of virtual work from the static to the dy-
namic case. In keeping with the formulation in the previous section, we will first derive
D’Alembert’s principle for a system of particles and then extend it to rigid bodies.

8.8.1 General Formulation


Consider the system of N components modeled as particles in the previous section. If the
system is not at rest, we can write Newton’s Second Law for the i-th particle as
d
Ri = mi ai = pi i = 1, 2, . . . , N (8.97)
dt
where pi = mi vi is the linear momentum of the i-th particle and Ri is the resultant of all
forces acting on the i-th particle. As in the static case, the resultant Ri is split into the
sum of the externally applied forces (that is, external to the entire system) and constraint
forces as

Ri = Fi + F0i (8.98)
402 Applied Dynamics

Introducing the above equation into Equation (8.97) gives

Fi + F0i − ṗi = 0 (8.99)

This equation is known as the dynamic equilibrium relationship, where the negative of
the rate of change of the linear momentum, −ṗi = −mi ai , is treated as a force, referred
to as the inertia force. We can now treat the dynamic system as if it is a static system
subjected to an inertia force and invoke the principle of virtual work. Equation (8.99) is
referred to by some (but not in this text) as the D’Alembert’s principle.
The next step is to take the dot product of Equation (8.99) with the variation of the
displacement, which results in

(Fi + F0i − mi ai ) · δri = 0 (8.100)

Summing over all components gives


N
X
(Fi + F0i − mi ai ) · δri = 0 (8.101)
i=1

Recalling that the work done by the constraint forces over virtual displacements is zero,
or
N
X
F0i · δri = 0 (8.102)
i=1

and subtracting Equation (8.102) from Equation (8.101), the resulting expression is
N
X
(Fi − mi ai ) · δri = 0 (8.103)
i=1

This we call the generalized principle of D’Alembert, or D’Alembert’s principle. The prin-
ciple of virtual work, given in Equation (8.93), becomes a special case of D’Alembert’s
principle.
The D’Alembert’s principle given in Equation (8.103) is a fundamental expression that
provides a compete formulation of all of the problems of mechanics. Hamilton’s principle
and Lagrange’s equations are all derived from D’Alembert’s principle, as will be shown
in subsequent sections. The advantage of using D’Alembert’s principle over a Newtonian
approach is that constraint forces are eliminated from the formulation. This advantage
becomes more pronounced for systems with several degrees of freedom or when the number
of components is larger than the number of degrees of freedom.

8.8.2 Extension to Rigid Bodies


We next extend D’Alembert’s principle to rigid bodies. Section 8.5 obtained the expression
for virtual work in terms of rigid bodies, and D’Alembert’s principle in Equation (8.103)
deals with accelerations. Consider a rigid body as a collection of an infinite number of
particles (or differential elements), where the position, velocity, and acceleration of each
point (or differential element) can be expressed in terms of the center of mass as (see Figure
8.15)

ri = rG + ρ i vi = vG + ω × ρ i ai = aG + α × ρ i + ω × (ω
ω × ρi) (8.104)
Analytical Mechanics 403

i G
mi

ri rG

FIGURE 8.15
Particle mi (or differential element dm) and geometry.

Noting the developments in Section 8.5 regarding the variation of a rotation, we can
write the virtual displacement of point i as

δri = δrG + δ θ × ρ i (8.105)

The next step is to extend the particle formulation to the differential element formula-
tion. Replacing the summation by integration in Equation (8.103) and mi and ρ i with dm
and ρ , respectively, leads to the acceleration term
N
X Z
mi ai · δri =⇒ a · δr dm (8.106)
i=1

and, substituting in the values of the acceleration in Equation (8.104), gives


Z Z
a · δr dm = (aG + α × ρ + ω × (ω ω × ρ )) · (δrG + δ θ × ρ ) dm (8.107)
R
Noting that by definition of the center of mass ρ dm = 0, expanding the terms in the
above equation and performing the integration leads to six terms. The first five are
Z Z Z
aG · δrG dm = maG · δrG aG · (δδθ × ρ ) dm = 0 α × ρ ) · δrG dm = 0

Z Z
ω × (ω
(ω ω × ρ )) · δrG dm = 0 ω × (ω
(ω ω × ρ )) · (δδθ × ρ ) dm = 0 (8.108)
R
The term (ω ω × (ωω × ρ )) ·(δδθ × ρ ) dm = 0 because ω and δ θ are parallel, so ω × ρ and
δ θ × ρ are parallel as well. It follows that ω × (ω ω × ρ ) and δ θ × ρ are perpendicular to each
other, so that their dot product vanishes. R
The sixth term in Equation (8.107) is (α α × ρ ) ·(δδθ × ρ ) dm, and it it is recognized as
the angular momentum about the center of mass. It will be shown in Chapter 11 that, for
three-dimensional motion, the angular momentum of a body about its center of mass is
defined as
Z
HG = [(ρρ · ρ ) ω − (ρρ · ω ) ρ ] dm (8.109)
404 Applied Dynamics

Using the vector identity


(a × b) · (c × d) = (a · c) (b · d) − (a · d) (b · c) (8.110)
and a = α , b = ρ , c = δ θ , d = ρ , the sixth term becomes
Z Z
α × ρ ) · (δδθ × ρ ) dm =
(α α · δ θ ) (ρρ · ρ ) − (α
[(α α · ρ ) (ρρ · δ θ )] dm

Z
= [(ρρ · ρ ) α − (ρρ · α ) ρ ] dm · δ θ = ḢG · δ θ (8.111)

Combining the above results, we obtain


N
X
mi ai · δri =⇒ maG · δrG + ḢG · δ θ (8.112)
i=1

so D’Alembert’s principle for a rigid body becomes


 
(F − maG ) · δrG + MG − ḢG · δ θ = 0 (8.113)

We can obtain the principle of virtual work for a rigid body by removing the acceleration
term and rate of change of angular momentum term from the above equation.
For the special case of plane motion, because the angular acceleration α and the rotation
θ are in the same direction, the vectors α ×ρρ and δθθ ×ρρ are parallel. Furthermore, the angular
acceleration vector α and position vector ρ are perpendicular to each other, so
Z Z
α × ρ ) · (δθθ × ρ ) dm =⇒ δθα ρ2 dm = IG α δθ = ḢG δθ
(α (8.114)

and D’Alembert’s principle for rigid bodies in plane motion becomes


 
(F − maG ) · δrG + MG − ḢG δθ = 0 (8.115)

8.8.3 Using D’Alembert’s Principle to Obtain Equations of Motion


D’Alembert’s principle in Equation (8.103) can be used in two ways: it can be used to derive
additional variational principles, such as Hamilton’s principle and Lagrange’s equations, or it
can be used directly to obtain the equations of motion. While in the past the D’Alembert’s
principle was primarily viewed as a tool to derive Hamilton’s principle and Lagrange’s
equations, the need to deal with complex multibody systems and advances in computers
and software have increased the attractiveness of the D’Alembert’s principle as a primary
method for obtaining equations of motion.
Consider a system of N particles that has n degrees of freedom. The virtual displace-
ments can be written as
n n
X ∂ri X ∂vi
δri = δqk = δqk i = 1, 2, . . . , N (8.116)
∂qk ∂ q̇k
k=1 k=1

∂ri
with the choice of the kinematic variables to use, ∂q k
or ∂∂vq̇ki , depending on the analyst.
Introducing this expression to D’Alembert’s principle gives
N N n
X X X ∂vi
(Fi − mi ai ) · δri = (Fi − mi ai ) · δqk
i=1 i=1
∂ q̇k
k=1
Analytical Mechanics 405
n N
" #
X X ∂vi
= (Fi − mi ai ) · δqk = 0 (8.117)
i=1
∂ q̇k
k=1

For a set of independent generalized coordinates q1 , q2 , . . . , qn , their variations δqk are


also independent of each other, so the coefficients of δqk in the above equation must vanish
independently, leading to the n equations of motion in the form
N
X ∂vi
(Fi − mi ai ) · = 0 k = 1, 2, . . . , n (8.118)
i=1
∂ q̇k

We can extend this formulation to rigid bodies. In view of the issues associated with
virtual rotations, it is preferable to expand δ θ in terms of the angular velocities and to write
the virtual rotations as
n
X ∂ω ω
δθ = δqk (8.119)
∂ q̇k
k=1

which leads to the equations of motion for a rigid body as


∂vG   ∂ω ω
(F − maG ) · + MG − ḢG · = 0 k = 1, 2, . . . , n (8.120)
∂ q̇k ∂ q̇k
The above equation can be expressed in terms of generalized forces as
∂vG ∂ω ω
maG · + ḢG · = Qk k = 1, 2, . . . , n (8.121)
∂ q̇k ∂ q̇k
where the generalized forces Qk can be calculated by one of the approaches summarized in
Table 8.1.
It follows that for a system of N rigid bodies, the equations of motion can be written as
N  
X ∂vGi   ∂ωωi
(Fi − mi aGi ) · + MGi − ḢGi · = 0 k = 1, 2, . . . , n (8.122)
i=1
∂ q̇k ∂ q̇k

or
N  
X ∂vGi ωi
∂ω
mi aGi · + ḢGi · = Qk k = 1, 2, . . . , n (8.123)
i=1
∂ q̇k ∂ q̇k

For plane motion the above equation reduces to


N  
X ∂rGi ∂θi
mi aGi · + IG θ̈i = Qk (8.124)
i=1
∂qk ∂qk

Equations (8.118)–(8.124), also called D’Alembert’s equations, represent direct use of


D’Alembert’s principle to obtain equations of motion. An extension of D’Alembert’s prin-
ciple to in terms of generalized speeds will be discussed in Section 8.12.

Example 8.7
Consider the T-shaped bar in Example 8.5 and find the equations of motion using
D’Alembert’s principle.
406 Applied Dynamics

The virtual work expression and the generalized forces were calculated in Example 8.5.
The D’Alembert’s principle can be written in terms of the virtual work as
maG · δrG + IG θ̈ δθ = δW [a]
From Example 8.5, the variation of rG has the form
3
δrG = δs (cos 30◦ i + sin 30◦ j) + Lδθ (cos θi + sin θj) [b]
4
Next, calculate the acceleration of the center of mass. Recall that the center of mass G
is at a distance 3L/4 from the pin joint. We can obtain the velocity expression for G by
simply replacing δs with ṡ and δθ with θ̇ in the above equation, which yields
3
vG = ṡ (cos 30◦ i + sin 30◦ j) + Lθ̇ (cos θi + sin θj) [c]
4
Differentiating the velocity one more time, the acceleration of the center of mass is obtained
as
3 3
aG = s̈ (cos 30◦ i + sin 30◦ j) + Lθ̈ (cos θi + sin θj) + Lθ̇2 (− sin θi + cos θj) [d]
4 4
At this point, we can proceed in two different ways to find the equations of motion. One
is to take the dot product of aG and δrG and then separate the terms that have δs and δθ.
The second approach is to calculate the kinematic variables and use Equation (8.124). Let
us use the second approach to illustrate the use of kinematic variables. The equations of
motion will be derived using
∂vG ∂θ ∂vG ∂θ
maG · + IG θ̈ = Qs maG · + IG θ̈ = Qθ [e]
∂ ṡ ∂s ∂ θ̇ ∂θ
The kinematic variables are
∂vG ∂vG 3 ∂θ ∂θ
= cos 30◦ i + sin 30◦ j = L (cos θi + sin θj) = 0 = 1 [f ]
∂ ṡ ∂ θ̇ 4 ∂s ∂θ
We calculate the mass moment of inertia of the T-shaped bar by taking the mass moment
of inertia for each segment of the bar, IG = I1 + I2 , and invoking the parallel axis theorem
for each part:
I1 = I1G + m1 d21 I2 = I2G + m2 d22 [g]
where m1 = m2 = m L
2 and d1 = d2 = 4 . It follows that the total mass moment of inertia
about the center of mass is
 2  2
1 m 2 m L 1 m 2 m L 7
IG = L + + L + = mL2 [h]
12 2 2 4 12 2 2 4 48
Taking the dot products in Equation [e] results in
∂vG 3 √  3  √ 
maG · = ms̈ + mLθ̈ 3 cos θ + sin θ + mLθ̇2 − 3 sin θ + cos θ
∂ ṡ 8 8
∂vG 9 3 √ 
maG · = mL2 θ̈ + mLs̈ 3 cos θ + sin θ
∂ θ̇ 16 8
∂θ ∂θ 7
IG θ̈ = 0 IG θ̈ = mL2 θ̈ [i]
∂s ∂θ 48
Combining the terms in Equation [i] gives the equations of motion as
3 √  3  √  1
For s: ms̈ + mLθ̈ 3 cos θ + sin θ + mLθ̇2 − 3 sin θ + cos θ = − mg + F cos 15◦
8 8 2
17 3  √  3
For θ: mL2 θ̈ + mLs̈ 3 cos θ + sin θ + mgL sin θ = 0 [j]
24 8 4
Analytical Mechanics 407

Example 8.8
A bead of mass m is free to slide on a ring (hoop) of radius R, as shown in Figure 8.16.
The ring rotates with constant angular velocity Ω. Find the equation of motion using
D’Alembert’s principle.

  



 


 


FIGURE 8.16
Bead on a rotating hoop. a) Geometry, b) free-body diagram.

This problem involves a single particle, so we can drop the subscript in Equation (8.103)
and write
(F − ma) · δr = 0 [a]
The free-body diagram is given in Figure 8.16b. The xyz axes are attached to the hoop and
θ is the generalized coordinate. The position vector is

r = R sin θj − R cos θk [b]

Because the motion of the hoop, or the xyz frame, is known, there is no variation of the
associated unit vectors i, j, and k. The variation of the position vector thus becomes

δr = R cos θ δθj + R sin θ δθk [c]

The velocity of the bead is


v = ω × r + vrel [d]
where

ω × r = Ωk × (R sin θj − R cos θk) = −RΩ sin θi vrel = R cos θθ̇j + R sin θθ̇k [e]

The term ω × r = −RΩ sin θi does not contribute to the virtual displacement, as Ω is
not the derivative of a variable. Rather, it is a known quantity. We see once again that
Equation [c] is the variation of the position vector, δr.
We can obtain the acceleration by differentiating the velocity directly, or by using a
relative acceleration expression. Selecting the latter, and noting that the angular velocity
of the hoop is constant, the acceleration becomes

a = ω × (ω
ω × r) + ω × vrel + arel [f ]

where
ω × r) = Ωk × (Ωk × (R sin θj − R cos θk)) = −RΩ2 sin θj
ω × (ω [g]
408 Applied Dynamics

ω × vrel = 2Ωk × (R cos θθ̇j + R sin θθ̇k) = −2RΩ cos θθ̇i


2ω [h]

arel = R cos θθ̈j + R sin θθ̈k − R sin θθ̇2 j + R cos θθ̇2 k [i]
The only force acting on the system which is not a constraint force is gravity, and it
has the form F = −mgk. The constraint forces N1 and N2 ensure that the bead stays
on the hoop. Substituting Equations [c] and [f] into the D’Alembert’s principle and using
Equations [g]–[i] gives
h   
(F − ma) · δr = −mgk + 2mRΩ cos θθ̇i − mR cos θθ̈ − sin θ θ̇2 + Ω2 j

  i
−mR sin θθ̈ + cos θθ̇2 k · (R cos θδθj + R sin θδθk) = 0 [j]
After evaluating the dot product and setting the coefficient of δθ equal to zero, the
equation of motion is obtained as
g 
θ̈ + sin θ − Ω2 cos θ = 0 [k]
R

8.9 Hamilton’s Principle


From D’Alembert’s principle it is possible to develop scalar variational principles that pro-
vide a complete formulation of the problems of mechanics. These principles are called Hamil-
ton’s principles, and are named after the Irish mathematician, physicist, and astronomer Sir
William Rowan Hamilton (1805–1865), who first formulated them in the middle of the 19th
century. This text discusses one of the principles. The reader is referred to more advanced
texts on mechanics for the other principles attributed to Hamilton.
Consider a system of N particles and D’Alembert’s principle
N
X
(mi r̈i − Fi ) · δri = 0 (8.125)
i=1
P
The virtual work δW = Fi · δri is due to all forces external to the system. To manipulate
the acceleration term in the above equation, consider the expression
d
(ṙi · δri ) = r̈i · δri + ṙi · δ ṙi (8.126)
dt
The second term on the right of Equation (8.126) can be written as
1
ṙi · δ ṙi = δ (ṙi · ṙi ) (8.127)
2
where the similarity is clear to the kinetic energy of the i-th particle
1
Ti = mi ṙi · ṙi (8.128)
2
The variation of the kinetic energy of the i-th particle becomes
1
δTi = mi δ (ṙi · ṙi ) = mi ṙi · δ ṙi (8.129)
2
Analytical Mechanics 409

and we can express Equation (8.126) as


d
mi (ṙi · δri ) = mi r̈i · δri + mi ṙi · δ ṙi = mi r̈i · δri + δTi (8.130)
dt
The variation of the total kinetic energy is
N N N  
X X 1 X d
δT = δTi = mi δ (ṙi · ṙi ) = mi (ṙi · δri ) − r̈i · δri (8.131)
i=1 i=1
2 i=1
dt

Using Equation (8.131), D’Alembert’s principle is expressed as


N N
X X d
(mi r̈i − Fi ) · δri = 0 = −δT + mi (ṙi · δri ) − δW (8.132)
i=1 i=1
dt

which results in an expression for the variation of the kinetic and energy and virtual work
N
X d
δT + δW = mi (ṙi · δri ) (8.133)
i=1
dt

The next step is to integrate the above equation over two points in time, say, t1 and t2 ,
thus
Z t2 Z N
t2 X
d
(δW + δT ) dt = mi (ṙi · δri ) dt
t1 t1 i=1
dt

N
Z X N
t2
X
= mi d (ṙi · δri ) = mi ṙi · δri (8.134)


i=1 i=1 t1

Consider the right side of the above equation and Figure 8.17, which shows the true
and varied paths of a particle (for simplicity consider motion in one direction). If the varied
path is defined so that it vanishes at the endpoints t1 and t2 , the right side of Equation
(8.134) vanishes,9 leading to the relationship
Z t2
(δW + δT ) dt = 0 (8.135)
t1

which is known as the extended Hamilton’s principle or Hamilton’s principle. Writing the
virtual work in terms of its parts due to conservative and nonconservative forces as δW =
δWc + δWnc = δWnc − δV , we can write the extended Hamilton’s principle also as
Z t2
(δT − δV + δWnc ) dt = 0 (8.136)
t1

Derived here for a system of particles, the extended Hamilton’s principle is valid for
particles as well as for rigid or deformable bodies. It is, again, a fundamental principle of
mechanics from which the motion of all bodies can be described. In this sense, Hamilton’s
9 The case when the variation is selected so that the virtual displacement does not vanish at t and
1
t2 leads to another variational principle, Hamilton’s principle of varying action. The interested reader is
encouraged to consult more advanced texts for details.
410 Applied Dynamics

Varied path

True path

t1 t2 t

FIGURE 8.17
True path and varied path.

principle is not exactly a derived principle. Rather, it is more like a law of nature, in the
same way that Newton’s Second Law is a law of nature. Further, only scalar quantities such
as work and energy are needed, with no need to calculate accelerations.
At this point, we introduce the Lagrangian, L = T −V . For conservative systems δWnc =
0 and we can write Hamilton’s principle in terms of the Lagrangian as
Z t2
δL dt = 0 (8.137)
t1

The above relationship was first stated by Lagrange and originally called the principle of
least action. When the system is holonomic, we can interchange the integration and variation
operations, which yields
Z t2
δ L dt = 0 (8.138)
t1

The principle of least action states that Ramong all the paths that a system can take,
t
the actual path renders the definite integral t12 L dt stationary. This integral is also known
as the action integral. The term action implies the transfer of potential energy into kinetic
energy and vice versa. Of all paths that a body can take, it will take the one that involves
the least amount of transfer of energy from kinetic to potential or from potential to kinetic.
As with D’Alembert’s principle, Hamilton’s principle can be used for two purposes:
i) to develop additional variational principles, which we will do in the next section, or
ii) to obtain equations of motion, which we will do in the next example. Direct use of
Hamilton’s principle to obtain equations of motion of rigid bodies is not as efficient as
using D’Alembert’s principle or Lagrange’s equations. However, direct use of Hamilton’s
principle is a powerful tool to obtain the equations of motion of deformable bodies, as use
of Hamilton’s principle yields the equations of motion as well as the boundary conditions.
The reader is referred to vibration texts for more details.

Example 8.9
Obtain the equation of the bead problem in Example 8.8 using Hamilton’s principle.
From Example 8.8, the only force that contributes to the virtual work is gravity. Further,
there are no nonconservative forces, so δWnc = 0. We proceed with calculating the kinetic
and potential energies. From Example 8.8, the velocity of the bead is

v = −RΩ sin θi + R cos θθ̇j + R sin θθ̇k [a]


Analytical Mechanics 411

so the kinetic energy is


1 1 1
T = mv · v = mR2 Ω2 sin2 θ + mR2 θ̇2 [b]
2 2 2
Using the center of the ring as the datum, the potential energy becomes
V = −mgR cos θ [c]
so the Lagrangian has the form
1 1
L = T −V = mR2 Ω2 sin2 θ + mR2 θ̇2 + mgR cos θ [d]
2 2
The variation of the Lagrangian is
∂L ∂L h g i
δL = δθ + δ θ̇ = mR2 Ω2 sin θ cos θ − sin θ δθ + mR2 θ̇ δ θ̇ [e]
∂θ ∂ θ̇ R
The last term in Equation [e], mR2 θ̇δ θ̇, involves δ θ̇. To invoke Hamilton’s principle, it
is necessary to state all the expressions in terms of δθ. To accomplish this, the second term
is integrated by parts, which gives
Z t2 Z t2 t2 Z t2
d
θ̇δ θ̇ dt = θ̇ (δθ) dt = θ̇δθ − θ̈ δθ dt [f ]

t1 t1 dt t1 t1

The integrated term on the right side of Equation [f] vanishes by virtue of the definition
of the variation operation (the values of the variation at the beginning and end of the path
are zero). The second term on the right, when used with Equation [e] and the extended
Hamilton’s principle, yields
Z t2 h  g i
−mR2 θ̈ + mR2 Ω2 sin θ cos θ − sin θ δθ dt = 0 [g]
t1 R
In order for the equality to hold, the integrand must vanish at all times. Because δθ is
arbitrary, for the integrand to be zero the coefficient of δθ must be identically zero. This we
recognize as the equation of motion
g 
θ̈ + sin θ − Ω2 cos θ = 0 [h]
R
Let us review the operations that were carried out to obtain the equations of motion using
Hamilton’s principle. The kinetic and potential energies were calculated, and appropriate
variations and partial derivatives were taken, which yielded the θ̇δ θ̇ term. Integration by
parts of the θ̇δ θ̇ term ensured that all expressions are in terms of δθ. Given a general problem,
we can integrate by parts the general expression ∂L ∂ θ̇
δ θ̇ rather than the corresponding specific
terms (in this example θ̇δ θ̇). The question arises as to whether by manipulating Hamilton’s
principle we can perform the integrations by parts in advance and develop a general form
for the equations of motion. This question will be explored in the next section.

8.10 Lagrange’s Equations


From Hamilton’s principle, we derive Lagrange’s equations, which present themselves as a
convenient way of obtaining the equations of motion. The extended Hamilton’s principle
412 Applied Dynamics

can be expressed as
Z t2 Z t2 Z t2
(δT − δV + δWnc ) dt = δL dt + δWnc dt = 0 (8.139)
t1 t1 t1

The Lagrangian L can be written in terms of the generalized coordinates qk and gener-
alized velocities q̇k as L = L(q1 , q2 , . . . , qn , q̇1 , q̇k , . . . , q̇n , t). The variation of L is
n  
X ∂L ∂L
δL = δqk + δ q̇k (8.140)
∂qk ∂ q̇k
k=1

and, using Equation (8.86), the virtual work due to the nonconservative forces is
n
X
δWnc = Qknc δqk (8.141)
k=1

Making use of the property that the variation and differentiation (with respect to time)
operations can be interchanged, integration by parts of the ∂∂L q̇k δ q̇k term in Equation (8.140)
leads to
Z t2 Z t2 δqk (t2 ) Z t2  
∂L ∂L d ∂L d ∂L
δ q̇k dt = (δqk ) dt = δqk
− δqk dt (8.142)
t1 ∂ q̇k t1 ∂ q̇k dt ∂ q̇k δqk (t1 ) t1 dt ∂ q̇k

The integrated term requires the evaluation of δqk (k = 1, 2, . . . , n) at the beginning


and end of the time intervals. By definition of the variation of a displacement from the
previous section, the varied path vanishes at the end points, thus δqk (t1 ) = δqk (t2 ) = 0
for all values of the index k. Considering this and introducing Equations (8.140)–(8.142),
Hamilton’s principle becomes
Z t2 Z t2 n    
X d ∂L ∂L
(δT − δV + δWnc ) dt = − + + Qknc δqk dt = 0 (8.143)
t1 t1 dt ∂ q̇k ∂qk
k=1

For the integral over time to vanish at all times, the integrand must be identically equal
to zero, which can be expressed as
n    
X d ∂L ∂L
− + + Qknc δqk = 0 (8.144)
dt ∂ q̇k ∂qk
k=1

It should be noted that this equation can be directly obtained from D’Alembert’s principle,
without using Hamilton’s principle. Because of this, Equation (8.144) has been referred to
as the Lagrange’s form of D’Alembert’s principle.
The derivations so far have dealt with a set of independent generalized coordinates.
Because the generalized coordinates are independent, so are their variations. It follows that
the only way Equation (8.144) can be equal to zero is if the coefficients of δqk vanish
individually for all values of the index k. Setting the coefficients equal to zero, Lagrange’s
equations are obtained and they have the form
 
d ∂L ∂L
− = Qknc k = 1, 2, . . . , n (8.145)
dt ∂ q̇k ∂qk

Equation (8.145) is the most general form of Lagrange’s equations. These equations can
Analytical Mechanics 413

also be expressed in terms of the kinetic and potential energies. Noting that the potential
energy is not a function of the generalized velocities, we can write Equation (8.145) as
 
d ∂T ∂T ∂V
− + = Qknc k = 1, 2, . . . , n (8.146)
dt ∂ q̇k ∂qk ∂qk
This form of Lagrange’s equations is preferred by many, as it reduces the possibility
of making a sign error when evaluating the partial derivatives. It is also similar to the
format in which Lagrange first presented these equations in 1788. Born in Italy, Joseph-
Louis Lagrange (1736–1813) lived in Prussia and France. His doctoral adviser was Leonhard
Euler. Lagrange made significant contributions to classical and celestial mechanics, as well
as to number theory, analysis, and variational calculus.
Under certain circumstances, it is more convenient to write Lagrange’s equations in
terms of kinetic energy alone, in the form of
 
d ∂T ∂T
− = Qk (8.147)
dt ∂ q̇k ∂qk
where the generalized forces Qk contain contributions from the conservative as well as
nonconservative forces. The principle of virtual work, given by Equation (8.96), is a special
case of Lagrange’s equations.
Lagrange’s equations can conveniently be expressed in column vector format. Introduc-
ing the n-dimensional generalized coordinate and generalized force vectors

{q} = [q1 q2 . . . qn ]T {Qnc } = [Q1nc Q2nc . . . Qnnc ]T (8.148)

Lagrange’s equations can be written in column vector form as


 
d ∂L ∂L
− = {Qnc }T (8.149)
dt ∂{q̇} ∂{q}
Let us compare the steps involved in obtaining the equations of motion using Lagrange’s
equations and using the Newtonian approach. When using Newton’s Second Law, the pro-
cedure is to:
1. Determine the number of degrees of freedom, and select motion variables and coordinate
systems.
2. Isolate the different bodies involved and draw free-body diagrams.
3. Relate the sum of forces and sum of moments to the translational and angular acceler-
ations.
4. Eliminate the constraint and reaction forces by manipulating the force and moment
balances and derive the equations of motion. Use kinematics, if necessary, to express the
accelerations in terms of the motion variables.
When using the analytical approach (Lagrange, Hamilton), the procedure is to
1. Determine the number of degrees of freedom and select the motion variables (generalized
coordinates).
2. Identify the forces that contribute to the virtual work and determine those forces that
are conservative and those that are not. Isolating the different bodies involved and
drawing free-body diagrams is not necessary because interactions between the bodies is
accounted for in the method of solution. On the other hand, free-body diagrams help in
identifying the degrees of freedom and the external forces that act on the system.
414 Applied Dynamics

3. Use kinematical relationships to find the velocities and virtual displacements.


4. Write the expressions for the kinetic and potential energies, as well as the virtual work.
5. Apply Hamilton’s principle or Lagrange’s equations to obtain the equations of motion.
Obtaining equations of motion using D’Alembert’s principle involves steps from both
approaches. The forces that do not contribute to the virtual work are identified, as in
the analytical approach. On the other hand, accelerations and angular accelerations are
calculated, as in the Newtonian approach. In addition, the kinematic variables, that is, the
derivatives of the velocities and angular velocities with respect to the generalized velocities,
are calculated.
There are three distinct differences between the Newtonian and analytical approaches:
• Order of steps involved. In the Newtonian approach, we first write the force and
moment balances for all of the bodies separately and then use kinematical relationships
to eliminate the constraint forces. In the analytical approach, constraints are accounted
for by choosing the generalized coordinates appropriately.
• Expressions developed. For Lagrangian mechanics and Hamilton’s principle, velocity
expressions are developed; for the Newtonian approach and for D’Alembert’s principle,
expressions for acceleration are needed.
• Treatment of constraint forces. While in Newtonian mechanics we calculate the val-
ues of the constraint forces, analytical mechanics automatically eliminates the constraint
forces from the formulation and their magnitudes are not calculated.
It may appear that analytical approaches should be preferable to the Newtonian ap-
proach at all times; but this is not always the case. By eliminating the constraint forces
from the formulation, the analytical approach does not calculate the amplitudes of these
forces. While this may be acceptable for classroom examples, it is not in real-life applica-
tions, where knowing the amplitudes of the reaction and constraint forces is necessary in
order to properly design and build a system.
The best way to determine which approach is most suited for a particular problem is by
experience. Looking at a problem from both an analytical and a Newtonian view increases
physical insight and provides a better understanding of the system characteristics.

Example 8.10
Obtain the equation of motion for the bead problem in Examples 8.8 and 8.9 using La-
grange’s equations.
From Example 8.9, the kinetic and potential energies are
1 1 1
T = mv · v = mR2 Ω2 sin2 θ + mR2 θ̇2 V = −mgR cos θ [a]
2 2 2
so the Lagrangian has the form
1 1
L = T −V = mR2 Ω2 sin2 θ + mR2 θ̇2 + mgR cos θ [b]
2 2
As there are no nonconservative forces, the virtual work due to nonconservative forces
is zero, or δWnc = 0, so the generalized force is Qnc = 0. The partial derivatives are
 
∂L d ∂L ∂L
= mR2 θ̇ = mR2 θ̈ = mR2 Ω2 sin θ cos θ − mgR sin θ [c]
∂ θ̇ dt ∂ θ̇ ∂θ
Analytical Mechanics 415

The next step is to apply Lagrange’s equations, with the result

mR2 θ̈ − mR2 Ω2 sin θ cos θ − mgR sin θ = 0



[d]

Dividing by mR2 and collecting terms gives the equation of motion as


g 
θ̈ + sin θ − Ω2 cos θ = 0 [e]
R
Example 8.11
Consider the T-shaped bar in Example 8.5 and obtain its equations of motion using La-
grange’s equations.
The generalized coordinates are the displacement of point A along the incline s, and the
angle θ line AG makes with the horizontal. The kinetic energy has the form
1 1
T = IG θ̇2 + mvG
2
[a]
2 2
7 2
where, from Example 8.7, IG = 48 mL . The velocity of the center of mass was found earlier
as    
3 3
vG = ṡ cos 30◦ + Lθ̇ cos θ i + ṡ sin 30◦ + Lθ̇ sin θ j [b]
4 4
so that √
2 9 2 2 3 
vG = vG · vG = ṡ2 + L θ̇ + Lṡθ̇ 3 cos θ + sin θ [c]
16 4
and the expression for the kinetic energy becomes
1 2 17 3 √ 
T = mṡ + mL2 θ̇2 + mLṡθ̇ 3 cos θ + sin θ [d]
2 48 8
The potential energy has the form
   
3 1 3
V = mg s sin 30◦ − L cos θ = mg s − L cos θ [e]
4 2 4

There is one nonconservative force, F = F cos 15◦ i0 , and the constraint force N does no
work. The virtual work due to the nonconservative force is δWnc = F · δrA = F cos 15◦ δs.
The next step is to take the partial derivatives. For the generalized coordinate s, the
partial derivatives are
∂T 3 √ 
= mṡ + mLθ̇ 3 cos θ + sin θ
∂ ṡ 8

d ∂T 3 √  3  √ 
= ms̈ + mLθ̈ 3 cos θ + sin θ + mLθ̇2 − 3 sin θ + cos θ
dt ∂ ṡ 8 8
∂T ∂V 1
= 0 = mg [f ]
∂s ∂s 2
and the generalized force due to the nonconservative force is Qsnc = F cos 15◦ , so the
equation of motion associated with the generalized coordinate s becomes
3 √  3  √  1
ms̈ + mLθ̈ 3 cos θ + sin θ + mLθ̇2 − 3 sin θ + cos θ = − mg + F cos 15◦ [g]
8 8 2
416 Applied Dynamics

For generalized coordinate θ, the partial derivatives are


∂T 17 3 √ 
= mL2 θ̇ + mLṡ 3 cos θ + sin θ
∂ θ̇ 24 8

z }| {
d ∂T 17 3 √  3  √ 
2
= mL θ̈ + mLs̈ 3 cos θ + sin θ + mLṡθ̇ − 3 sin θ + cos θ
dt ∂ θ̇ 24 8 8
∂T 3  √  ∂V 3
= mLṡθ̇ − 3 sin θ + cos θ = mgL sin θ [h]
∂θ 8 ∂θ 4
d ∂T
Note that the term with an overbrace in dt ∂ θ̇
is the same as the expression for ∂T ∂θ , so
these two terms cancel each other when putting the terms together to obtain the equation
of motion. This is a common feature of Lagrange’s equations, and a source of their criticism
d ∂T ∂T
as some terms in the calculation of dt ∂ q̇k and ∂qk (k = 1, 2, ..., n) cancel each other, leading
to wasted manipulations.
The generalized force due to the nonconservative force is zero, Qθnc = 0, so the second
equation of motion becomes
17 3 √  3
mL2 θ̈ + mLs̈ 3 cos θ + sin θ + mgL sin θ = 0 [i]
24 8 4
which, of course, is the same as the equation of motion for θ in Example 8.7.
You should compare the algebraic effort to solve this problem using D’Alembert’s prin-
ciple and Lagrange’s equations and determine which one you prefer.

Example 8.12

g k! k"
m! m" m3
c! c"

x! G! x" G" x#
M, L " M, L
!

P! P"
k#

FIGURE 8.18
Mass-spring-damper system.

Consider the system in Figure 8.18 and calculate the kinetic and potential energies, as
well as the generalized forces associated with the nonconservative forces. The springs are
unstretched when the motion variables are all zero.
This system has five degrees of freedom and a suitable choice for the generalized coor-
dinates is x1 , x2 , x3 , θ1 , and θ2 . The kinetic energies of the carts are simply
1 1 1
T1 = m1 ẋ21 T2 = m2 ẋ22 T3 = m3 ẋ23 [a]
2 2 2
Analytical Mechanics 417

To find the kinetic energies of the rods, it is necessary to calculate the velocities of their
centers of mass. For the rod attached to mass 1, the position of the center of mass is
L
rG1 = x1 i + (sin θ1 i − cos θ1 j) [b]
2
so the velocity of the center of mass of the rod attached to mass 1 is
 
L L
vG1 = ẋ1 + θ̇1 cos θ1 i + θ̇1 sin θ1 j [c]
2 2

The associated kinetic energy becomes


1 1
T4 = M vG1 · vG1 + I1 θ̇12
2 2
" 2  2 #
1 L L 1 1
= M ẋ1 + θ̇1 cos θ1 + θ̇1 sin θ1 + M L2 θ̇12
2 2 2 2 12
 
1 2 1 2 2
= M ẋ1 + L θ̇1 + Lẋ1 θ̇1 cos θ1 [d]
2 3
The kinetic energy of the second rod is obtained the same way. Indeed, substituting x2
for x1 and θ2 for θ1 , we obtain for the second rod
 
1 1
T5 = M ẋ22 + L2 θ̇22 + Lẋ2 θ̇2 cos θ2 [e]
2 3
The total kinetic energy of the system is obtained by adding the individual kinetic energies
5
X
T = Ti [f ]
i=1

Five components contribute to the potential energy: the three springs and the weight of
the two rods. The potential energies of the two springs connecting the carts are
1 2 1 2
V1 = k1 (x2 − x1 ) V2 = k2 (x3 − x2 ) [g]
2 2
To find the potential energy associated with the spring that connects the two rods,
we need to calculate the tip positions of the rods. Ignoring the contribution to the spring
deflection of the vertical displacement of the tip points and denoting the horizontal distance
between the points P2 and P1 by ∆ yields

∆ = (rP2 − rP1 ) · i = x2 + L sin θ2 − (x1 + L sin θ1 ) [h]

so the potential energy associated with the bottom spring becomes


1 1 2
V3 = k3 ∆2 = k3 (x2 + L sin θ2 − x1 − L sin θ1 ) [i]
2 2
Considering the datum position as the pin joints on the carts, the gravitational potential
energies associated with the rods are simply
L L
V4 = −M g cos θ1 V5 = −M g cos θ2 [j]
2 2
418 Applied Dynamics

m1 . . m2 . . m3
c1(x 2 – x1) c2(x 3 – x 2)

x1 x2 x3

FIGURE 8.19
Free-body diagrams of carts showing only the damper forces.

The total potential energy is


5
X
V = Vi [k]
i=1

The next step is to find the generalized forces associated with the damping forces. To
this end, Figure 8.19 shows the free-body diagrams of the carts with only the damper forces.
As discussed in Chapter 4, damper forces are proportional to the difference in the velocities
of the carts. As the motion is one-dimensional, we can dispense with the vector location
and write the nonconservative work as

δWnc = F1 δx1 + F2 δx2 + F3 δx3 [l]

where

F1 = c1 (ẋ2 − ẋ1 ) F2 = −c1 (ẋ2 − ẋ1 ) + c2 (ẋ3 − ẋ2 ) F3 = −c2 (ẋ3 − ẋ2 ) [m]

are the damping forces acting on each mass. Since xi are the generalized coordinates qi , Fi
are the generalized forces Qi (i = 1, 2, 3) with Q4 = Q5 = 0.

8.11 Constrained Systems


As discussed earlier, it usually is preferable to work with a set of independent generalized
coordinates and generalized velocities. Sometimes, we must or we prefer to work with a set
of constrained coordinates. Listed below are cases where dealing with a set of constrained
coordinates is unavoidable or preferable:
• When the constraints are nonholonomic. Nonholonomic constraints involve velocity ex-
pressions that cannot be integrated to displacement expressions, so it may not be possi-
ble to find a set of generalized coordinates that lead to a set of unconstrained generalized
velocities.
• When the constraints are holonomic and we cannot eliminate the redundant (also re-
ferred to as surplus) coordinates with ease, for one or more of the following reasons:
– The constraint equation is complicated.
– Finding the transformations that lead to unconstrained equations makes the equa-
tions of motion lengthy and complex.
– Some of the forces acting on the system are functions of constraint forces. For
example, sliding friction forces are dependent on normal forces.
Analytical Mechanics 419

• When the constraints are holonomic but we do not want to eliminate the constraint
forces from the formulation, usually because of the need to know the amplitudes of the
reaction forces.
This section discusses analytical approaches that deal with constrained systems. The
next section will discuss Kane’s equations, which is another way to obtain equations of mo-
tion in terms of unconstrained variables, an approach of particular significance for systems
acted upon by nonholonomic constraints.
Consider a system described by m generalized coordinates, to which p equality con-
straints act, so that the number of degrees of freedom is n = m − p. The constraints are
expressed in the general velocity form as
n
X
ajk q̇k + aj0 = 0 j = 1, 2, . . . , p (8.150)
k=1

and the variations of the constraint equations are


n
X
ajk δqk = 0 j = 1, 2, . . . , p (8.151)
k=1

At this point, no distinction is being made regarding whether the constraints are holonomic
or nonholonomic.

8.11.1 Lagrange Multiplier Method


In the Lagrange multiplier method, the p constraint equations in Equation (8.151) are
multiplied by the Lagrange multipliers λj , (j = 1, 2, . . . , p) and summed, with the result
p X
X m
λj ajk δqk = 0 (8.152)
j=1 k=1

No restrictions on the magnitudes of the Lagrange multipliers are imposed at this stage.
Subtracting the above equation from the extended Hamilton’s principle in Equation (8.136)
gives
Z t2 Z t2 Z t2 p X
X m
δL dt + δWnc dt − λj ajk δqk dt = 0 (8.153)
t1 t1 t1 j=1 k=1

Following the same procedure used when deriving Lagrange’s equations for uncon-
strained systems, taking the appropriate partial derivatives, and performing the integration
by parts results in
 
Z t2 Xm   p
− d ∂L ∂L X
+ + Qknc − λj ajk  δqk dt = 0 (8.154)
t1 dt ∂ q̇k ∂qk j=1
k=1

In order for the above expression to be valid for any two time instances, the integrand must
be identically equal to zero so
 
m   p
X
− d ∂L ∂L X
+ + Qknc − λj ajk  δqk = 0 (8.155)
dt ∂ q̇k ∂qk j=1
k=1
420 Applied Dynamics

The generalized coordinates are no longer independent; hence, their variations are not
independent of each other either. At this point, a restriction on the Lagrange multipliers is
placed and their magnitudes are defined such that the coefficients of the virtual coordinates
δqk (k = 1, 2, . . . , m) vanish in Equation (8.155). This leads to a modified form of Lagrange’s
equations, written as
  p
d ∂L ∂L X
− + λj ajk = Qknc k = 1, 2, . . . , m (8.156)
dt ∂ q̇k ∂qk j=1

where λj ajk are the generalized constraint forces. They have the same units as the gener-
alized forces (which do not necessarily have the units of force).
The m equations in Equation (8.156) and the p constraint equations in Equation (8.150)
add up to a total of m + p = n + 2p equations that need to be solved for the unknowns
of m generalized coordinates q1 , q2 , . . . , qm and p Lagrange multipliers λ1 , λ2 , . . . , λp . The
resulting m + p equations are not a set of differential equations, as there is no derivative
of the Lagrange multipliers involved. Such equations are known as differential-algebraic
equations.
In column vector format, using the constraint equations in matrix form [a] {q̇} + {b} =
{0} and Equation (8.149), the Lagrange’s equations in the presence of constraints can be
written as
 
d ∂L ∂L
− + {λ}T [a] = {Qnc }T (8.157)
dt ∂ {q̇} ∂ {q}

where {λ} = [λ1 λ2 . . . λp ]T is the vector of Lagrange multipliers.


When the constraints are holonomic and expressed in terms of generalized coordinates
(also known as in configuration form), as

cj (q1 , q2 , . . . , qm , t) = 0 j = 1, 2, . . . , p (8.158)

a procedure similar to the one above is followed. The variation of the constraints (δcj ) are
taken, multiplied with the Lagrange multipliers λj , and summed and the resulting expression
p
X
λj δcj (q1 , q2 , . . . , qm ) (8.159)
j=1

is added to Hamilton’s principle. Taking the appropriate derivatives leads to the same result
as in Equation (8.156), where ajk are replaced by ∂cj /∂qk .
We can follow two approaches in solving the m + p equations discussed above:
1. Algebraically manipulate the m + p equations to eliminate the Lagrange multipliers and
obtain a set of n = m − p unconstrained equations of motion. This task is easier said
than done and may result in lengthy equations. We may also need to introduce new
motion variables to the problem in order to simplify the resulting equations.
2. Solve the m + p differential-algebraic equations directly. While cumbersome to do by
hand, this approach has become more popular in recent years because of the introduction
of powerful software that makes it possible to obtain a numerical solution.
The descriptions of constrained systems above are applicable for static as well as dynamic
systems. It should be noted that we can also introduce the Lagrange multiplier approach to
the D’Alembert’s principle. The procedure will be outlined in the next section, within the
context of Kane’s equations.
Analytical Mechanics 421

8.11.2 Constraint Relaxation Method


When the objective is to obtain the amplitude of a constraint force, an analytical approach
that can be used is the constraint relaxation method. A typical example of such cases is
problems involving friction. The friction force is a function of the normal force, which is
a constraint force. The constraint relaxation method is mathematically equivalent to the
Lagrange multiplier approach. However, it is more intuitive and particularly useful when
dealing with holonomic constraints expressed in terms of generalized coordinates.
In the constraint relaxation approach the constraint equation, which is a geometric or
kinematic relationship, is relaxed and we represent the effects of the constraint in terms
of an external force. The Lagrangian and the virtual work are written in terms of this
force. The constraint force enters the formulation via the virtual work and it appears in the
equations as part of the generalized forces. Invoking Lagrange’s equations leads to a set of
m equations.
The constraint equation is then imposed, which yields the equations of motion and also
permits calculation of the magnitude of the constraint force. The method is best illustrated
by means of an example.

Example 8.13
Figure 8.20 shows a collar of mass m sliding outside a long, slender rod of mass M and
length L. The coefficient of friction between the rod and collar is µ. There is a force F
acting at the tip of the rod. Find the equations of motion using the constraint relaxation
method.

a) b) Oy

O g
Ox
r r

N
m Ff
mg
L G Ff
M
r
Mg
y

P
F F x

FIGURE 8.20
a) Collar on a rod, b) free-body diagrams.

This system has two degrees of freedom and the generalized coordinates are selected as
the angle θ and the distance r of the collar from point O. The free-body diagrams of the
rod and collar are shown in Figure 8.20b. When using the constraint relaxation approach,
we treat the rod and collar separately. Let us describe the motion of the rod by the angle
φ and the motion of the collar by r and θ. The angles θ and φ are, of course, the same.
422 Applied Dynamics

Treating these two angles separately leads to a three-degrees-of-freedom system, to which


the constraint θ = φ is applied.
The kinetic and potential energies of the rod are
1 1
Trod = M L2 φ̇2 Vrod = − M gL cos φ [a]
6 2
and for the collar
1  2 
Tcollar = m ṙ + r2 θ̇2 Vcollar = −mgr cos θ [b]
2
The normal force N is included in the virtual work as acting on both the rod and collar.
The other two forces that contribute to the virtual work are the friction force and the force
at the tip of the rod. The virtual work due to the nonconservative forces becomes

δWnc = −Ff sign (ṙ) δr + F L sin ψ δθ + N r (δθ − δφ) [c]

Application of Lagrange’s equations gives

For r: mr̈ − mrθ̇2 − mg cos θ = −Ff sign (ṙ) [d]

For θ: mr2 θ̈ + 2mrṙθ̇ + mgr sin θ = N r [e]

1 1
For φ: M L2 φ̈ + M gL sin φ = −N r + F L sin ψ [f ]
3 2
These equations have to be combined with the constraint equation φ − θ = 0. Equation
[d] is one of the equations of motion. Adding Equations [e] and [f], and setting θ = φ
eliminates the constraint force N and results in
   
2 1 2 1
mr + M L θ̈ + 2mrṙθ̇ + mr + M L g sin θ = F L sin ψ [g]
3 2

Equation [e], when solved for N , gives the expression for the constraint force

N = mrθ̈ + 2mṙθ̇ + mg sin θ [h]

so that the friction force becomes


 
Ff = µN = µm rθ̈ + 2ṙθ̇ + g sin θ [i]

Introducing Equation [i] into Equation [d] gives the equation of motion for r as
 
mr̈ − mrθ̇2 − mg cos θ = −µm rθ̈ + 2ṙθ̇ + g sin θ sign (ṙ) [j]

8.12 Kane’s Equations


Previous sections discussed issues associated with dealing with constrained systems, and
the problems that arise when nonholonomic constraints are involved. The introduction of
generalized speeds (or quasi-velocities) was shown to facilitate the kinematic description of
Analytical Mechanics 423

systems acted upon by nonholonomic constraints, as well as in three-dimensional rotation


problems. This section explores the use of generalized speeds with D’Alembert’s principle.
Consider a system consisting of N rigid bodies which has m degrees of freedom. A set
of m dependent generalized coordinates q1 , q2 , . . . , qm is used to describe the motion, and
there are p constraints written in the general velocity form [a] {q̇} + {b} = {0}, where
[a] is the constraint matrix of order p × m. From the equations of motion obtained from
D’Alembert’s principle, Equation (8.122), and considering the derivations in the previous
section for constrained systems, the equations of motion, expressed in terms of the velocity
form of the kinematic variables, can be written as
N  
X ∂vGi   ∂ωωi
(Fi − mi aGi ) · + MGi − ḢGi · = Rk k = 1, 2, . . . , m (8.160)
i=1
∂ q̇k ∂ q̇k

where the contributions of the constraints to the equations of motion are accounted for by
adding the term Rk to the right side of Equation (8.122), where

Rk = {λ}T {ak } (8.161)

in which {λ} = [λ1 λ2 . . . λp ]T is an array of order p containing the Lagrange multipliers


and {ak } is the k-th column of [a], [a] = [{a1 } {a2 } . . . {am }].
Now consider n = m−p independent generalized speeds u1 , u2 , . . . , un . From Section 8.3
and Equation (8.32), we can relate the kinematic variables and partial velocities associated
with the i-th body using
   
∂vGi ∂vGi ∂vGi ∂vGi ∂vGi ∂vGi
... = ... [Y ] i = 1, 2, . . . , N (8.162)
∂ q̇1 ∂ q̇2 ∂ q̇m ∂u1 ∂u2 ∂un

where [Y ] is of order n × m and is defined in Equation (8.19). Considering all the bodies,
we can write
 ∂vG1 ∂vG1 ∂vG1   ∂vG1 ∂vG1 ∂vG1 
∂ q̇1 ∂ q̇2 ... ∂ q̇m ∂u1 ∂u2 ... ∂un
 ∂vG2 ∂vG2 ∂vG2   ∂vG2 ∂vG2 ∂vG2 
 ∂ q̇ ∂ q̇2 . . . 
∂ q̇m 
 ∂u1 ∂u2 ... ∂un

 =
 1   [Y ] (8.163)
 ...  ... 
   
∂vGN ∂vGN ∂vGN ∂vGN ∂vGN ∂vGN
∂ q̇1 ∂ q̇2 ... ∂ q̇m ∂u1 ∂u2 ... ∂un

or
q u
[vG ] = [vG ] [Y ] (8.164)

where the notation is obvious. Next, introduce the row vectors of order n

[F] = [(F1 − m1 aG1 ) (F2 − m2 aG2 ) . . . (FN − mN aGN )]

h     i
[M] = MG1 − ḢG1 MG2 − ḢG2 . . . MGN − ḢGN (8.165)

It follows that we can write Equations (8.160)–(8.161) in column vector format as


q
[F]T [vG ] + [M]T [ω
ω q ] − {λ}T [a]

= [F]T [vG
u
] [Y ] + [M]T [ω
ω u ] [Y ] − {λ}T [a] = {0}T (8.166)
424 Applied Dynamics

ω q ] and [ω
in which [ω ω u ] are the partial angular velocities discussed in Section 8.3, expressed
in array form and obeying the relationship [ω ω q ] = [ω
ω u ] [Y ].
Next, right multiply the above equation by [W ], which is of order m × n, and recall from
Equations (8.24)–(8.25) that [Y ] [W ] = [1], [a] [W ] = [0], so
[F]T [vG
u
] + [M]T [ω
ω u ] = {0}T (8.167)
which can be recognized as the column vector representation of the n equations of motion
in terms of the partial velocities. Hence, we can write the equations of motion in terms of
a set of independent generalized speeds as
N  
X ∂vGi   ∂ωωi
(Fi − mi aGi ) · + MGi − ḢGi · = 0 k = 1, 2, . . . , n (8.168)
i=1
∂uk ∂uk
or
N h
X   i
k k
(Fi − mi aGi ) · vG i
+ MG i
− ḢG i
· ω i = 0 k = 1, 2, . . . , n (8.169)
i=1

The above equations are known as Kane’s equations or Gibbs-Appell equations. Note
the ease with which it was possible to eliminate the constraints from the formulation and
to arrive at a set of unconstrained equations of motion. Kane’s equations are particularly
useful when dealing with nonholonomic constraints and in multibody systems undergoing
three-dimensional motion.
We note the following:
• The above equations are valid whether we begin with a set of dependent generalized
coordinates (and associated constraints) or, for unconstrained systems, with a set inde-
pendent generalized coordinates. For unconstrained systems, using a set of generalized
speeds other than generalized velocities may simplify the problem or enhance the anal-
ysis. A typical example is angular velocity components for three-dimensional motion.
• Following the definition of generalized forces in Equation (8.90), for a system with N
components, we can define generalized forces associated with the generalized speeds as
N  
X ∂vGi ωi
∂ω
Uk = Fi · + MGi · k = 1, 2, . . . , n (8.170)
i=1
∂uk ∂uk

which are in terms of resultant forces Fi and resultant moments MGi , or


0
N  
X ∂vi ωi
∂ω
Uk = Fi · + Mi · k = 1, 2, . . . , n (8.171)
i=1
∂uk ∂uk

which are in terms of individual forces Fi and individual moments Mi , where N 0 denotes
the total numbers of forces and moments. Note that the term Fi denotes different
quantities in the above two equations.
Writing these generalized forces in column vector form, {U } = [U1 U2 . . . Un ]T , and
considering the derivations above, the generalized forces associated with generalized
coordinates can be related to the generalized speeds by
{U }T = {Q}T [W ] or {U } = [W ]T {Q} (8.172)
where {Q} = [Q1 Q2 . . . Qm ]T , and it is helpful to recall that [W ] is of order m × n,
as defined in Equation (8.21).
Analytical Mechanics 425

• When using Kane’s equations, we first determine the velocity degrees of freedom, then
select a set of generalized speeds that are unconstrained, and then apply Equation
(8.168) directly.
• If desired, we can also write the Kane’s equations using a set of generalized speeds that
are constrained. The derivation for this formulation can be found in advanced texts,
such as Kane or Baruh.

• The Kane’s (or Gibbs-Appell) equations can also be derived from a scalar principle
that involves accelerations. The reader is encouraged to consult the texts by Baruh or
Ginsberg for details.

Example 8.14
Consider the vehicle in Example 8.2 and obtain its equations of motion by means of Kane’s
equations. The mass of the vehicle is m and mass moment of inertia about the center of
mass is IG .
The generalized speeds are u1 = vA , u2 = θ̇, and the velocity of point A is vA = vA i.
The velocity of the center of mass is

vG = vA + θ̇k × di = vA i + dθ̇j = u1 i + du2 j [a]

Differentiating Equation [a] gives the acceleration of the center of mass as


 
aG = v̇A i + dθ̈j + θ̇k × vA i + dθ̇j

   
v̇A − dθ̇2 i + dθ̈ + θ̇vA j = u̇1 − du22 i + (du̇2 + u1 u2 ) j

= [b]

The angular momentum is HG = IG θ̇k = IG u2 k, whose derivative is simply ḢG =


IG u̇2 k, and the resultant force and moment are
t
F = (FC + FD ) i MG = (FD − FC ) k [c]
2
The partial velocities are

1 ∂vG 2 ∂vG ω
∂ω
vG = = i vG = = dj ω1 = = 0
∂u1 ∂u2 ∂u1

ω
∂ω
ω2 = = k t
vG = 0 ωt = 0 [d]
∂u2
The Kane’s equations are invoked next. The force and moment balances are

F − maG = (FC + FD ) i − m u̇1 − du22 i − m (du̇2 + u1 u2 ) j




 
t
MG − ḢG = (FD − FC ) − IG u̇2 k [e]
2
For the first generalized speed, u1 = vA ,
∂vG   ∂ωω
= FC + FD − m u̇1 − du22 = 0

(F − maG ) · + MG − ḢG · [f ]
∂u1 ∂u1
426 Applied Dynamics

so the first equation of motion is

m u̇1 − du22

= FC + FD [g]

For the second generalized speed, u2 = θ̇,


∂vG   ∂ωω
(F − maG ) · + MG − ḢG ·
∂u2 ∂u2
 
t
= −md (du̇2 + u1 u2 ) + (FD − FC ) − IG u̇2 = 0 [h]
2
The second equation of motion becomes
t
IG + md2 u̇2 + mdu1 u2 = (FD − FC )

[i]
2
Next, let us calculate the generalized forces U1 and U2 using Equation (8.171). To this
end, we write the external forces that do work as

FD = FD i FC = FC i [j]

The velocities of points C and D are


   
t t
vC = vA + ω × rC/A = vA − θ̇ i vD = vA + ω × rD/A = vA + θ̇ i [k]
2 2

so that the partial velocities are

1 ∂vC 2 ∂vC t 1 ∂vD 2 ∂vD t


vC = = i vC = = − i vD = = i vD = = i [l]
∂u1 ∂u2 2 ∂u1 ∂u2 2
The generalized forces become

1 1 2 2 t
U1 = FC · vC + FD · v D = FC + FD U2 = FC · vC + FD · v D = (FD − FC ) [m]
2

8.13 Natural and Nonnatural Systems, Equilibrium


Consider the kinetic energy of a system of N particles that has n degrees of freedom
N
1X
T = mi vi · vi (8.173)
2 i=1

For the most general case, and as discussed earlier in this chapter, the position vector of
particle i can be expressed as

ri = ri (q1 , q2 , . . . , qn , t) (8.174)

The possibility that displacements ri (i = 1, 2, . . . , N ) may have an explicit dependance


on time is included in the above formulation. This may seem strange at first, as the gener-
alized coordinates should be sufficient to describe the evolution of the system completely.
Analytical Mechanics 427

Viewing ri as explicit functions of time implies that there is prescribed motion or that part
of the motion is treated as known.
From Newton’s Third Law, the motion of a body is related to the motion of every other
body with which it interacts; thus, in absolute reality such prescribed motion does not
(and cannot) exist. However, in several cases reasonable approximations can be made. For
example, the motion of a body on Earth is influenced by the motion of the earth. However,
the motion of the Earth is not affected by the motion of the body10 and hence can be
treated as known.
Treating part of the motion of a system as known leads to interesting results for the
expression of the kinetic energy. The equilibrium equations and stability properties are also
affected. The velocity of the i-th particle can be written as
dr ∂ri ∂ri ∂ri ∂ri
vi = = q̇1 + q̇2 + . . . + q̇n + (8.175)
dt ∂q1 ∂q2 ∂qn ∂t
or
n
X ∂ri ∂ri
vi = q̇k + (8.176)
∂qk ∂t
k=1

Introduction of the above equation into the expression for the kinetic energy, Equation
(8.173), results in
N
" n ! n
!#
1X X ∂ri ∂ri X ∂ri ∂ri
T = mi q̇k + · q̇s +
2 i=1 ∂qk ∂t s=1
∂qs ∂t
k=1

N n Xn n
!
1X X ∂ri ∂ri X ∂ri ∂ri ∂ri ∂ri
= mi · q̇k q̇s + 2 · q̇k + · (8.177)
2 i=1 s=1
∂qk ∂qs ∂qk ∂t ∂t ∂t
k=1 k=1

Denoting by
N N N
X ∂ri ∂ri X ∂ri ∂ri 1X ∂ri ∂ri
αks = mi · βk = mi · τ = mi · (8.178)
i=1
∂qk ∂qs i=1
∂qk ∂t 2 i=1 ∂t ∂t

where αks , βk , and τ are functions of the generalized coordinates and time, we can express
the kinetic energy as

T = T2 + T1 + T0 (8.179)

in which
n n n
1 XX X
T2 = αks q̇k q̇s T1 = βk q̇k T0 = τ (8.180)
2 s=1
k=1 k=1

Examining the terms that contribute to the kinetic energy, T2 is quadratic in the gener-
alized velocities, T1 is linear in the generalized velocities and T0 has no generalized velocity
terms. Both T1 and T0 are present because of the explicit dependence of the position vectors
10 Massive earthquakes have an effect on the orientation of the axis about which the Earth rotates. The

Indian Ocean earthquake in 2004 had just such an effect. The energy released during that earthquake was
equivalent to 1500 atom bombs like the one dropped on Hiroshima.
428 Applied Dynamics

ri on time. In rotating systems where the rotation of the system is treated as known, T1 is
related to Coriolis effects and T0 to centrifugal effects.
The kinetic energy can be expressed in matrix form as
1 T
T = {q̇} [M ] {q̇} + {β}T {q̇} + τ (8.181)
2
in which {q̇} = [q̇1 q̇2 . . . q̇n ]T is the generalized velocity vector, {β} = [β1 β2 . . . βn ]T is
referred to as the gyroscopic vector, and the matrix
 
α11 α12 . . . α1n
 α21 α22 . . . α2n 
[M ] =  . (8.182)
 
..
 ..

. 0 
αn1 αn2 ... αnn
is referred to as the mass matrix or the inertia matrix. We can show that [M ] is symmetric
and positive definite.
When T1 = T0 = 0, the system is called a natural system, and when T1 or T0 are not
zero, the system is called nonnatural. The name nonnatural is associated with the fact that
in a nonnatural system, a component of the motion is treated as, or assumed to be, known;
hence, the system being observed is artificial and not a natural one. Always keep in mind
the procedures used and the assumptions made when viewing a system as nonnatural and
evaluate whether the assumptions are realistic or not. Most systems that we deal with are
modeled as natural systems.
As an illustration of what constitutes a nonnatural system, consider a particle of mass m
sliding with speed q̇ along a wedge of angle σ, as shown in Figure 8.21. The wedge (of mass

q
m
. .
w q
M

x

FIGURE 8.21
Mass sliding on a moving wedge.

M ) is moving in the horizontal direction with speed ẇ. The total velocity of the particle is
v = ẇi + q̇ cos σi − q̇ sin σj (8.183)
If the velocity of the wedge is a known quantity, say ẇ = const, we can write the position
of the mass m as
r (t) = ẇti + q cos σi − q sin σj (8.184)
The explicit dependence of the position vector on time, r = r (q, t), is noted. The kinetic
energy becomes
1 1 1 h 2 2
i 1
T = mv · v + M ẇ2 = m (ẇ + q̇ cos σ) + (q̇ sin σ) + M ẇ2
2 2 2 2
Analytical Mechanics 429
1  2  1
= m q̇ + 2q̇ ẇ cos σ + ẇ2 + M ẇ2 (8.185)
2 2
When ẇ is known a priori the system is nonnatural with one degree of freedom and
1 1
T2 = mq̇ 2 T1 = mq̇ ẇ cos σ T0 = (m + M ) ẇ2 (8.186)
2 2
When w is treated as a variable, the system has two degrees of freedom and every term
in the kinetic energy is quadratic in the generalized velocities.
Nonnatural systems can be categorized into two distinct groups: i) when both T1 and
T0 are nonzero and ii) when only T0 is nonzero. The second case (T1 = 0, T0 6= 0) can be
treated as an otherwise natural system with an equivalent kinetic energy T2 and equivalent
potential energy U = V − T0 , referred to as modified potential energy or dynamic potential.
The Lagrangian for such a system can be written as

L = T − V = T2 + T0 − V = T2 − U (8.187)

The case when T1 6= 0 describes more complex problems, usually associated with rotating
systems. The response characteristics are quite different from when T1 = 0.
Let us reexamine equilibrium within the context of nonnatural systems. Static equilib-
rium was considered in Chapter 5 and in this chapter, and it was defined as the state at
which velocities and accelerations of all components of the system are zero. Because all
physical velocities and accelerations are zero, so too are the generalized velocities and their
time derivatives.
For a nonnatural system, equilibrium is defined as the state where all generalized ve-
locities and accelerations are zero. The velocities or accelerations of certain components of
the system are not zero at equilibrium, as part of the motion of a nonnatural system is
a known quantity. Consider an n-degrees-of-freedom system with generalized coordinates
q1 , q2 , . . . , qn and rewrite Lagrange’s equations as given in Equation (8.146)
d ∂ (T2 + T1 + T0 ) ∂ (T2 + T1 + T0 ) ∂V
− + = Qknc k = 1, 2, . . . , n (8.188)
dt ∂ q̇k ∂qk ∂qk
At equilibrium, all generalized velocities and accelerations are zero. In addition, all forces
and moments that act on the system that are functions of time disappear and the remaining
nonconservative forces are not functions of time, that is, Qknc 6= f (time). It follows that
Lagrange’s equations at equilibrium reduce to
∂T0 ∂V ∂U
− + = = Qknc k = 1, 2, . . . , n (8.189)
∂qk ∂qk ∂qk
For a natural system, T1 = T0 = 0 and the equilibrium relationship becomes
∂V
= Qknc k = 1, 2, . . . , n (8.190)
∂qk
which is the same as Equation (8.96).
The energy integral associated with nonnatural systems is denoted by the Jacobi integral
H and is defined by

H = T2 + U = T2 − T0 + V (8.191)

For a conservative nonnatural system, the Jacobi integral is constant and is a first integral
of the motion. For a natural system T0 = 0 and the Jacobi integral becomes the total energy
E =T +V.
430 Applied Dynamics

Example 8.15
Consider the bead-hoop system in Example 8.10 and find the equilibrium position.
The bead-hoop assembly represents a nonnatural system. From Example 8.10, the kinetic
and potential energies are
1 1 1
T = mv · v = mR2 Ω2 sin2 θ + mR2 θ̇2 V = −mgR cos θ [a]
2 2 2
and the components of the kinetic energy are identified as
1 1
T2 = mR2 θ̇2 T1 = 0 T0 = mR2 Ω2 sin2 θ [b]
2 2
We can calculate the equilibrium equation by using Equation (8.189) or by setting the
velocity and acceleration terms to zero in the equation of motion. The equation of motion
was obtained as g 
θ̈ + sin θ − Ω2 cos θ = 0 [c]
R
Let us use the second approach and set all first and second time derivatives of θ to zero.
The resulting equilibrium equation is
g 
sin θ − Ω2 cos θ = 0 [d]
R
Using the first approach, the equilibrium equation is obtained from
 
dU d (V − T0 ) d 1
= = −mgR cos θ − mR2 Ω2 sin2 θ = 0 [e]
dθ dθ dθ 2
Solving for the equilibrium equations yields the results
g  g 
sin θ = 0 =⇒ θe = 0, π − Ω2 cos θ = 0 =⇒ θe = ± cos−1 [f ]
R RΩ2
g g
The last equilibrium position is valid only when RΩ2 ≤ 1, or Ω2 ≥ R.

Example 8.16
Consider the bead-hoop system again. The Jacobi integral has the form
1 1
H = T2 + U = mR2 θ̇2 − mgR cos θ − mR2 Ω2 sin2 θ [a]
2 2
which obviously is different from the sum of the kinetic and potential energies of the bead.
It should be reiterated that for nonnatural systems the energy integral is not the sum of
the kinetic and potential energies. For this problem T + V is not constant.

Example 8.17
Consider Example 8.11 and calculate the mass (inertia) matrix that is used to describe the
kinetic energy.
From Equation [d] of Example 8.11, the kinetic energy is
1 17 3 √ 
T = mṡ2 + mL2 θ̇2 + mLṡθ̇ 3 cos θ + sin θ [a]
2 48 8
This is a natural system. The kinetic energy consists of T2 only, with T1 = T0 = 0
1
T = T2 = {q̇}T [M ] {q̇} [b]
2
Analytical Mechanics 431

Introducing the position vector {q} = [s θ]T , the inertia matrix becomes
"
3
√ #
m 8 mL 3 cos θ + sin θ
[M ] = 3
√  17 2
[c]
8 mL 3 cos θ + sin θ 24 mL

The inertia matrix is symmetric. We can show that the matrix is positive definite as
well, by evaluating its determinant. It should also be noted that the inertia matrix can be
directly obtained from the coefficients of the acceleration terms in the equations of motion.

8.14 Small Motions about Equilibrium


It is of interest to analyze the behavior of dynamical systems in the neighborhood of equi-
librium. This section extends the developments of Chapter 5 to analytical mechanics and
obtains linearized equations of motion in terms of generalized coordinates. The same proce-
dure as Chapter 5 is followed: the equations of motion are obtained, the equilibrium positions
denoted by qre (r = 1, 2, . . . , n) are found, and the equations of motion are linearized in the
neighborhood of equilibrium. Note that at equilibrium all generalized velocities are zero, so
q̇re = 0.
From the results of the previous section, and using Equations (8.189) and (8.190), the
equilibrium equations in the absence of nonconservative forces can be written as
∂V ∂ (V − T0 )
Natural systems: = 0 Nonnatural systems: = 0 (8.192)
∂qr ∂qr
Denoting the value of the potential energy V (or the dynamic potential U = V − T0 ) at
equilibrium by
Ve = V (q1e , q2e , . . . , qne ) (8.193)
we can write the Taylor series expansion of the potential energy as
n  
X ∂V
V (q1 , q2 , . . . , qn ) = Ve + (qr − qre )
r=1
∂qr e

n n
∂2V
 
1 XX
+ (qr − qre )(qs − qse ) + h.o.t. (8.194)
2 r=1 s=1 ∂qr ∂qs e

Without loss of generality, the datum, or reference position for the potential energy,
∂V
can be selected so that Ve = 0. Partial derivatives of the potential energy vanish, ∂q r
= 0,
∂U
at equilibrium ( ∂qr = 0 for nonnatural systems). Introducing the small quantities (local
variables) r = qr − qre (r = 1, 2, . . . , n), we can approximate the potential energy in
quadratic form as
n n 
∂2V

1 XX
V ≈ r s (8.195)
2 r=1 s=1 ∂qr ∂qs e

The same procedure is followed for nonnatural systems. Defining the stiffness coefficients
krs as
 2 
∂ V
Natural systems: krs =
∂qr ∂qs e
432 Applied Dynamics

∂ 2 (V − T0 )
 
Nonnatural systems: krs = (8.196)
∂qr ∂qs e

Introducing the local generalized coordinate vector {} = [1 2 . . . n ]T , the potential
energy (or modified potential energy) is expressed in quadratic form as
1
V ≈ {}T [K] {} (8.197)
2
in which [K] is known as the stiffness matrix, whose elements are krs . The stiffness matrix
is symmetric. Furthermore, if for a conservative system [K] is positive definite, the potential
energy (or dynamic potential) has a minimum at the equilibrium configuration. This can be
concluded by comparing the potential energy V with the Hessian that is calculated when
seeking the minimum value of a quadratic function.
A stationary value of a function of a single variable is a minimum if the second derivative
is positive at the stationary value. For a function of several variables, all first derivatives
must vanish at a stationary point. Also, to be considered a local minimum, the Hessian
must be positive definite. It should be reiterated that the results above are valid only for
small values of the coordinates r and for small motions around equilibrium.
As discussed in Chapter 5, a theorem from stability theory states that for a natural
conservative system, if the potential energy has a local minimum at equilibrium, then the
equilibrium position is critically stable, implying that if the system is disturbed from its
equilibrium position it hovers around equilibrium. In the presence of an energy dissipation
mechanism, the system returns to equilibrium. The corollary states that if the potential
energy does not have a local minimum about the equilibrium position, then the equilibrium
position is unstable. It follows that the equilibrium position of a natural conservative system
is critically stable if the stiffness matrix associated with that equilibrium position is positive
definite.
For nonnatural conservative systems the corresponding stability theorem depends on
the value of T1 . When T1 = 0, the system can be treated as an otherwise natural system
with kinetic energy T2 and potential energy U = V − T0 . An equilibrium position is stable
only if U = V − T0 has a minimum.
When T1 6= 0, an equilibrium position is still stable when U = V −T0 has a minimum, but
we cannot conclude that the equilibrium position is not stable when the dynamic potential
is not a minimum. This is because T1 generates gyroscopic (or Coriolis-like) terms, which
may enhance stability. Table 8.2 summarizes stability results for conservative systems.

TABLE 8.2
Summary of stability theorems for conservative dynamical systems

Is potential energy V (or U ) minimum?


Yes No
Natural systems Stable Unstable
Nonnatural systems
T1 = 0 Stable Unstable
T1 6= 0 Stable No conclusion

Consider linearization of the kinetic energy. For natural systems, the kinetic energy can
be written in the quadratic form
1
T = T2 = {q̇}T [M ] {q̇} (8.198)
2
Analytical Mechanics 433

in which [M ] is the inertia matrix or mass matrix, [M ] = [M (q1 , q2 , . . . , qn )]. The above
form is valid whether the kinetic energy is linear or nonlinear in terms of the generalized
coordinates. It follows that, for small motions about equilibrium, we can directly linearize
the kinetic energy as
1
T = T2 ≈ ˙ T [Me ] {}
{} ˙ (8.199)
2
in which the subscript e denotes that the inertia matrix is evaluated at equilibrium

[Me ] = [M (q1e , q2e , . . . , qne )] (8.200)

For small motions of a natural system about equilibrium, the Lagrangian can be written
as
1 T 1 T
L = T −V = {}
˙ [Me ] {}
˙ − {} [K] {} (8.201)
2 2
For nonnatural systems with T1 = 0 we can write L = T − V = T2 − U .
Lagrange’s equations in column vector format are given in Equation (8.149). Using the
properties of the derivative of a scalar with respect to a column vector, we can write
 
d ∂L ∂L
= {¨ }T [Me ] = −{}T [K] (8.202)
dt ∂ {} ˙ ∂{}
so that the linearized equations of motion of a natural system (or nonnatural system with
T1 = 0) can be expressed in matrix form as

[Me ] {¨
} + [K] {} = {Qnc } (8.203)

After calculating the kinetic and potential energies about equilibrium, we can use them
directly to obtain the equations of motion.
For nonnatural systems with T1 6= 0, we need to express T1 about equilibrium. Because
T1 is linear in the generalized velocities, it can be written as
n
X
T1 = βs q̇s = {β ({q})}T {q̇} (8.204)
s=1

where the coefficients βs are functions of the generalized coordinates, βs = βs (q1 , q2 , . . . , qn ).


It follows that when expanding T1 in terms of a Taylor series expansion, derivatives need to
be taken with respect to the generalized coordinates, as well as the generalized velocities.
The procedure is straightforward but cumbersome, so only the results are given here.11 The
quadratic approximation to T1 in the neighborhood of equilibrium becomes
n
X n X
X n
T1 ≈ βse ˙s + Brs ˙r ˙s (8.205)
s=1 r=1 s=1

where
   
∂T1 ∂βr
Brs = = (8.206)
∂qs ∂qr e ∂qs e

are evaluated at equilibrium. In column vector format

T1 ≈ {βe }T {} ˙ T [B] {}


˙ + {} (8.207)
11 Details can be found in the text Analytical Dynamics by Baruh.
434 Applied Dynamics

where the entries of [B] are Brs . Introducing T1 into the Lagrangian and writing the La-
grangian in column vector format results in
1 T 1 T
L = {} ˙ + {βe }T {}
˙ [Me ] {} ˙ T [B] {} − {} [K] {}
˙ + {} (8.208)
2 2
Taking the appropriate derivatives yields
 T  T
∂L ∂L
= [Me ] {}
˙ + {βe } + [B] {} = − [K] {} + [B]T {}
˙ (8.209)
∂ {}
˙ ∂{}
which leads to the equations of motion as
} + [B] − [B]T {}

[Me ] {¨ ˙ + [K] {} = {Qnc } (8.210)

˙ is [B] − [B]T . A matrix
The coefficient of the local generalized velocity vector {}
subtracted from its transpose results in a skew-symmetric matrix (a null matrix if the
original matrix is symmetric). Denoting [B]−[B]T by [G], where [G] is called the gyroscopic
matrix, the linearized equations of motion for a nonnatural system become
[Me ] {¨
} + [G] {}
˙ + [K] {} = {Qnc } (8.211)
For relative motion problems, when the motion of the relative frame is treated as known,
T1 leads to the Coriolis effect.
The first step associated with understanding the nature of the motion in the neighbor-
hood of equilibrium is to linearize the equations of motion about equilibrium and see if
the linearized equations imply significant behavior. If there is no significant behavior, then
higher-order analysis needs to be conducted, including the stability theorems summarized
in Table 8.2.
When studying multi-degrees-of-freedom systems in Chapter 7, the linearized equations
of motion comprised three matrices; mass, stiffness, and damping. There was no gyroscopic
matrix. Further, in the analysis above there is no damping matrix. To explain this difference,
recall that the types of problems considered in Chapter 7 did not include gyroscopic systems.
Also, viscous damping forces, such as the ones generated by dampers, are included here in
the generalized forces, as we cannot use the mass and stiffness matrices to describe them.
Once the equations of motion are obtained and linearized, the viscous damping effects
can be expressed in matrix form. The next section will discuss an addition to the Lagrangian
formulation, which makes it possible to have a matrix representation of viscous damping
forces.

Example 8.18
Consider the bead-hoop system in Example 8.15, find the equilibrium positions and analyze
their stability.
Because the system is nonnatural with T1 = 0, if U = V − T0 has a local minimum
at equilibrium, the equilibrium position is stable. Otherwise it is unstable. To investigate
stability, it is necessary to examine the second derivative of the dynamic potential at equi-
librium. The second derivative must be greater than zero for a local minimum.
From Example 8.15, the dynamic potential is
1
U = V − T0 = −mgR cos θ − mR2 Ω2 sin2 θ [a]
2
The first derivative of U gives the equilibrium equation obtained earlier:
dU
= mgR sin θ − mR2 Ω2 sin θ cos θ = 0 [b]

Analytical Mechanics 435

Differentiating Equation [b] with respect to θ yields the second derivative as

d2 U
= mgR cos θ + mR2 Ω2 sin2 θ − cos2 θ

2
[c]

2
Next, let us evaluate ddθU2 at the equilibrium positions. The equilibrium positions were
shown to be θe = 0, π, ± cos−1 g/RΩ2 . Divide the above expression by mR2 . For θe = 0


the second derivative becomes


1 d2 U

g
= − Ω2 [d]
mR2 dθ2 θe =0 R
g
so θe = 0 is stable when Ω2 < R . The physical explanation is that for low rotational speeds
of the hoop, the bead stays at the bottom.
For θe = π,
1 d2 U

g
= − − Ω2 < 0 [e]
mR2 dθ2 θe =π R
so θe = π is unstable at all rotational speeds. This is to be expected because at the equilib-
rium position θe = π the bead is at the top of the hoop. q
g g
2
For the third and fourth equilibrium positions, cos θe = RΩ 2 and sin θe = ± 1 − RΩ 2

so that  g 2
sin2 θe − cos2 θe = 1 − 2 [f ]
RΩ2
The second derivative, when evaluated at equilibrium, becomes

1 d2 U
  g 2 
2
= Ω 1− [g]
mR2 dθ2 θe =± cos−1 ( g ) RΩ2
RΩ2

g g
2 g
so that the equilibrium positions ± cos θe = RΩ 2 are stable when 1− RΩ2 > 0 or Ω2 > R .
Table 8.3 summarizes the stability results as a function
p of the rotational speed of the
hoop. The angular velocity has a critical value of Ω = g/R that dictates p which equilibrium
point is stable. When the angular velocity is in the range 0 < Ω < pg/R, the stable
equilibrium position is θe = 0. When the angular velocity goes over g/R, the stable
g
equilibrium point becomes θe = ± cos−1 ( RΩ 2 ). The equilibrium point moves higher with

increasing angular velocity, approaching θe = ±π/2 for high values of the angular velocity
Ω.

TABLE 8.3
Stability results for bead-hoop problem

g g
Equilibrium Position Ω2 < R Ω2 > R
θe = 0 Stable Unstable
θe = π Unstable Unstable
g
θe = ± cos−1

RΩ2 Unstable Stable
436 Applied Dynamics

Example 8.19
Consider Example 8.11, find the equilibrium positions, and linearize the equations of motion
about equilibrium.
From Example 8.11, the kinetic and potential energies are
1 17 3 √ 
T = mṡ2 + mL2 θ̇2 + mLṡθ̇ 3 cos θ + sin θ [a]
2 48 8
   
3 1 3
V = mg s sin 30◦ − L cos θ = mg s − L cos θ [b]
4 2 4
and the virtual work due to the nonconservative force F is δW = F · δrA = F cos 15◦ δs.
The inertia matrix is
"
3
√ #
m 8 mL 3 cos θ + sin θ
[M ] = 3
√  17 2
[c]
8 mL 3 cos θ + sin θ 24 mL

and the equations of motion are


3 √  3  √  1
ms̈ + mLθ̈ 3 cos θ + sin θ + mLθ̇2 − 3 sin θ + cos θ = − mg + F cos 15◦ [d]
8 8 2

17 3 √  3
mL2 θ̈ + mLs̈ 3 cos θ + sin θ + mgL sin θ = 0 [e]
24 8 4
We can find the equilibrium positions two ways. In the first, the potential energy is
calculated, its derivatives with respect to the generalized coordinates are taken, and use is
made of Equation (8.190). The second approach considers the equations of motion and sets
all derivative terms involving the generalized coordinates to zero.
Derivatives of the potential energy w.r.t. the generalized coordinates are
∂V ∂V 3
= mg sin 30◦ = mgL sin θ [f ]
∂s ∂θ 4
and the generalized forces are Qs = F cos 15◦ , Qθ = 0, leading to the equilibrium equations
mg sin 30◦ = F cos 15◦ sin θ = 0 [g]
The equilibrium condition for s simply indicates that the acting force F should be
sufficient to counter the effect of gravity. We can then measure s from any point where
the cart is at rest and set se = 0. The equilibrium equation for θ gives two solutions,
θe = 0, θe = π. The realistic (and stable) position is θe = 0.
Let us linearize the equations of motion about se = 0, θe = 0. Because values of the
generalized coordinates at equilibrium are zero, we can use a small angles assumption of
cos θ ≈ 1, sin θ ≈ θ. Noting that
√  √  √
s̈ 3 cos θ + sin θ ≈ s̈ 3 + θ ≈ 3s̈ θ̇2 ≈ 0 [h]

the linearized equations of motion are obtained as


√ √
3 3 1 17 3 3 3
ms̈ + mLθ̈ = − mg + F cos 15◦ mL2 θ̈ + mLs̈ + mgLθ = 0 [i]
8 2 24 8 4
The corresponding inertia matrix is
" √ #
3 3
m 8 mL
[M ] = √ [j]
3 3 17 2
8 mL 24 mL
Analytical Mechanics 437

The inertia matrix can also be obtained directly from Equation [c] by substituting the
values of the equilibrium positions into the inertia matrix.
It should be noted that when linearizing about equilibrium, one is linearizing also about
ṡ = 0, s̈ = 0, θ̇ = 0, and θ̈ = 0. This is the reason the term θ̇2 in Equation [h] vanishes,
as the linearization of a quadratic function about zero is zero. For example, f (x) = x2
linearized about x = 0 is f (x) ≈ 0.

8.15 Rayleigh’s Dissipation Function


An important class of nonconservative forces is forces that dissipate energy. As discussed
in Chapters 4 and 5, the primary models of dissipative forces are friction, viscous damping,
aerodynamic and hydrodynamic drag, and hysteresis. This section develops a convenient
way of treating viscous damping forces in Lagrangian mechanics by means of Rayleigh’s
dissipation function.
A body moving slowly in a viscous medium experiences a resistive force whose magnitude
is proportional to the velocity of the body. This resistive force is referred to as the viscous
damping force. Denoting the displacement of the body by x, the linear approximation to the
viscous damping force has the form Fd = −cẋ, in which c is the viscous damping coefficient.
The viscous damping model is discussed in Chapter 4.
The Rayleigh’s dissipation function F is a quadratic function that is defined as
1 2
F = cẋ (8.212)
2
and the associated generalized force Q can be obtained from it by
dF
Q = − (8.213)
dẋ
Note the similarity between Rayleigh’s dissipation function and the expressions for po-
tential energy of springs (kx2 /2) and kinetic energy (mẋ2 /2). Consider a viscous damping
force generating component, such as a damper (shock absorber) in a vehicle is used to con-
nect two bodies, as shown in Figure 8.22a. The damping force acting on each body (Figure
8.22b) is Fd = c (ẋ2 − ẋ1 ), and the Rayleigh’s dissipation function has the form

a) b)
x1 x2
Fd Fd
m1 m2 m1 m2
c . .
c(x 2 – x 1)

FIGURE 8.22
a) Two bodies connected by a damper. b) Free-body diagrams.

1 2
F = c (ẋ2 − ẋ1 ) (8.214)
2
The Rayleigh’s dissipation function can be written in terms of the generalized velocities
438 Applied Dynamics

as
n n
1 XX 1 T
F = drs q̇r q̇s = {q̇} [D] {q̇} (8.215)
2 r=1 s=1 2

where [D] is the damping matrix, whose entries are dks . We can show that the damping
matrix is symmetric and positive definite (or positive semi-definite).
The contribution of viscous damping forces to Lagrange’s equations can be obtained by
differentiating Rayleigh’s dissipation function with respect to the generalized velocities
n
∂F X
= drs q̇s (8.216)
∂ q̇r s=1

We can then express Lagrange’s equations in the presence of viscous damping forces as
 
d ∂L ∂L ∂F
− + = Qrnc r = 1, 2, . . . , n (8.217)
dt ∂ q̇r ∂qr ∂ q̇r
where Qknc no longer include contributions from viscous damping forces.
For small motions about equilibrium, because the form of the Rayleigh’s dissipation
function is similar to the kinetic energy ( 21 {q̇}T [M ]{q̇}), we can linearize the Rayleigh’s
dissipation function by evaluating [D] at the equilibrium position. Denoting this matrix by
[De ], the Rayleigh’s dissipation function can be written in matrix form as
1 T
F ≈ {}
˙ [De ] {}
˙ (8.218)
2
It follows that the equations of motion in the neighborhood of equilibrium can be ex-
pressed as
[Me ] {¨
} + ([G] + [De ]) {}
˙ + [K] {} = {Qnc } (8.219)
Let us examine the units of the Rayleigh’s dissipation function F. Consider a single
damper so that F = cẋ2 /2. The damping force is Fd = −cẋ. It follows that the unit of
Rayleigh’s dissipation function is force × velocity, or power.

Example 8.20
Consider Example 8.12 and write the associated Rayleigh’s dissipation function.
There are two dampers, of constants c1 and c2 , and the relative speeds that affect the
dampers are ẋ2 − ẋ1 and ẋ3 − ẋ2 . The Rayleigh’s dissipation function is
1 2 2

F = c1 (ẋ2 − ẋ1 ) + c2 (ẋ3 − ẋ2 ) [a]
2
Recalling that the generalized coordinates associated with the three carts are qi = xi (i =
1, 2, 3), q4 = θ1 , and q5 = θ2 , the Rayleigh’s dissipation function can be written in matrix
form as
1
F = {q̇}T [D]{q̇} [b]
2
where the associated damping matrix is
 
c1 −c1 0 0 0
 −c1 c1 + c2 −c2 0 0 
 
[D] =   0 −c2 c2 0 0   [c]
 0 0 0 0 0
0 0 0 0 0
Analytical Mechanics 439

8.16 Generalized Momentum, First Integrals


First integrals, also known as integrals of the motion, are expressions that are obtained
by integrating the describing equations of a system once (hence, the name first integral).
They are helpful for analyzing the behavior of a dynamical system qualitatively, and they
give insight into the system behavior without solving for the response. For example, for
conservative systems, total energy is a first integral, as is the Jacobi integral H discussed
in Section 8.13 for nonnatural systems. Linear momentum and angular momentum are first
integrals when they are conserved.
Given a describing equation, such as an equation of motion, if we can integrate that
equation to find an expression whose value is constant, then that expression is a first integral.
This section is concerned with first integrals associated with Lagrangian mechanics.
Define by πk the generalized momentum associated with the k-th generalized coordinate
as
∂L
πk = k = 1, 2, . . . , n (8.220)
∂ q̇k
The relationship between the generalized coordinates and generalized momenta is similar
to the relationship between a translational coordinate and linear momentum or between a
rotation angle and angular momentum. Because the potential energy does not contain any
terms in the generalized velocities, we can express the generalized momenta as
∂L ∂T
πk = = (8.221)
∂ q̇k ∂ q̇k
Consider now a system where the l-th generalized coordinate is absent from the for-
mulation. Such a coordinate is referred to as cyclic or ignorable. The name cyclic is due
to the fact that such coordinates are encountered mostly in systems with rotating compo-
nents. It follows that in the l-th equation of motion, the relationship ∂L/∂ql = 0 holds and
Lagrange’s equations become
 
d ∂L
= Qlnc (8.222)
dt ∂ q̇l

In the special case when the generalized force associated with the ignorable coordinate
is zero we can write
 
d ∂L d
= πl = 0 (8.223)
dt ∂ q̇l dt

leading to the conclusion that πl is constant. Hence, the generalized momentum associated
with a cyclic coordinate is an integral of the motion when the associated generalized force
is zero.
Unfortunately, a set procedure for identifying integrals of the motion does not exist.
You should begin with examining energy and linear and angular momentum, ascertain the
presence of nonconservative forces, and look for constants involving position and velocity
expressions.
440 Applied Dynamics

8.17 Impulsive Motion


This section explores impulse-momentum relationships when a system is described using
analytical mechanics. We consider applications to Lagrange’s equations as well as Kane’s
equations.
The linear impulse-momentum relationship for a particle of mass m acted upon by a
force F is obtained by integrating Newton’s Second Law over time from t1 to t2
Z t2
mv (t1 ) + F (t) dt = mv (t2 ) (8.224)
t1

As discussed in Chapter 5, an impulsive force is one that is of very high magnitude


applied over a very short period of time. An impulsive force is mathematically idealized
by means of the Dirac delta function. An ideal impulsive force causes a sudden change in
velocity, with no change in position.

8.17.1 Impulsive Excitation in Lagrangian Mechanics


Consider Lagrange’s equations, an impulsive force applied at time t0 and an impulse dura-
tion of . Integrating Lagrange’s equations over time and taking the limit as  approaches
zero gives
Z t0 +    
d ∂T ∂T ∂V
lim − + − Qknc dt = 0 (8.225)
→0 t
0
dt ∂ q̇k ∂qk ∂qk

Evaluating each of the terms above and noting from Equation (8.220) the definition of
the generalized momentum πk gives
Z t0 +   Z t0 +  
d ∂T ∂T t +
lim dt = lim d = lim (πk )|t00 = ∆πk (8.226)
→0 t
0
dt ∂ q̇ k →0 t0 ∂ q̇ k →0

Z t0 +  
∂T ∂V
lim − + dt = 0 (8.227)
→0 t0 ∂qk ∂qk

Z t0 +
lim Qknc dt = Q̂k k = 1, 2, . . . , n (8.228)
→0 t0

Equation (8.227) is equal to zero because both ∂T /∂qk and ∂V /∂qk have finite magni-
tudes. The term Q̂k is defined as the generalized impulse. The only forces that contribute to
the generalized impulse are impulsive forces; the contribution of all other forces disappears
as the duration of the impulse becomes infinitesimally small. The generalized impulse can
be obtained from the impulsive forces by
N N
X ∂ri X ∂vi
Q̂k = F̂i · = F̂i · (8.229)
i=1
∂qk i=1
∂ q̇k

It follows that the change in the generalized momentum is equal to the generalized impulse,
or

∆πk = Q̂k k = 1, 2, . . . , n (8.230)


Analytical Mechanics 441

Equation (8.230) can be expressed in matrix form. Considering the matrix form of the
kinetic energy in Equation (8.181)
1 T
T = {q̇} [M ] {q̇} + {β}T {q̇} + τ (8.231)
2
the generalized momentum becomes
∂T
{π}T = [π1 π2 . . . πn ] = = {q̇}T [M ] + {β}T (8.232)
∂{q̇}
or {π} = [M ] {q̇} + {β}. Considering that the inertia matrix and the vector {β} do not
change during the impulse (both are functions of position variables only), we can write the
change in generalized coordinates during impulse as

{∆π} = [M ] {∆q̇} = {Q̂} (8.233)

in which {Q̂} = [Q̂1 Q̂2 . . . Q̂n ]T is the generalized impulse vector. Hence, given the
impulsive forces, we can calculate the generalized impulses, and by inverting Equation
(8.233), we calculate the generalized velocities immediately after the impulse as

{∆q̇} = [M ]−1 {Q̂} (8.234)

8.17.2 Impulse-Momentum Relationships for Kane’s Equations


Here, the formulation above is extended to generalized speeds (quasi-velocities). The gen-
eralized momentum associated with a generalized speed is defined as
∂T
πk0 = (8.235)
∂uk
For a system of N particles we can show that the generalized momenta have the form
N
X
πk0 = mi vi · vik k = 1, 2, . . . , n (8.236)
i=1

∂vi
where vik = ∂u k
are the partial velocities introduced in Section 8.3.
To relate the two generalized momenta πk and πk0 note from Equation (8.21) that {q̇} =
[W ] {u} + {X}, so ∂{q̇}/∂{u} = [W ]. Also, considering from Equation (8.232) that {π}T =
∂T /∂{q̇}, we invoke the chain rule of differentiation, resulting in
∂T ∂T ∂{q̇}
{π 0 }T = = = {π}T [W ] (8.237)
∂{u} ∂{q̇} ∂{u}
so that
T
{π 0 } = [W ] {π} (8.238)

The generalized impulses associated with the generalized speeds are defined as
Z t2
Ûk = Uk dt (8.239)
t1

From Equation (8.172) we can relate the generalized impulses associated with the general-
ized speeds to those associated with the generalized coordinates by
T
{Û } = [W ] {Q̂} (8.240)
442 Applied Dynamics

When acted upon by impulsive forces, the change in the generalized momenta associated
with generalized speeds are related to the generalized impulses by

{∆π 0 } = {Û } (8.241)

which is analogous to its counterpart associated with generalized velocities. For nonholo-
nomic systems, it is more convenient to use Equation (8.241) than Equation (8.230).
We can also express the above equation in terms of the generalized speeds. Introducing
the relationship {q̇} = [W ] {u} + {X} and Equation (8.240) into Equation (8.233), and
making the same assumption that position variables do not change during the impulse,
gives

{∆π 0 } = [M 0 ] {∆u} = {Û } (8.242)

where [M 0 ] = [W ]T [M ] [W ] is the inertia matrix associated with the generalized speeds.


Inverting Equation (8.242) gives the generalized speeds immediately after an impulsive force
is applied.

Example 8.21
The massless collar in Figure 8.23 is free to slide over a guide bar. Attached to the collar
with a pin joint is a rod of mass m and length L. A ball of mass M is attached to the
tip of the rod. The system is at rest with θ = 0 when a horizontal impulsive force F̂ is
applied to the mass at the tip. Find the velocity of the collar and angular velocity of the
rod immediately after the impulsive force is applied.

a) b)
x O Nx y
N

G x
g
m, L
x + L sin
2
mg F^
F^
M
Mg

FIGURE 8.23
a) Rod-collar system, b) free-body diagram.

This system has two degrees of freedom. The generalized coordinates are selected as the
translation of the collar x and the rotation angle θ. For this problem, it is not necessary to
obtain the equations of motion; we need only to calculate the kinetic energy and the virtual
work due to the impulsive force. The kinetic energy is due to the kinetic energy of the ball
and the kinetic energy of the rod:
1  2 2
 1 
2 2
 1
T = M vM + vM + m vG + vG + IG θ̇2 [a]
2 x y
2 x y
2
From Figure 8.23b, we express the displacements of the center of mass and of the ball
Analytical Mechanics 443

as
L L
xG = x + sin θ yG = − cos θ xM = x + L sin θ yM = −L cos θ [b]
2 2
Differentiating the above terms gives the velocity components in the x and y directions
as
L L
vGx = ẋ + θ̇ cos θ vGy = θ̇ sin θ vMx = ẋ + Lθ̇ cos θ vMy = Lθ̇ sin θ [c]
2 2
1
Substituting Equation [c] into Equation [a] and noting that IG = 12 mL2 gives the kinetic
energy as 
1 2  2 
T = M ẋ + Lθ̇ cos θ + Lθ̇ sin θ
2
" 2  2 #
1 L L 1
+ m ẋ + θ̇ cos θ + θ̇ sin θ + mL2 θ̇2 [d]
2 2 2 24
Next, take the derivatives of the kinetic energy withe respect to the generalized velocities,
which yields
∂T 1
= M ẋ + M Lθ̇ cos θ + mẋ + mLθ̇ cos θ
∂ ẋ 2
∂T 1 1
= M L2 θ̇ + M Lẋ cos θ + mL2 θ̇ + mLẋ cos θ [e]
∂ θ̇ 3 2
It is not necessary to consider the potential energy for impulsive motion analysis and
when calculating the virtual work we need to take only the impulsive force into account.
The virtual work associated with the impulsive force is

δW = F δxM = F δ (x + L sin θ) = F δx + F L cos θδθ = Qx δx + Qθ δθ [f ]

so the generalized forces are Qx = F and Qθ = F L cos θ. It follows that the generalized
impulses are
Q̂x = F̂ Q̂θ = F̂ L cos θ [g]
At the point of application of the impulse, the rod is vertical and θ = 0. Combining
Equations [e] and [g] and setting θ = 0, the two equations for the two unknowns ẋ and θ̇
are obtained as  
1
(M + m) ẋ + M L + mL θ̇ = F̂
2
   
1 2 1 2
M L + mL ẋ + M L + mL θ̇ = F̂ L [h]
2 3
The impulsive motion equations are coupled, so they need to be solved simultaneously.
Doing so, the velocities immediately after the impulse are obtained as

2F̂ 6F̂
ẋ = − θ̇ = [g]
4M + m (4M + m) L

Note that the velocity of the collar immediately after the impulse is in the opposite
direction of the impulse (similar to the whiplash reaction encountered when a car is rear-
ended by another vehicle).
444 Applied Dynamics

Example 8.22
The cart in Example 8.14 is at rest when it is acted upon by an impulsive force F̂D with
F̂C = 0 in Figure 8.7. Find the resulting velocity of point A and angular velocity of the
cart.
From Example 8.14, the equations of motion in terms of the generalized speeds u1 = vA
and u2 = θ̇ are
m u̇1 − du22 = FC + FD

[a]
and
IG + md2 u̇2 + mdu1 u2 = (FD − FC ) h

[b]
To obtain the impulse equations, replace u̇1 and u̇2 by ∆u1 and ∆u2 in Equation [a]
and Equation [b], FD by F̂D , set FC = 0, and eliminate all the other terms in the equations
of motion that do not have derivatives, with the result

IG + md2 ∆u2 = F̂D h



m∆u1 = F̂D [c]

For this problem the impulsive motion equations are uncoupled, so they can be solved
independently with the result

F̂D F̂D h
∆u1 = ∆vA = ∆u2 = ∆ω = [d]
m (IG + md2 )

Because the force hitting the cart is applied at the right of the rear side, the resulting
velocity is positive (forward) and the angular velocity is counterclockwise.

8.18 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Benaroya, H., and Nagurka, M., Mechanical Vibration, 3rd Edition, CRC Press, 2009.
Bottega, W.J., Engineering Vibrations, CRC Press, 2006.
Ginsberg, J., Engineering Dynamics, 3rd Edition, Cambridge University Press, 2007.
Greenwood, D.T., Principles of Dynamics, 2nd Edition, Prentice-Hall, 1988.
Kane, T.R., and Levinson, D.A., Dynamics: Theory and Applications, McGraw-Hill, 1985.
Meirovitch, L., Methods of Analytical Dynamics, McGraw-Hill, 1970.

8.19 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 8.2—Generalized Coordinates


8.1 (M) The tip of the double link mechanism in Figure 8.24 is constrained to lie on the
inclined plane. Derive the constraint equation. Then, express the constraint in velocity form.
Analytical Mechanics 445

P
2L
y
B 2
L s

O 1 
x
3L/2

FIGURE 8.24
Figure for Problems 8.1 and 8.44.

8.2 (M) A system with generalized coordinates q1 and q2 is subjected to the constraint in
velocity form
q2
 
3q1 sin q2 + 2 + 2 q̇1 + q12 cos q2 + 2q2 q̇2 = 0

q1
Show that this constraint is holonomic. Hint: Write the constraint as a1 q̇1 + a2 q̇2 = 0 and
look for an integrating factor g(q1 , q2 ) such that ∂f /∂qk = gak (k = 1, 2).
8.3 (M) The radar tracking of a moving vehicle by another moving vehicle is a common
problem. Consider the vehicles A and B in Figure 8.25. The orientation of vehicle A must
always be toward vehicle B. Express the constraint relationship between velocities and show
that the constraint is nonholonomic.

y
B
vB

yB
vA
A 
yA xB
x
xA

FIGURE 8.25
Figure for Problem 8.3.

8.4 (M) Find the constraint equation for the front wheel (point B) of the bicycle model of
a vehicle in Figure 8.5.
8.5 (D) A block of mass m is attached to a cord of original length L and is rotating around
a thin hub, as shown in Figure 8.26. Friction is negligible. Find the constraint force, which
is the tension in the cord, if a) the cord is not wrapping around the hub, and b) the cord is
wrapping around the hub.
446 Applied Dynamics

   

   

 
 
 

FIGURE 8.26
Figure for Problem 8.5. a) Cord is not wrapping around the hub, b) cord is wrapping around
the hub.

Section 8.3—Velocity Representation


8.6 (E) Consider a 3-1-3 Euler angle transformation from frame XY Z to xyz via the angles
φ, θ, ψ. Write the angular velocity in terms of the generalized coordinates (the Euler
angles). Then, consider the angular velocity components in the xyz frame, ωx , ωy , ωz
as the generalized speeds. Calculate the partial angular velocities associated with both
descriptions.
8.7 (E) Consider a 3-1-2 Euler angle transformation from frame XY Z to xyz via the angles
φ, θ, ψ. Write the angular velocity in terms of the generalized coordinates (the Euler
angles). Then, consider the angular velocity components in the xyz frame, ωx , ωy , ωz
as the generalized speeds. Calculate the partial angular velocities associated with both
descriptions.
8.8 (M) Consider the bicycle model of a ground vehicle in Figure 8.5. Now, treat the steer
angle δ as an input and not as a motion variable. This reduces the number of position
variables to three, XA , YA , and θ. Because there are two constraints, the vehicle has one
d.o.f. Using vA as the generalized speed, obtain the partial velocities of the center of mass
motion and the angular velocity.
8.9 (D) Figure 3.32 shows the top view of a trailer. Both the front and rear vehicles are
modeled by assuming that all wheels roll without slip or slide. Now, treat the steer angle δ
as an input (and not as a motion variable), so that the trailer now has one d.o.f. Using vA
as the generalized speed, write the partial velocities of the centers of mass GF and GR of
the front and rear vehicles, as well as the partial angular velocities.

Section 8.4—Virtual Displacements and Virtual Work


8.10 (E) Find the virtual displacement of the slider in the slider-crank mechanism in Figure
8.27 using i) relative velocity relationships, and ii) the analytical expressions. Then, find
the virtual work of the force F .
8.11 (E) Calculate the virtual displacement of point P in Figure 8.28a.
8.12 (M) Calculate the virtual work associated with the force F2 acting through point P in
Figure 8.28a. Also calculate the virtual work due to the gravitational forces acting on the
two rods.
8.13 (M) Calculate the virtual displacement of point P in Figure 8.28b. The mass is sus-
pended from an arm that is attached to a rotating column. The pendulum swings in the
plane generated by the column and arm.
Analytical Mechanics 447

B
g
m1, L2 m2, L2

O 
P F

FIGURE 8.27
Figure for Problem 8.10.

a) b)
y z
x M A F1 x L1 g
x
m 1, L 1 . L2

1
m 2, L 2 P m
B F2 o
2 30
P

FIGURE 8.28
Figures for a) Problems 8.11, 8.12, 8.18, 8.43, 8.58, 8.69, and 9.37, b) Problem 8.13.

Section 8.5—Virtual Displacements and Virtual Work for Rigid Bodies


8.14 (M) Calculate the virtual displacement of point A in Figure 8.29a, which is at the tip
of a “scissor platform.” All rods inside the mechanism are of length L.
8.15 (M) Calculate the virtual work associated with the forces F1 and F2 in Figure 8.29b
by i) treating each force individually, and ii) calculating the resultant moment and using
the counterpart of Equation (8.63) for rotational motion. Show that the two results are
equivalent.
8.16 (M) A cart of mass M moves without friction, as shown in Figure 8.30. Inside the cart
there is a slot and a mass m slides inside the slot with friction. Draw the free-body diagrams
of the two masses and denote the normal force exerted by the ground to the cart by N1 ,
the normal force the mass m exerts on the cart by N2 , and the friction force between the
mass and cart by Ff . Calculate the virtual work of the entire system. Show that the normal
forces N1 and N2 do not contribute to the virtual work, while the friction force Ff does.

Section 8.6—Generalized Forces


8.17 (M) Calculate the generalized force associated with the pendulum in Figure 8.29b.
8.18 (M) Calculate the generalized forces associated with the forces and gravity that act in
Figure 8.28a.
8.19 (M) Consider the slider-crank mechanism in Figure 8.11. Now, there is friction between
the slider and the surface, characterized by the friction coefficient µ. Calculate the contri-
448 Applied Dynamics

30o g
a) b) F2
A x
y L/3
B
x
Y L z
L
All inside
O X rods of length L F1
P

FIGURE 8.29
Figures for a) Problem 8.14, b) Problems 8.15 and 8.17.










FIGURE 8.30
Figure for Problems 8.16 and 8.20.

bution of the friction force to the generalized force. How do we characterize the direction
of the friction force using variational notation?
8.20 (M) Calculate the generalized forces associated with all the forces that act in Figure
8.30.
8.21 (M) Calculate the potential energy and the generalized forces of the two pendulums
connected by a spring in Figure 7.12.

Section 8.7—Principle of Virtual Work for Static Equilibrium


8.22 (E) Find the equilibrium positions for the system in Figure 5.56a by means of analytical
techniques.
8.23 (E) Find the equilibrium positions for the system in Figure 5.56b by means of analytical
techniques.
8.24 (M) Each rod in Figure 8.31 weighs 40 lb and is of length 2 ft. The springs each have
constant of k = 60 lb/in. The springs are unstretched when the rods are vertical. Someone
pushes the rods down with force F . Find the magnitude of F so that the rods acquire an
angle of θ = 15◦ .
8.25 (M) Calculate the value of θ in Figure 8.32a at equilibrium. The spring is unstretched
when θ = 60◦ .
Analytical Mechanics 449

B
g

k A C k
 m, L

FIGURE 8.31
Figure for Problem 8.24.

a) b) g
g k
m

m, L m, L 2m =0
x1
k k
m
x2

FIGURE 8.32
Figures for a) Problem 8.25, b) Problem 8.26.

8.26 (M) Find the equilibrium positions of the pulley system in Figure 8.32b for the value
mg/k = 3.
8.27 (D) A rod of mass M and length L is pivoted at point O, as shown in Figure 8.33.
At the end of each rod is a concentrated mass m. The concentrated mass on the right is
connected via a spring to another point mass 2m. Using the angle θ and displacement x of
mass 2m (x = 0 when springs are not deflected) as generalized coordinates, calculate the
potential energy. Then, find the static equilibrium positions when M = 3m.

Section 8.8—D’Alembert’s Principle


8.28 (M) Using D’Alembert’s principle, obtain the equation of motion of the bead in Figure
8.34a sliding without friction in the parabolic tube described by z = x2 /4, while the tube
is rotating about the z axis with constant angular velocity Ω.
8.29 (D) Find the equation of motion of the system in Figure 8.34b using direct application
of D’Alembert’s principle.

Section 8.9—Hamilton’s Principle


8.30 (M) Obtain the equation of motion of the bead in Figure 8.34a using Hamilton’s
principle. The bead slides without friction in the parabolic tube described by z = x2 /4,
while the tube is rotating about the z axis with constant angular velocity Ω.
8.31 (M) Using Hamilton’s principle, obtain the equations of motion of the system in Figure
450 Applied Dynamics

B
g 2L/3 m
M
k k

L/3 O x 2m
A
m
k

FIGURE 8.33
Figure for Problems 8.27, 8.38, 8.60, and 4.5.

a) z x2 b)
z=
g 4 g y

m 2m
k
x
1 2

FIGURE 8.34
Figures for a) Problems 8.28, 8.30, and 8.33, b) Problem 8.29.

8.35, which consists of a disk of mass M and radius R, to which a pendulum of mass m and
length L is attached by a pin joint. Friction is sufficient to prevent slip of the disk while it
rolls. The spring is unstretched when φ = 0 and the force F is always horizontal.
8.32 (M) Obtain the equations of motion of the cart to which a pendulum is attached in
Figure 7.1 using Hamilton’s principle.

Section 8.10—Lagrange’s Equations


8.33 (M) Using Lagrange’s equations, obtain the equation of motion of the bead in Figure
8.34a sliding without friction in the parabolic tube described by z = x2 /4, while the tube
is rotating about the z axis with constant angular velocity Ω.
8.34 (M) Obtain the equations of motion of the two pendulums connected by a spring in
Figure 7.12.
8.35 (M) Obtain the equations of motion of the system in Figure 8.35, which consists of a
disk of mass M and radius R, to which a pendulum of mass m and length L is attached by
a pin joint at D, using Lagrange’s equations. The configuration shown is for φ = 0. Friction
is sufficient to prevent slip of the disk while it rolls. The spring is unstretched when φ = 0
and the force F is always horizontal.
8.36 (D) The platform in Figure 8.36a is rotating with constant angular velocity Ω. Attached
to the platform are two identical pendulums of length L and mass m. The pendulums swing
in the xz plane, and they are connected by a spring of constant k. Derive the e.o.m. using
Analytical Mechanics 451



 
 



 



 

FIGURE 8.35
Figure for Problems 8.31, 8.35, and 8.46.

Lagrange’s equations. Assume that the spring deforms only in the x direction and that it
is unstretched when θ1 = θ2 = 0.

a) b) y
g
O g x
a a x
1 m1, L1 F
L 
z
P
1 2 L m2, L2
m k m B
2

FIGURE 8.36
Figures for a) Problems 8.36, 8.56, and 8.57, b) Problems 8.37 and 8.66.

8.37 (M) Using Lagrange’s equations, obtain the equations of motion of the double pendulum
in Figure 8.36b.
8.38 (M) Obtain the equations of motion of the rod of mass M and length L is pivoted at
point O, shown in Figure 8.33.
8.39 (M) Obtain the equations of motion of the pendulum in Example 8.3.
8.40 (M) The pendulum in Figure 8.37 swings on the inclined plane xy. The incline angle
γ of the plane (between X and x axes) is being raised at the constant rate of γ̇. The pivot
point of the pendulum, point O, is at a distance of 1.2L from the y axis. Find the equation
of motion of the pendulum.
8.41 (M) Consider the bead on a circular hoop in Figure 8.16. Now, the rotation rate of the
hoop is no longer a known quantity and is denoted by the variable φ̇. The mass of the hoop
is M . Obtain the equations of motion.
8.42 (M) Two masses m1 and m2 are attached to springs k1 and k2 and they move without
friction in a slot that is attached to a rotating table with mass moment of inertia IG , as
shown in Figure 8.38. The springs are unstretched when the displacements of the masses
are a. Obtain the equations of motion using Lagrange’s equations.
452 Applied Dynamics
x

g O
X L

1.2L 
m
Z
Y, y

FIGURE 8.37
Figure for Problem 8.40.

y x Top view
Y
x1 y
IG m1
x2 k1 x
a
k2
. X
m2

FIGURE 8.38
Figure for Problems 8.42, 8.47, 8.55, and 8.64.

8.43 (D) Obtain the equations of motion of the double pendulum connected to a cart in
Figure 8.28a.

Section 8.11—Constrained Systems


8.44 (M) Find the equation of motion of the double link in Figure 8.24 that is constrained
to move on an incline by using the constraint relaxation method. The mass of rod OB is m
and the mass of rod BP is 2m.
8.45 (M) The bead of mass m in Figure 8.6 moves in the spiral of radius r. The coefficient
of friction between the bead and spiral is µ. Obtain the equation of motion of the bead via
a constraint formulation and by using θ as the generalized coordinate.

Section 8.12—Kane’s Equations


8.46 (M) Obtain the equations of motion of the system in Figure 8.35, which consists of a
disk of mass M and radius R, to which a pendulum of mass m and length L is attached by
a pin joint, using Kane’s equations. Friction is sufficient to prevent slip of the disk while it
rolls. The spring is unstretched when φ = 0 and the force F is always horizontal.
8.47 (M) Two masses m1 and m2 are attached to springs k1 and k2 and they move without
friction in a slot that is attached to a rotating table with mass moment of inertia IG , as
shown in Figure 8.38. The springs are unstretched when the displacements of the masses
are a. Obtain the equations of motion using Kane’s equations.
8.48 (M) The wheelbarrow of mass m and mass moment of inertia IG in Figure 8.39 operates
Analytical Mechanics 453
y

L
E x
C
G y Y
FC 
t d
 X x
D  X
FD

FIGURE 8.39
Figure for Problem 8.48.

on the XY plane. Find the nonholonomic constraint for the motion of the wheelbarrow
shown in terms of X, Y , and θ, given that the wheel of the wheelbarrow rolls without
slipping. Then, using Kane’s equations and vE and θ̇ as the generalized speeds, obtain the
equations of motion of the wheelbarrow.

g
m, L
z y
 G
x

A
 vA

FIGURE 8.40
Figure for Problem 8.49.

8.49 (M) The rod of mass m and length L in Figure 8.40 is released from rest making an
angle of θ with the horizontal. There is no friction (µ = 0) between point A and the surface.
Obtain the equations of motion of the rod by means of Kane’s method, using the velocity
of point A, vA , and θ̇ as generalized speeds.
8.50 (M) Consider the bicycle model of a ground vehicle in Figure 8.5. Now, treat the
steer angle δ as an input and not as a motion variable. This reduces the number of position
variables to three, XA , YA , and θ. Because there are two constraints, the vehicle has one d.o.f.
Using vA as the generalized speed, obtain the equation of motion using Kane’s equations.
Show that this process is equivalent to summing moments about the instant center of the
vehicle. Hint: Do not forget to express the angular velocity θ̇ in terms of vA .
8.51 (D) Figure 3.32 shows the top view of a trailer. Both the front and rear vehicles are
modeled by assuming that all wheels roll without slip or slide. Now, treat the steer angle
δ as an input (and not as a motion variable), so that the trailer now has one d.o.f. Using
vA as the generalized speed, write the equation of motion. The masses and centroidal mass
moments of inertia are mF , mR , IF , and IR .
8.52 (D) The cart of mass m and mass moment of inertia IG in Figure 8.41, which is similar
454 Applied Dynamics

y L

C x
k 
FC X Y
G B y
s 
M
d k
t A e
x
D  
X X
FD FA

FIGURE 8.41
Figure for Problem 8.52.

to Figure 8.7, has a block of mass M which is connected to two springs and moves without
friction in the y direction. Find the equations of motion using Kane’s equations.
8.53 (M) The boat in Figure 8.42a of mass m and centroidal moment of inertia IG is
propelled by an outboard motor, which provides thrust F acting through the variable angle
γ. The hydrodynamic properties of the boat restrict the velocity of point D on the boat to
be along the x axis, so that vD = vD i. Obtain the equations of motion of the boat using
Kane’s equations.

Section 8.13—Natural and Nonnatural Systems, Equilibrium

Y
a) x b) Z
g
y s
 c
m
G k
D b m, IG
Y
a 
A X 2
1
X

F

FIGURE 8.42
Figures for a) Problem 8.53 and b) Problem 8.54.

8.54 (M) A ball of mass m is inside a frictionless tube, and it is attached to the ends of the
tube by a spring of constant k and a damper of constant c, as shown in Figure 8.42b. When
the spring is unstretched, the mass is at a distance L from end O of the tube. The end O of
the tube is attached to a joint, which makes the tube behave like a joystick. Given that the
orientation of the joystick is described by the angles θ1 and θ2 , and that the time histories
of these angles are known, derive the equation of motion of the ball.
8.55 (M) For the two masses moving freely in a rotating slot in Figure 8.38, calculate the
Analytical Mechanics 455

energy integral (Jacobi integral) when a) the rotation rate of the table is a variable and b)
the rotation rate of the table is constant, θ̇ = Ω = const.
8.56 (M) Consider the pendulum pair in Problem 8.36 and obtain the equilibrium equations.
Then, linearize the equilibrium equations about θ1 = θ2 = 0 and calculate the equilibrium
positions.
8.57 (C) Consider the pendulum pair in Problem 8.36 and obtain the energy integral. Given
the values of L = 2 ft, m = 1/50 slug, a = 0.6 ft, k = 12 lb/ft, Plot Jacobi’s integral as a
function of the rotation speed Ω in the range 0 ≤ Ω ≤ 3 rad/sec.

Section 8.14—Small Motions about Equilibrium


8.58 (M) Write the kinetic energy for the two rods attached to the cart in Figure 8.28a and
obtain the mass matrix for when the system is linearized about x = 0, θ1 = θ2 = 0.
8.59 (M) Linearize the equation of motion in Examples 8.4 and 8.6 and assess the stability
of the equilibrium position. Begin your analysis with energy expressions.
8.60 (M) Linearize the equations of motion of the system in Figure 8.33 about the position
x = 0, θ = 0.

Section 8.15—Rayleigh’s Dissipation Function


8.61 (M) Write Rayleigh’s dissipation function for the mass-spring-damper system in Figure
7.23 and from that express the mass, stiffness, and damping matrices without writing the
equations of motion.

x
O! L#" O"
g z

m, L m, L

"
!
c
F
B C

FIGURE 8.43
Figure for Problem 8.62.

8.62 (M) Write Rayleigh’s dissipation function for the two pendulums connected by a
damper in Figure 8.43. Then, linearize the Rayleigh’s dissipation function and express it in
matrix form.

Section 8.16—Generalized Momentum, First Integrals


8.63 (M) Consider the bead on a circular hoop in Figure 8.16. Now, the rotation rate of the
hoop is no longer a known quantity and it is denoted by the variable φ̇. The mass of the
hoop is M . Determine the motion integrals.
8.64 (M) For the two masses moving freely in a rotating slot in Figure 8.38, calculate the
generalized momenta and integrals of the motion when the rotation rate of the table is
constant, θ̇ = Ω = const.
456 Applied Dynamics

Section 8.17—Impulsive Motion


8.65 (E) Consider Figure 8.23. Now, the impulsive force F̂ is applied to the collar. Calculate
the ensuing velocity ẋ and angular velocity θ̇.
8.66 (M) The double pendulum in Figure 8.36b is at rest when an impulsive force F̂ is
applied to it at an angle of φ = 30◦ . Calculate the angular velocities immediately after the
impulse.
8.67 (M) Solve Problem 5.15 using a Lagrangian approach, by using the angular velocities
of the two rods (call them θ̇1 and θ̇2 ) as generalized velocities.
8.68 (M) Figure 8.44 is a simple schematic of what happens to a driver’s head when the
vehicle is hit from the rear. Consider a driver of mass m1 + m2 = 80 kg, where the head
weighs m1 = 5 kg and is assumed to be a disk of radius R = 11 cm hinged to the body.
The vehicle has a mass of M = 1500 kg and the driver is wearing her seatbelt. The vehicle
is traveling at a speed of 45 km/hr and the collision can be modeled by an impulsive force
(F is replaced by F̂ in the figure) of magnitude 1000 N·s. Find the angular velocity of the
driver’s head immediately after the collision.





 

FIGURE 8.44
Figure for Problem 8.68.

8.69 (C) The two rods attached to the cart in Figure 8.28a is at rest with x = 0, θ1 = θ2 = 0
when an impulsive force F̂1 = 2 N·s is applied to the cart (F̂2 = 0). Calculate the ensuing
velocities immediately after impact. Use values of M = 5 kg, m1 = m2 = 2 kg, L1 = 0.8 m,
L2 = 1.2 m.
Analytical Mechanics 457


9
Three-Dimensional Kinematics of Rigid Bodies

9.1 Introduction
This chapter is concerned with the geometry and kinematics of rigid bodies that undergo
three-dimensional motion. In Chapter 2 we studied basic kinematics, which included coordi-
nate transformations, angular velocity, and relative motion. The kinematics of rigid bodies
was restricted to plane motion. So were the inertia properties of rigid bodies in Chapter
4. This chapter extends the analysis of the kinematics of rigid bodies to three-dimensional
motion.
The chapter begins with a discussion of the kinematics of three-dimensional motion.
Angular velocity is quantified by means of Euler angles, Euler parameters, and Rodrigues
parameters. Euler angle sequences and reference frames are discussed. The all-important
cases of axisymmetric bodies and of rolling are considered. The chapter continues with the
geometric properties of a rigid body and extends the concept of mass moment of inertia to
three dimensions by means of the inertia matrix. Chapter 11 brings together the concepts
in Chapters 2, 3, 4, 5, this chapter and Chapter 10 in order to study the kinetics of three-
dimensional rigid body motion.

9.2 Basic Kinematics of Rigid Bodies


The general motion of a rigid body can broadly be classified into three categories:

• pure translation,
• pure rotation,
• combined translation and rotation.

Consider Figure 9.1, which describes a rigid body and its center of mass G; two points
on the body, B and P ; and a fixed reference point O from which the motion of the body is
observed. A reference frame xyz is attached to the body. The angular velocity and angular
acceleration of the reference frame coincide with the angular velocity and acceleration of
the body, and the axes of the reference frame are referred to as body-fixed axes or body
axes.
Let us specify that point B does not move with respect to the body and point P may
move with respect to the body. The relative velocity and relative acceleration of P with
respect to the body are referred to as v(P/B)rel and a(P/B)rel , respectively. When P is fixed
on the body, then v(P/B)rel = 0 and a(P/B)rel = 0. It follows from the developments of
Chapter 2 that we can write the relative position, velocity, and acceleration relationships

459
460 Applied Dynamics
ȦĮ
$
(
!$)"

!$ "##### %
'
!"
&
!

FIGURE 9.1
A rigid body.

as

rP = rB + rP/B vP = vB + ω × rP/B + v(P/B)rel

aP = aB + α × rP/B + ω × (ω
ω × rP/B ) + 2ω
ω × v(P/B)rel + a(P/B)rel (9.1)

In the majority of dynamics problems involving rigid body motion, we attach the refer-
ence frame to the body. Under certain circumstances it is preferable to employ a reference
frame that is not attached to the body. Rotation of axisymmetric bodies is the most com-
mon use of such a frame. When using a reference frame not attached to the body, we need
to distinguish between the angular velocity of the body and the angular velocity of the
reference frame.

9.2.1 Pure Translation


In this case, the rigid body moves with no angular velocity and no angular acceleration,
that is, ω = 0, α = 0. Every point on the body has the same translational velocity and
acceleration, and the kinematics can be analyzed similarly to particle motion. Three trans-
lational coordinates are sufficient to describe the motion. A body may move in a curved
trajectory without any rigid body rotation. An example of this is the approach and landing
of an airplane.1

9.2.2 Pure Rotation


In the case of pure rotation, the motion of the rigid body is described using rotational
parameters alone. The velocity and acceleration of any point on the body can be expressed
in terms of the angular velocity, angular acceleration, and the distance of the point from
the rotation center. This type of motion is separated into two categories: i) rotation about
a fixed axis, and ii) rotation about a fixed point. Rotation about a fixed axis is considered
a special case of rotation about a fixed point. For plane motion the two categories coincide.
Three-Dimensional Kinematics of Rigid Bodies 461

a) b)
t
P vP, a
h P
b
rP an
b
B
h
O
h
n

FIGURE 9.2
Rotation about a fixed axis. a) General view, b) view with axis h out of the plane.

Rotation about a Fixed Axis


Consider a body rotating about a fixed axis h, as shown in Figure 9.2a. Figure 9.2b shows
the normal and tangential coordinates when looking into the axis of rotation. The axes t, n,
and h form a right-handed coordinate system, with et ×en = eh , where eh is the unit vector
along the fixed axis of rotation. The angular velocity and angular acceleration of the body
are

ω = ωeh ω = ω̇eh
α = ω̇ (9.2)

Next, consider a point P on the body and express its position in terms of its components
along the h axis and the plane perpendicular to the h axis. The normal and tangential
coordinates are defined on this plane. The position vector is

rP = h + b (9.3)

in which h = heh and b = ben , h is the distance between O and B and b is the distance
between B and P . When P is fixed on the body, its velocity is

vP = ω × rP = ωeh × (heh − ben ) = bωet (9.4)

The magnitude of the velocity is bω. The velocity of any point on the body is dependent
only on the perpendicular distance between that point and the axis of rotation. In terms of
normal-tangential coordinates, the acceleration can be expressed as

aP = at + an = α × rP + ω × (ω
ω × rP ) (9.5)

The tangential acceleration is at = at et = bαet , and the normal acceleration is an = an en =


bω 2 en . A body that rotates about a fixed axis has one degree of freedom.

Rotation about a Fixed Point


In the case of rotation about a fixed point, the angular velocity vector does not lie on a
fixed axis. The rate of change of the angular velocity depends on the change in direction of
1 A few seconds before touchdown, the pilot pitches the plane up to cause the rear wheels to make contact

with the runway first and then pitches down to cause the front wheels to touch down.
462 Applied Dynamics

the angular velocity, as well as on the change in magnitude. A body rotating about a fixed
point has three degrees of freedom and the angular velocity is usually the combination of
two or more rotation components.
Consider the collar in Figure 9.3a. The base is rotating about the fixed Z axis with
angular velocity ω1 , so that the XY Z axes are attached to the rotating base. The arm,
around which the collar moves, rotates on the Y Z plane by pivoting around O and makes
an angle of β with the Y axis. Denote the arm axis by y, so that the xyz axes are obtained
by rotating the XY Z coordinates about X. From Figure 9.3b, the unit vectors are related
by

i = I j = cos βJ + sin βK k = cos βK − sin βJ (9.6)

a) b)
Z y
z
Z
z
O
3
y
y
Y Y

X 1

FIGURE 9.3
Rotation about a fixed point. a) General view, b) coordinate system.

The collar rotates about the rod (y axis) with an angular velocity of ω3 . The angular
velocity of the collar is

ω = ω base + ω rod/base + ω collar/rod = ω 1 + ω 2 + ω 3 (9.7)

in which

ω 1 = ω1 K = ω1 sin βj + ω1 cos βk ω 2 = β̇I = β̇i ω 3 = ω3 j (9.8)

are the components of the angular velocity, so that the angular velocity vector becomes

ω = β̇i + (ω1 sin β + ω3 ) j + ω1 cos βk (9.9)

The collar can be viewed as rotating about the fixed point O, as all components of the
angular velocity are about axes that go through point O. Note that the coordinate system
used to represent the angular velocity is not attached to the collar. The angular velocity of
the reference frame is

ω f = ω 1 + ω 2 = β̇i + ω1 sin βj + ω1 cos βk (9.10)

Let us next calculate the angular acceleration of the collar. As discussed in Chapter 2,
we can find the angular acceleration several ways, as outlined below.
Three-Dimensional Kinematics of Rigid Bodies 463

• Using the transport theorem for the entire angular velocity. The angular ac-
celeration becomes

α = ω̇ ω rel + ω f × ω
ω = ω̇ (9.11)

in which
   
ω rel = β̈i + ω̇3 + ω̇1 sin β + ω1 β̇ cos β j + ω̇1 cos β − ω1 β̇ sin β k
ω̇ (9.12)

and ω f is given in Equation (9.10). The term ω f × ω becomes

ω f × ω = ω f × (ω
ωf + ω 3) = ω f × ω 3

 
= β̇i + ω1 sin βj + ω1 cos βk × ω3 j = −ω1 ω3 cos βi + ω3 β̇k (9.13)

so that the angular acceleration becomes


   
α = β̈ − ω1 ω3 cos β i + ω̇3 + ω̇1 sin β + ω1 β̇ cos β j

 
+ ω̇1 cos β − ω1 β̇ sin β + ω3 β̇ k (9.14)

• Using the transport theorem on each term of the angular velocity individu-
ally. Using the description of the angular velocity Equation (9.7), the angular acceler-
ation is written as

ω 1 + ω̇
α = ω̇ ω 2 + ω̇
ω3 (9.15)

The individual derivative terms are calculated next. For the base, because the direction
of the angular velocity ω 1 is fixed, we can use direct differentiation so that

ω 1 = ω̇1 K = ω̇1 (sin βj + cos βk)


ω̇ (9.16)

For the rod with respect to the base, the transport theorem gives

ω̇ ω 2rel + ω 1 × ω 2
ω 2 = ω̇ (9.17)

It follows that using the transport theorem on each component of the angular velocity
leads to the angular acceleration in the form

ω 1 + ω̇
α = ω̇ ω 3rel + ω 1 × (ω
ω 2rel + ω̇ ω2 + ω 3) + ω 2 × ω 3 (9.18)

• Direct differentiation of each term in the angular velocity. Direct differentiation


ω 2 can be carried out using the XY Z or xyz frames, leading to
to find ω̇

ω 2 = β̈1 I + β̇ İ = β̈1 i + β̇ i̇
ω̇ (9.19)

where
 
İ = i̇ = ω f × i = β̇i + ω1 (sin βj + cos βk) × i = ω1 (cos βj − sin βk) (9.20)
464 Applied Dynamics

The derivative of ω 2 becomes

ω 2 = β̈1 i + ω1 β̇ cos βj − ω1 β̇ sin βk


ω̇ (9.21)

For the collar with respect to the rod, using the transport theorem, the angular accel-
eration has the form

ω 3 = ω̇
ω̇ ω1 + ω 2) × ω 3
ω 3rel + (ω (9.22)

where

ω 3rel = ω̇3 j
ω̇ (9.23)

ω 3 results in
Using direct differentiation to find ω̇

ω 3 = ω̇3 j + ω3 j̇
ω̇ (9.24)

where
 
j̇ = ω f × j = β̇i + ω1 sin βj + ω1 cos βk × j = −ω1 cos βi + β̇k (9.25)

so that the derivative of ω 3 becomes

ω 3 = −ω1 ω3 cos βi + ω̇3 j + ω3 β̇k


ω̇ (9.26)

Adding the individual angular acceleration components leads to Equation (9.14).


• Differentiating Equation (9.9) directly. To this end, we can use the xyz coordinates
or the XY Z coordinates. The angular velocity of the XY Z frame is ω 1 .
An observation from Equation (9.18) is that even if the individual components of the
angular velocity are constant in magnitude, the angular acceleration is not zero, because
ω1 =
the direction of the angular velocity vector is changing. Indeed, setting the quantities ω̇
ω 2rel = ω̇
ω̇ ω 3rel = 0, we are left with

α = ω 1 × (ω
ω2 + ω 3) + ω 2 × ω 3 (9.27)

which represents the change in direction of the angular velocity. These terms are known
as gyroscopic effects. Even when β is constant (ω ω 2 = 0, in addition to ω1 and ω3 being
constant), the expression for angular acceleration becomes

α = ω 1 × ω 3 = −ω1 ω3 cos βi (9.28)

This term is perpendicular to the yz plane and describes the change in direction of the
angular velocity. It is depicted in Figure 9.4.
The line specifying the direction of the angular velocity vector ω is known as the instan-
taneous axis of acceleration, or the instantaneous axis of rotation. The unit vector along
this axis is defined as
ω
n = (9.29)
ω
in which ω = |ω
ω | is the magnitude of the angular velocity. The rigid body can be viewed as
rotating about the axis defined by the angular velocity vector at any instant.
The trajectory of the instantaneous axis of rotation defines the body and space cones.
When the trajectory of the angular velocity vector is viewed from the rotating body, the
Three-Dimensional Kinematics of Rigid Bodies 465

Z
z

 1x  3
1

3 y


Y

X, x

FIGURE 9.4
Gyroscopic effect when β is constant.

cone that is generated by the instantaneous axis is the body cone. When the trajectory is
viewed from an inertial frame, the space cone is generated. The body and space cones are
always in contact with each other. The line of contact, which is the angular velocity vector
ω , is also referred to as the generatrix. Figure 9.5 shows the body and space cones for an
arbitrary body.2 Body and space cones will be discussed in Chapter 11 when describing
torque-free motion.

!"#$%&
$'(%

)'*+& Ȧ
$'(%

FIGURE 9.5
Body and space cones.

9.2.3 Combined Translation and Rotation


A body undergoing combined translation and rotation requires both translational and
angular parameters to describe its motion. An unrestricted rigid body undergoing three-
dimensional motion has six degrees of freedom. Combined translation and rotation of an
arbitrary rigid body is too general to be described in broad terms.
2 As the cones drawn in Figure 9.5 are for arbitrary motion, the cones are not depicted as completely

closed.
466 Applied Dynamics

It turns out that we can extend the results of Euler’s theorem into general translation
and rotation. This extension is known as Chasles’ theorem, which states:

The most general displacement of a rigid body is equivalent to a translation of some


point on the body, plus a rotation about an axis that goes through that point.

While for rotation about a fixed point the axis of rotation is unique (it goes through the
fixed point), for general motion the translation point and rotation axis are not unique. We
can illustrate this by a simple example from plane motion. Consider the rectangle in Figure
9.6, which is shown in its initial position ABCD and is moved to position A0 B 0 C 0 D0 .
There are several ways to move the rectangle. One is to first have a translation from
any point on it to its displaced location, say, from point A to A0 , and then a rotation by θ
about the displaced point, here A0 . Another is to find the pole of the displacement, denoted
by P in Figure 9.6, and to rotate about P by θ. To find the pole of the displacement, we
draw extensions to lines AD and A0 D0 (or any two lines), mark their intersection by E, and
locate the circle that goes through D, D0 , and E. The point at which the perpendicular
bisector of the line DD0 intersects the circle is the pole of the displacement.

P
E


D C

Aʹ θ
rAʹ/A

A B

FIGURE 9.6
Combined translation and rotation on a plane.

For general three-dimensional motion, it is customary to select the rotation axis and
point on the rotation axis using screw geometry, which will be discussed later on in this
chapter. In screw geometry, the motion is represented by a rotation about an axis plus a
translation along the same axis. The rotation axis is called the screw axis.

Example 9.1
Consider the disk in Figure 9.7 spinning with angular velocity ω about point G. The pivot
arm of length L is attached to a column that rotates with angular velocity Ω. Calculate the
angular velocity and angular acceleration of the disk. Does this problem describe rotation
about a fixed point? If so, which point?
The xyz coordinates move with the rotating shaft. The angular velocity is

ω = ω shaft + ω disk/shaft = Ωk + ωi [a]


Three-Dimensional Kinematics of Rigid Bodies 467

Z, z

O G y
L

FIGURE 9.7
Spinning disk on a pivot.

The xyz coordinates are rotating with ω shaft = ω f = Ωk, so that


d d
i = Ωk × i = Ωj k = Ωk × k = 0 [b]
dt dt
The angular acceleration is found by differentiating Equation [a], which yields
ω = Ω̇k + Ωk̇ + ω̇i + ω i̇
α = ω̇ [c]

Introducing the derivatives of the unit vectors in Equation [b] into Equation [c], the angular
acceleration becomes
α = Ω̇k + ω̇i + Ωωj [d]
The rotation of the disk cannot be considered as rotation about a fixed point. The
rotation axes of the shaft (Z) and the rotation axis of the disk (x) do not intersect. Also,
although point O is fixed, it is not the center of rotation. The rotation axis of the disk with
respect to the shaft does not go through point O.

9.3 Euler Angles


We demonstrated in Chapter 2 that three successive rotations about nonparallel axes are
sufficient to define the most general transformation of one coordinate system into another.
The transformation can be expressed in many ways and is not unique. In Chapter 2, we nar-
rowed our interest to body-fixed and space-fixed rotation sequences. This section quantifies
the different choices for carrying out the rotation transformations.

9.3.1 Euler Angle Sequences


In a body-fixed rotation sequence, we begin with an initial reference frame attached to the
body, say, A, with coordinate axes a1 a2 a3 and unit vectors a1 , a2 , and a3 . This reference
frame is rotated about one of its axes. There are three choices, one for each coordinate axis.
The resulting coordinate frame is called A0 with its axes a01 a02 a03 . For example, if we rotate
about the a3 axis by angle ψ, in the resulting a01 a02 a03 frame the a3 and a03 axes coincide, as
shown in Figure 9.8a.
468 Applied Dynamics

a) b) c)
a3, a3
a3 a3 a3 , b3

a1
a2 a1 b2
a1
a1
a2 a2, a2 b1 a2
a1

FIGURE 9.8
3-2-3 rotation sequence: a) first rotation about a3 by ψ, b) second rotation about a02 by θ,
c) third rotation about a003 by φ.

The second transformation rotates the A0 frame about one of its axes. At this stage,
there are only two choices, as the axis about which the previous rotation took place needs
to be excluded. Otherwise, we cannot distinguish between the first and second rotations.
Continuing with the example of having the first rotation about the a3 axis, the second
rotation cannot take place about a03 = a3 , leaving rotation about a01 and a02 as the choices.
Let us select the second rotation to take place about a02 by angle θ and denote the resulting
frame by A00 and corresponding axes by a001 a002 a003 , as shown in Figure 9.8b.
The third and final rotation takes place about one of the axes of A00 frame, keeping
in mind the need to exclude the axis about which the previous rotation took place. The
previous rotation was about a02 , with a002 = a02 , so the third rotation needs to take place
about a001 or a003 . Select the third rotation axis as a003 by φ and denote the rotated frame by
B and the rotated axes by b1 b2 b3 in Figure 9.8c. This transformation sequence is called a
3-2-3 sequence. The plane defined by a1 a2 (and also a01 a02 ) intersects the plane defined by
b1 b2 (and a001 a002 ) along the axis a02 = a002 . Shown in Figure 9.9, this axis is known as the line
of nodes, a term first coined in orbital mechanics.
It follows that there are 3×2×2 = 12 possible ways in which we can perform a body-fixed
rotation sequence. These combinations are

1-2-1, 1-2-3, 1-3-1, 1-3-2, 2-1-2, 2-1-3,


2-3-1, 2-3-2, 3-1-2, 3-1-3, 3-2-1, 3-2-3

These twelve sets are called Euler angle sequences. They can be classified into two categories,
each showing similar characteristics. The first category is when the first and third indices
are the same (e.g., 1-2-1, 2-3-2) and the second category consists of rotations where the first
and third indices are different (e.g., 3-2-1, 2-3-1) We select the sequence that offers better
visualization and that reduces the number of singularities that are encountered, as will be
discussed shortly.
The most commonly used Euler angle sequences are 3-1-3, 3-2-3, and 3-2-1. Table 9.1
describes these sequences and their significance.
Let us continue with the 3-2-3 transformation sequence. The rotation angles are denoted
by ψ (precession), θ (nutation), and φ (spin). It is convenient to use matrix notation and
to successively apply the transformations. The transformations are

{a0 } = [R1 ] {a} {a00 } = [R2 ] {a0 } {b} = [R3 ] {a00 } (9.30)
Three-Dimensional Kinematics of Rigid Bodies 469
!‫"މމ‬
1 1 !#$#!‫މ‬
1 1

‫׋‬Ú
ȥÚ
ș
.*%&'/)
'%/(&0 "#$%&'(&)*%&('+,
'"!"-'#$%(#!!'‫!މމ‬-‫މމ‬

!!

ȥ
"-
!‫!މ‬ ș
ș șÚ

!‫!މމ‬ ȥ !‫މމ!މ‬
‫׋‬ - -

!-
"!
"#$%&'(&)*%&('+,'
!!!-'$%('!‫މ!'!މ‬-

FIGURE 9.9
Combined 3-2-3 rotation sequence.

where
     
cψ sψ 0 cθ 0 −sθ cφ sφ 0
[R1 ] =  −sψ cψ 0  [R2 ] =  0 1 0 [R3 ] =  −sφ cφ 0  (9.31)
0 0 1 sθ 0 cθ 0 0 1
T T T
are the rotation matrices. Also, {a} = [a1 a2 a3 ] , {a0 } = [a01 a02 a03 ] , {a00 } = [a001 a002 a003 ] ,
T
and {b} = [b1 b2 b3 ] . Note that the notation for sines and cosines as s and c, respectively.
Also, [R1 ] and [R3 ] are matrices associated with 3 rotations, and [R2 ] is a 2 rotation.
The three transformations are combined to obtain the rotation matrix for the entire
transformation, with the result
{b} = [R3 ] {a00 } = [R3 ] [R2 ] {a0 } = [R3 ] [R2 ] [R1 ] {a} = [R3-2-3 ] {a} = [R] {a} (9.32)

TABLE 9.1
Commonly used Euler angle sequences

Seq. Rotation Angle Names Significance and Applications


Angles
3-1-3 φ, θ, ψ Precession, nu- Celestial mechanics (historically earliest used),
tation, spin axisymmetric bodies
3-2-3 ψ, θ, φ Precession, nu- Axisymmetric bodies (a.k.a. NASA Standard
tation, spin Aerospace)
3-2-1 ψ, θ, φ Heading, atti- Bodies undergoing small rotations (a.k.a.
tude, bank NASA Standard Aeroplane), vehicle motion
470 Applied Dynamics

where the combined transformation matrix [R] is


 
cψcθcφ − sψsφ sψcθcφ + cψsφ −sθcφ
[R] = [R3-2-3 ] =  −cψcθsφ − sψcφ −sψcθsφ + cψcφ sθsφ  (9.33)
cψsθ sψsθ cθ

Recall from Chapter 2 that the rotation matrix is unitary, so its inverse is equal to its
transpose. It follows that

{a} = [R]−1 {b} = [R]T {b} (9.34)

We can use Equations (9.32) and (9.34) to relate any vector expressed in the B frame in
terms of its coordinates in the A frame and vice versa.

9.3.2 Angular Velocity


The next step is to analyze the angular velocity and angular acceleration of the rotated
frame. Because angular velocities are vectors, they obey the commutativity rule. The angular
velocity vector has a component due to each rotation
0
A0 00
A00
A
ωB = A
ωA + ωA + ω B = ψ̇a3 + θ̇a02 + φ̇a003 = ψ̇a03 + θ̇a002 + φ̇b3 (9.35)

It is convenient to express the angular velocity components in terms of the same reference
frame. There are several choices when selecting the reference frame to use:
• The initial frame. Using the notation in Equation (9.35), the initial frame is the A
frame. The advantage of using this frame, which usually is selected as an inertial frame,
is that the derivatives are easy to obtain because the coordinate axes are fixed. On the
other hand, such a frame gives little insight for dynamics problems. Also, the inertia
matrix of a body written in terms of such a frame is not constant, creating algebraic
difficulties. The initial (or inertial) frame comes in handy in mechanism analysis and
robotics, where we are interested in the final position of a body with respect to its initial
position.
• The final frame. Using the notation in Equation (9.35), this is the B frame and it is
attached to the body. The advantage of using this frame is that the angular velocity
and acceleration are in terms of the coordinates attached to the body. As will be shown
later, a reference frame attached to the body leads to an inertia matrix that is constant,
which simplifies the equations of motion.
A similar advantage of body-fixed coordinates is observed when measuring angular ve-
locities. The angular velocities of a rotating body are taken by on-board (or body-fixed)
sensors, so the angular velocity measurements are in terms of body axes. The body-fixed
frame is usually preferred in dynamics problems.

• An intermediate frame. In the notation here, this would be one of the A0 or A00
frames. We prefer to use such frames when it is easier to state the expressions for
velocities and angular velocities. A very important application is rotating axisymmetric
bodies.
Consider the final frame, the B frame, which is attached to the body (thus the name
body-fixed). The components of the angular velocity are obtained in terms of the body-fixed
frame by using the transformation between the A and B frames, that is, Equation (9.34).
For the term ψ̇a3 = ψ̇a03 in Equation (9.35), we need to express a3 in terms of the B frame.
Three-Dimensional Kinematics of Rigid Bodies 471

This can be accomplished by reading the third row of [R3-2-3 ]T (or the third column of
[R3-2-3 ]) in Equation (9.33), which gives

a3 = −sθcφb1 + sθsφb2 + cθb3 (9.36)

To express the term θ̇a02 = θ̇a002 in Equation (9.35) in terms of the B frame, it is necessary
to consider the relationship between the A0 and B frames. Noting that {b} = [R3 ][R2 ]{a0 },
we can accomplish this by reading the second row of [R3-2-3 ]T (or the second column of
[R3-2-3 ]), and setting ψ = 0 (as the A0 frame is obtained after the first rotation is performed).
The result is

a02 = sφb1 + cφb2 (9.37)

The term φ̇a003 = φ̇b3 in Equation (9.35) needs no manipulation, since a003 = b3 .
The angular velocity can be expressed in terms of the body-fixed frame by introducing
Equation (9.36) and Equation (9.37) into Equation (9.35), with the result
A
ω B = ψ̇ (−sθcφb1 + sθsφb2 + cθb3 ) + θ̇ (sφb1 + cφb2 ) + φ̇b3

     
= −ψ̇sθcφ + θ̇sφ b1 + ψ̇sθsφ + θ̇cφ b2 + ψ̇cθ + φ̇ b3 (9.38)

which can be written in matrix form as


A B 
ω1 
−sθcφ sφ 0
 
ψ̇
A B 
 ω2  =  sθsφ cφ 0   θ̇  (9.39)
A B
ω cθ 0 1 φ̇
3

in which A ωiB is the i-th component of Aω B expressed in terms of the body-fixed axes. The
above equation can also be written as {ω} = [B] {θ̇}, where the notation is obvious.
The following are observations concerning the matrix [B] that relates the angular veloc-
ities to the rates of the Euler angles:

• The matrix [B] is not orthogonal. The rotations ψ, θ, and φ are performed about the
a3 , a02 , and a003 axes, which do not form an orthogonal set, even though the rotation
angles ψ, θ, and φ are independent of each other.
• The sines and cosines of the first rotation angle ψ are absent from [B].
• The matrix [B] becomes singular when the second angle, θ, is equal to zero or to a
multiple of π. This can be explained by noting that when θ = 0, the rotation reduces
to a 3-3 sequence, and the third (ψ) rotation cannot be distinguished from the first (φ).
Because [B] can be singular at times, the rates of change of the Euler angles cannot al-
ways be obtained from the components of the angular velocities, which causes problems
when integrating the equations of motion or when taking measurements. Angular ve-
locity measurements are usually taken by means of sensors attached to the body. These
measurements need to be translated to the rates of change of the Euler angles.
To visualize the singularity, we can invert Equation (9.39), to yield
 cφ sφ
 A B 
− sθ 0 ω1
 
ψ̇ sθ
sφ cφ 0  A B 
 θ̇  =     ω2  (9.40)
cθcφ cθsφ
φ̇ sθ − sθ 1 A B
ω3
472 Applied Dynamics

Equation (9.40) contains the kinematic differential equations relating the angular ve-
locities to the Euler angles. Existence of the singularity at θ = 0 and at multiple values
of π is obvious. Moreover, the kinematic differential equations are nonlinear, reducing
their suitability for hand calculations.
• Each of the Euler angle sequences results in a [B] matrix that becomes singular at certain
values of the second rotation angle. One of the objectives when selecting an Euler angle
sequence is to avoid singularities as much as possible. For example, in sequences where
no index is repeated (3-2-1, 1-3-2, etc.), the singularity occurs when the second rotation
angle has the value ±π/2. For dynamics problems where rotation angles are small, as
in aircraft or ground vehicle dynamics, a sequence such as 3-2-1, where no two indices
repeat, is preferred. For axisymmetric bodies undergoing rotation, as in a spinning top
or satellite, a sequence such as 3-2-3 or 3-1-3, where the first and third indices are the
same, is preferred.
• The singularity associated with a particular Euler angle sequence can be overcome by
switching to another Euler angle sequence in the neighborhood of the singularity, but
this makes the analysis cumbersome. When singularity is frequently encountered, it has
become preferable to use Euler parameters or Rodrigues parameters as alternate sets of
rotation variables.

9.3.3 Angular Acceleration


The angular acceleration is obtained by differentiating the angular velocity. This differenti-
ation requires use of the transport theorem. Interesting results are obtained depending on
the reference frame that is used:
• An inertial (fixed) frame. Because the unit vectors associated with the fixed reference
frame do not change direction, their derivatives are zero. The angular acceleration is
obtained by direct differentiation of the angular velocity components.
• A body-fixed frame. The angular acceleration is again obtained by direct differentia-
tion of the angular velocity components. This is because the second term in the transport
theorem drops out of the formulation. Indeed, invoking the transport theorem yields
A B A d A d d
ωB B A
ωB A
ω B × Aω B = B A
ωB
  
α = = + (9.41)
dt dt dt
For example, given the angular velocity components along the body axes for the 3-2-3
transformation in Equation (9.38), components of the angular acceleration become
A B d  
α1 = −ψ̇sθcφ + θ̇sφ = −ψ̈sθcφ − ψ̇ θ̇cθcφ + ψ̇ φ̇sθsφ + θ̈sφ + θ̇φ̇cφ
dt

A B d  
α2 = ψ̇sθsφ + θ̇cφ = ψ̈sθsφ + ψ̇ θ̇cθsφ + ψ̇ φ̇sθcφ + θ̈cφ − θ̇φ̇sφ
dt

A B d  
α3 = ψ̇cθ + φ̇ = ψ̈cθ − ψ̇ θ̇sθ + φ̈ (9.42)
dt
• An intermediate frame other than inertial or body-fixed. The terms in the trans-
port theorem do not vanish. For an intermediate reference frame, say, F , the expression
for the angular acceleration becomes
A B A d A d
ωB F A
ωB A
ω F × Aω B
 
α = = + (9.43)
dt dt
Three-Dimensional Kinematics of Rigid Bodies 473

As discussed earlier, we select to use a reference frame other than an inertial or body-
fixed frame only when it is convenient to do so. The next section discusses an important
application of such a reference frame in conjunction with axisymmetric bodies.

Example 9.2

!" #"
$%&'()*+

! !

# ȥ

ȥÚ !‫މ‬
"

#‫މ‬ ȥ #

FIGURE 9.10
Axes for aircraft. a) Before rotations, b) after rotation by ψ.

Derive the angular velocity expressions for aircraft (and also ground vehicle) dynamics
problems, using a 3-2-1 Euler angle transformation.
The coordinate system traditionally used in air or ground vehicles is shown in Figures
9.10–9.12. The inertial coordinates are denoted by XY Z and the body-fixed coordinates
attached to the aircraft are xyz. The Z axis is the local vertical with the positive direction
pointing down. The Euler angle transformations are
• A rotation about the Z axis by ψ (the heading or yaw angle), leading to the X 0 Y 0 Z 0
axes (rotation matrix [R1 ]).
• A rotation about the Y 0 axis by θ (the attitude or pitch angle), leading to the X 00 Y 00 Z 00
axes (rotation matrix [R2 ]).
• A rotation about the X 00 axis by φ (the bank or roll angle), leading to the body-fixed
xyz coordinate system (rotation matrix [R3 ]).
Rotation by the attitude angle is shown in Figure 9.11, for when the airplane pitches up.
The final rotation is by φ, the roll angle. Using this configuration, the xyz axes are attached
to the aircraft, and the xz plane is the plane of symmetry of the aircraft, as shown in Figure
9.12.
The associated rotation matrices are
     
cψ sψ 0 cθ 0 −sθ 1 0 0
[R1 ] =  −sψ cψ 0  [R2 ] =  0 1 0 [R3 ] =  0 cφ sφ  [a]
0 0 1 sθ 0 cθ 0 −sφ cφ
so the relationship between the inertial and body-fixed coordinates is written as
     
x X X
 y  = [R3 ] [R2 ] [R1 ]  Y  = [R]  Y  [b]
z Z Z
474 Applied Dynamics

Ú
‫׋‬ ;үү[

ș

șÚ ș
==ү =үү
<ү<үү

FIGURE 9.11
Rotation by attitude angle θ about the Y 0 axis.

Ȧ#!"+,--

Ȧ!!"()*

Ȧ"!"#$%&' !
"

FIGURE 9.12
Body-fixed axes for aircraft. The xz plane is the plane of symmetry.

and the combined transformation matrix is


 
cψcθ sψcθ −sθ
[R] = [R3-2-1 ] =  −sψcφ + cψsθsφ cψcφ + sψsθsφ cθsφ  [c]
 
sψsφ + cψsθcφ −cψsφ + sψsθcφ cθcφ
The angular velocity vector is
ω = ψ̇K + θ̇J0 + φ̇I00 [d]
0
Using the approach in the previous section, we obtain K and J in terms of the unit vectors
associated with the body axes by reading the third and second columns of [R], respectively,
and setting ψ = 0 when reading the second column. Also note that I00 = i. Hence, the
angular velocity becomes
ω = ωx i + ωy j + ωz k = ψ̇ (−sθi + cθsφj + cθcφk) + θ̇ (cφj − sφk) + φ̇i

     
= −ψ̇sθ + φ̇ i + ψ̇cθsφ + θ̇cφ j + ψ̇cθcφ − θ̇sφ k [e]
Three-Dimensional Kinematics of Rigid Bodies 475

The body angular velocities ωx , ωy , and ωz are commonly referred to as roll, pitch, and
yaw rates, respectively, and are shown in Figure 9.12. These angles are related to the time
derivatives of the heading, attitude, and bank angles ψ, θ, and φ by Equation [e] as
    
ωx −sθ 0 1 ψ̇
 ωy  =  cθsφ cφ 0   θ̇  [f ]
ωz cθcφ −sφ 0 φ̇

When the attitude angle becomes θ = π/2, the first and third columns of the coefficient
matrix [B] above become identical, leading to a singularity. For civilian aircraft or ground
vehicles, this is not a major problem, as the attitude angle θ rarely exceeds π/2.

9.4 Axisymmetric Bodies


When dealing with axisymmetric bodies, such as disks, cones, cylinders, and spinning tops,
we usually are more interested in the velocity and acceleration of a point on the symmetry
axis, such as the center of mass, than the velocity or acceleration of an actual point on the
body. For example, when viewing a rolling disk from a distance, the interest is in how fast
the disk is moving, how the disk’s orientation is changing with respect to the vertical, or
with the trajectory of the center of the disk. Unless there is a compelling reason, we are
not interested in the motion of a certain point on the disk. This observation becomes even
more relevant when considering rapidly rotating bodies.
It turns out that we can analyze relevant properties of the motion of an axisymmetric
body by means of a reference frame obtained after rotation by the precession and nutation
angles. For example, when using a 3-1-3 transformation, the reference frame is obtained
after the 3 and 1 rotations (3-1 transformation).

a3, a3
f3
f2

a2
a2
a1

a1, f1

FIGURE 9.13
The shape (or F ) frame for a spinning top.

Consider the spinning top in Figure 9.13 and a 3-1-3 transformation sequence with
a1 a2 a3 as the inertial frame. The first rotation is about a3 by the precession angle φ,
leading to the a01 a02 a03 (or A0 ) frame. The second rotation is about a01 by the nutation angle
476 Applied Dynamics

θ and the resulting frame is called the shape frame or the F frame, with coordinate axes
f1 f2 f3 . The axis f3 is the axis of symmetry. When looking from a distance at a spinning
top that does not have any discernible markings on its outside, it is the nutation of this
axis that is observed.
The angular velocity of the shape frame is
0
A0 00
ωf = A
ωF = A
ωA + ωA = φ̇a3 + θ̇a01 (9.44)

Noting that a3 = a03 and a01 = f1 , and


    0 
f1 1 0 0 a1
 f2  =  0 cθ sθ   a02  (9.45)
f3 0 −sθ cθ a03

so that a03 = sin θf2 + cos θf3 , the angular velocity of the shape frame can be written as
A
ωF = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 (9.46)

The spin angular velocity is


F
ωs = ω B = ψ̇f3 (9.47)

It follows that the total angular velocity of the top is


 
A
ω = ωB = ωf + ωs = A
ωF + F
ωB = θ̇f1 + φ̇ sin θf2 + ψ̇ + φ̇ cos θ f3 (9.48)

The velocity of a point along the symmetry axis, say, at a distance L from the base so
that the position vector is r = Lf3 , can be written as
   
v = ω × r = θ̇f1 + φ̇ sin θf2 + ψ̇ + φ̇ cos θ f3 × Lf3 = Lφ̇ sin θf1 − Lθ̇f2 (9.49)

As expected, the spin rate of the top does not affect the velocity of a point on the symmetry
axis.
Consider next the angular acceleration. Using the shape frame, we can find the angular
acceleration by invoking the transport theorem. Before doing so, observe that

ω f × ω = ω f × (ω
ωf + ω s) = ω f × ω s (9.50)

The transport theorem for the angular acceleration gives


d d d
α = ω = ω rel + ω f × ω = ω rel + ω f × ω s (9.51)
dt dt dt
or, in terms of the A and B frame notation,

A B A dA B F dA B A
α = ω = ω + ωF × F ωB (9.52)
dt dt
Expanding the above equation, the relative angular acceleration becomes
d d    
ω rel = F Aω B = θ̈f1 + φ̈ sin θ + φ̇θ̇ cos θ f2 + φ̈ cos θ − φ̇θ̇ sin θ + ψ̈ f3 (9.53)
dt dt
and the transport term is
 
A
ωf × ωs = ωF × F ωB = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 × ψ̇f3
Three-Dimensional Kinematics of Rigid Bodies 477

= φ̇ψ̇ sin θf1 − θ̇ψ̇f2 (9.54)


Adding the above two equations gives the angular acceleration as
     
α = Aα B = θ̈ + φ̇ψ̇sθ f1 + φ̈sθ + φ̇θ̇cθ − θ̇ψ̇ f2 + φ̈cθ − φ̇θ̇sθ + ψ̈ f3 (9.55)

which is considerably simpler than its counterpart when the spin angle transformation is
included, that is, the third coordinate transformation in the 3-1-3 set.
Given the expression for the angular velocity in Equation (9.48), we can differentiate it
directly to arrive at the angular acceleration, noting that the derivatives of the unit vectors
of the shape frame have the form
d  
fi = ω f × fi = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 × fi i = 1, 2, 3 (9.56)
dt
Example 9.3

a3
f3
g

– f2
G

L a2
O

f1, a1

FIGURE 9.14
Spinning top.

The spinning symmetric top in Figure 9.14 has the following rotational parameters:
precession rate φ̇ = 0.3 rad/s, increasing with the rate of φ̈ = 0.05 rad/s2 , constant nutation
angle θ = 30◦ , and constant spin rate of ψ̇ = 5 rad/s. Assuming that the bottom point
(pivot) of the top does not move, find the velocity and acceleration of the center of mass of
the top, as well as the angular velocity and acceleration.
The angular velocity and angular acceleration are given by Equation (9.48) and Equation
(9.55), respectively, for a 3-1-3 transformation. From the given data, θ̇ = 0, θ̈ = 0, ψ̈ =
0, and the angular velocity and angular acceleration expressions in Equation (9.48) and
Equation (9.55) reduce to
 
ω = φ̇ sin θf2 + ψ̇ + φ̇ cos θ f3 = 0.15f2 + 5.2598f3 rad/s [a]

2
α = φ̇ψ̇sθf1 + φ̈sθf2 + φ̈cθf3 = 0.75f1 + 0.025f2 + 0.0433f3 rad/s [b]
The position vector from pivot O to the center of mass is rG = Lf3 . The velocity of the
center of mass is found by using the expression for rotation about the fixed point O
vG = ω × rG = (0.15f2 + 5.2598f3 ) × Lf3 = 0.15Lf1 [c]
478 Applied Dynamics

The acceleration is calculated in a similar way, by writing

aG = α × rG + ω × (ω
ω × rG ) = α × rG + ω × vG [d]

Substituting the appropriate values, the acceleration becomes

aG = (0.75f1 + 0.025f2 + 0.0433f3 ) × Lf3 + (0.15f2 + 5.2598f3 ) × 0.15Lf1

= 0.025Lf1 + 0.03897Lf2 − 0.0225Lf3 [e]


It is clear that the shape frame results in expressions that are simpler and more mean-
ingful than using a body-fixed frame. Also, as discussed earlier, we can obtain the velocity
and acceleration by direct differentiation of the position vector. To do this, it is necessary
to consider the actual expressions for position and velocity and to differentiate them before
their actual values are substituted. For example, the velocity of the center of mass can be
written as
d
vG = rG = Lḟ3 [f ]
dt
Noting that, since θ̇ = 0, the angular velocity of the shape frame is

ω f = φ̇ sin θf2 + φ̇ cos θf3 [g]

the rate of change of f3 is found from


d  
f3 = ω f × f3 = φ̇ sin θf2 + φ̇ cos θf3 × f3 = φ̇ sin θf1 [h]
dt

Hence, the velocity of the center of mass is vG = Lφ̇ sin θf1 .


Directly differentiating the velocity yields an expression for the acceleration as
d   d
aG = vG = L φ̈ sin θ + φ̇θ̇ cos θ f1 + Lφ̇ sin θ f1 [i]
dt dt
The rate of change of f1 is found via
d
f1 = ω f × f1 = φ̇ cos θf2 − φ̇ sin θf3 [j]
dt

so that the acceleration (for θ̇ = 0) becomes

aG = Lφ̈ sin θf1 + Lφ̇2 sin θ cos θf2 − Lφ̇2 sin2 θf3 [k]

We can show that substitution of the values for the nutation angle θ = 30◦ and for the
precession and spin rates, φ̇ and ψ̇, gives the same answer as Equation [e].

9.5 Rolling
Rolling was defined in Chapter 3 as a continuous sequence of points on one body coming
into continuous contact with a continuous sequence of points on another body. For this to
happen, the bodies in contact must have smooth contours, with no sharp corners. Here, the
developments of Chapter 3 are extended to general three-dimensional motion.
Three-Dimensional Kinematics of Rigid Bodies 479

Plane of contact

C1 2
C2
Common
normal n

FIGURE 9.15
Contact surface for rolling.

Contact of two bodies takes place in a plane of contact or contact plane, as shown in
Figure 9.15. The contacting points are denoted by C1 and C2 , with velocities of vC1 and
vC2 , respectively. Perpendicular to this plane is the common normal n. The unit vector
along the common normal is denoted by n. This axis is perpendicular to the curvatures of
both contacting points, so the centers of curvature of both contacting bodies at the contact
point lie along the common normal.
The rolling constraint basically states that the contacting points do not have a relative
velocity with respect to each other along the common normal, or

(vC1 − vC2 ) · n = 0 (9.57)

Table 9.2 lists the requirements for contact to take place. A body rolling over a fixed surface
with slipping and sliding has five degrees of freedom.

TABLE 9.2
Requirements for contact to take place

Requirement Description
Smooth contour on Body 1 Radius of curvature and center of curvature exist
Smooth contour on Body 2 Radius of curvature and center of curvature exist
Continuous contact Contacting bodies do not separate from each other

The motion of the contacting points with respect to each other defines the type of rolling.
If the contacting points have the same velocity, the motion is referred to as roll without
slip. If the contacting points have different velocities, this means that the contact points
slide with respect to each other, and the motion is referred to as roll with slip or roll with
slide. The existence of slipping depends on the forces acting on the rolling bodies, as well
as on the amount of friction between them. The no-slip condition can be expressed as

vC1 = vC2 (9.58)

which represents three constraints. A body rolling without slipping on a fixed surface has
three degrees of freedom.
The accelerations of the contacting points are different, whether there is slipping or not.
The contacting points approach each other, they come into contact, and then they separate
480 Applied Dynamics

and move away from each other. The constraint force associated with the rolling constraint
is applied to the contacting points only for one instant. Also, accelerations of the contacting
points usually have components along the contact plane.
Our primary interest is flat axisymmetric bodies, such as disks, that roll on fixed surfaces.
Two types of coordinate transformations will be used to describe rolling. In each case, the
first two rotations define the shape frame.

9.5.1 Disk Originally Lying Flat

==ү
‫׋‬Ú

‫׋‬
#
" !
‫׋‬

%
$

FIGURE 9.16
Initial position and precession by φ.

When the initial position of the disk is flat, it is preferable to select an Euler angle
rotation sequence whose first and third indices are the same, such as 3-1-3. Begin with an
inertial set of axes XY Z, with the disk lying on the XY plane, as shown in Figure 9.16. The
disk has radius R and thickness h. The first Euler angle transformation is by the precession
angle φ about Z, leading to the X 0 Y 0 Z 0 coordinates. Figure 9.17a depicts the disk after the
second rotation, by the nutation angle θ, about X 0 . This second rotation leads to the xyz
axes. The xyz axes depict the shape frame, so the spin ψ̇ is about the z axis, as shown from
the side in Figure 9.17b.
The transformation matrix [R1 ], which rotates the XY Z coordinates to X 0 Y 0 Z 0 by the
precession angle φ about Z, and matrix [R2 ], which rotates the X 0 Y 0 Z 0 coordinates to xyz
by the nutation angle θ, are
   
cos φ sin φ 0 1 0 0
[R1 ] =  − sin φ cos φ 0  [R2 ] =  0 cos θ sin θ  (9.59)
0 0 1 0 − sin θ cos θ

so the combined rotation matrix is


 
cos φ sin φ 0
[R] = [R2 ] [R1 ] =  − sin φ cos θ cos φ cos θ sin θ  (9.60)
sin φ sin θ − cos φ sin θ cos θ

The descriptions of column vector {r} in the xyz and XY Z frames are related by
{xyz r} = [R] {XY Z r}. It follows that the unit vectors of the inertial frame and shape frame
are related by

i = cos φI + sin φJ j = − sin φ cos θI + cos φ cos θJ + sin θK


Three-Dimensional Kinematics of Rigid Bodies 481

a) b)
Z Z
y
z
z y
h

G
R R
Rsin G
Rsin

C C
,x

FIGURE 9.17
a) Rotation by nutation angle θ, b) side view.

k = sin φ sin θI − cos φ sin θJ + cos θK (9.61)

From Equation (9.48), the angular velocity of the disk is


 
ω = θ̇i + φ̇ sin θj + ψ̇ + φ̇ cos θ k = ωx i + ωy j + ωz k (9.62)

so ωx = θ̇, ωy = φ̇ sin θ, and ωz = ψ̇ + φ̇ cos θ. The angular velocity of the reference frame
xyz is
ωy
ω f = θ̇i + φ̇ sin θj + φ̇ cos θk = ωx i + ωy j + k (9.63)
tan θ
The angular velocity associated with spin is
 ωy 
ω s = ω − ω f = ψ̇k = ωz − k (9.64)
tan θ
Denote the position of the center of mass G by rG = XI + Y J + ZK. For a disk
of thickness h, the position vector between the center of mass and the contact point is
rC/G = −Rj − h2 k, where R is the radius. When the disk is thin, its thickness can be
ignored and rC/G ≈ −Rj. Thus, the relative velocity expression between the center of mass
and the contact point becomes

vC = vG + ω × rC/G

   
= ẊI + Ẏ J + ŻK + θ̇i + φ̇ sin θj + ψ̇ + φ̇ cos θ k × −Rj (9.65)

Carrying out the calculations and expressing in terms of the inertial axes, the velocity of
the contact point becomes
   
vC = Ẋ − Rθ̇ sin φ sin θ + R φ̇ cos θ + ψ̇ cos φ I
482 Applied Dynamics
     
+ Ẏ + Rθ̇ cos φ sin θ + R φ̇ cos θ + ψ̇ sin φ J + Ż − Rθ̇ cos θ K (9.66)

Noting that the common normal is the Z axis, the fundamental constraint associated
with rolling (with or without slip) states that the velocity of point C has no component
along the common normal, so vC · K = 0. Setting the components of vC in the Z direction
to zero yields the constraint equation

Ż = Rθ̇ cos θ (9.67)

This constraint is the time derivative of

Z = R sin θ (9.68)

so the fundamental constraint associated with roll is holonomic; that is, it can be represented
in terms of position variables. The constraint has a physical explanation: It represents the
height of the center of mass from the contact surface. A disk rolling with slip has five degrees
of freedom.
Next, consider roll without slip, so that the velocity of the contact point C is zero. Using
the expression for the velocity of the contact point in terms of the inertial axes in Equation
(9.66), and setting vC = 0, two additional constraint equations are obtained, which are
 
Ẋ = Rθ̇ sin φ sin θ − R φ̇ cos θ + ψ̇ cos φ

 
Ẏ = −Rθ̇ cos φ sin θ − R φ̇ cos θ + ψ̇ sin φ (9.69)

These equations are not perfect differentials, and they cannot be integrated to a form
similar to Equation (9.68). Hence, the constraints associated with the no-slip condition are
nonholonomic.
For roll without slip on a fixed surface, since vC = 0, we can write vG = ω × rG/C .
Expressing this relationship in terms of the shape frame yields

vG = ω × rG/C = (ωx i + ωy j + ωz k) × Rj

 
= −Rωz i + Rωx k = −R φ̇ cos θ + ψ̇ i + Rθ̇k (9.70)

Differentiation of the above equation gives the acceleration of the center of the disk for
roll without slip as
 ωy 
aG = v̇Grel + ω f × vG = −Rω̇z i + Rω̇x k + ωx i + ωy j + k × (−Rωz i + Rωx k)
tan θ
 ωy ωz 
= R (−ω̇z + ωx ωy ) i + R −ωx2 − j + R (ω̇x + ωy ωz ) k (9.71)
tan θ
In a similar way, the angular acceleration of the disk (whether there is roll without slip
or not) is obtained as
 ω   ω 
α = ω̇ω rel + ω f × ω s = ω̇x i + ω̇y j + ω̇z k + ωx i + ωy j + y k × ωz − y k
tan θ tan θ
!
ωy2  ωx ωy 
= ω̇x + ωy ωz − i + ω̇y − ωx ωz + j + ω̇z k (9.72)
tan θ tan θ
Three-Dimensional Kinematics of Rigid Bodies 483

The above two equations can also be expressed in terms of the Euler angles. Two types
of slip are possible for the general motion of a disk. One is spin slip, where the disk rotates
too fast for friction to prevent sliding. This motion is in the direction of the heading (in the
X 0 direction). The second type of slip is slide slip and it involves sideways sliding (in the
Y 0 direction) of the contact point. We will discuss different types of slipping in more detail
in Chapter 11 within the context of the kinetics of roll.

9.5.2 Beginning with the Disk as Vertical


An alternate, and sometimes more representative, description of the rotation of a disk can
be accomplished by considering the initial position of the disk as vertical, as shown in Figure
9.18a.

a) b)
Heading
X z
Z
z
Initial
After
heading
rotation y

y Wheel
plane
Y

Top view Side view

FIGURE 9.18
Rotation sequence: a) top view showing heading angle ψ, b) lean angle θ.

The first rotation is about the Z axis (coming out of the page) and by the angle ψ,
known as the heading angle, resulting in the X 0 Y 0 Z 0 axes. This rotation is shown in Figure
9.18a, viewed from the top. In the absence other rotations, the disk would travel upright
and in a straight line along the X 0 axis. The second rotation, shown in Figure 9.18b, is
about the X 0 axis by θ, which is known as the lean angle, resulting in the xyz axes, or the
shape frame. The disk lies on the xz plane and the spin is about the y axis by φ.
The rear view of the disk is shown in Figure 9.19a. For this reference frame, the disk is
moving forward (in the x direction) when φ̇ > 0, as depicted in Figure 9.19b.
The transformation matrices are
   
cψ sψ 0 1 0 0
[R1 ] =  −sψ cψ 0  [R2 ] =  0 cθ sθ  (9.73)
0 0 1 0 −sθ cθ
The combined rotation matrix [R] = [R2 ][R1 ] is the same as in Equation (9.60). However,
the geometry is quite different, as the initial starting geometries are different. The lean angle
is measured from the vertical and the spin is about the y axis.
The angular velocity of the disk is
ω = ψ̇K0 + θ̇i + φ̇j (9.74)
484 Applied Dynamics

a) b)
Z, Z
z z

R G

x
C
Rear view Side view

FIGURE 9.19
a) Rear view of rolling disk, b) side view of xz plane and roll rate φ̇.

and, using the definitions of the unit vectors in Equation (9.61), and noting from Figure
9.18b that K0 = cθk + sθj, we can write the angular velocity in terms of the axes of the
shape frame as
 
ω = θ̇i + φ̇ + ψ̇ sin θ j + ψ̇ cos θk (9.75)

The position vector from the contact point C to the center of mass G is rG/C = Rk.
For roll with slip, the velocity of the contact point C can be expressed as vC = ẊC I + ẎC J,
where XC and YC are the coordinates of the contact point C in the inertial XY Z frame.
The velocity of the center of mass becomes

vG = vC + ω × rG/C = ẊC I + ẎC J + ω × rG/C = ẊC I + ẎC J

     
+ θ̇i + φ̇ + ψ̇ sin θ j + ψ̇ cos θk × Rk = ẊC I + ẎC J + R φ̇ + ψ̇ sin θ i − Rθ̇j (9.76)

As discussed earlier, roll with slip is a five-degrees-of-freedom system. For roll without
slip, the velocity of the contact point is zero, and the velocity of the center of the disk
becomes
 
vG = ω × rG/C = R φ̇ + ψ̇ sin θ i − Rθ̇j (9.77)

The advantages of the above formulation is that the way we arrive at the shape frame
is more intuitive and the constraints on XC and YC have a simpler form for the case of no
slip. However, the constraints in terms of the coordinates of the center of the disk XG and
YG are nonholonomic.
To find the angular acceleration of the disk, express the angular velocity as ω = ωx i +
ωy j + ωz k so that the angular velocity of the shape frame becomes

ω f = θ̇i + ψ̇ sin θj + ψ̇ cos θk = ωx i + ωz tan θj + ωz k (9.78)


Three-Dimensional Kinematics of Rigid Bodies 485

Direct differentiation of the angular velocity gives the angular acceleration as


α = ω̇x i + ω̇y j + ω̇z k + ωx i̇ + ωy j̇ + ωz k̇ (9.79)
The local derivatives are
   
ω̇x i + ω̇y j + ω̇z k = θ̈i + φ̈ + ψ̈ sin θ + ψ̇ θ̇ cos θ j + ψ̈ cos θ − ψ̇ θ̇ sin θ k (9.80)

and the time derivatives of the unit vectors are


i̇ = ω f × i = ψ̇ cos θj − ψ̇ sin θk j̇ = ω f × j = −ψ̇ cos θi + θ̇k

k̇ = ω f × k = ψ̇ sin θi − θ̇j (9.81)


The above terms are combined to calculate the angular acceleration.

9.5.3 Steady Precession


An interesting special case of rolling without slip is steady precession, also known as steady
motion, where the nutation angle (or lean angle) remains constant while the disk rolls. This
section examines the kinematics of steady motion.
Using the 3-1-3 transformation sequence, the velocity of the center of the disk is given
by Equation (9.70) for roll without slip. Setting the nutation rate to zero, θ̇ = 0, the velocity
of the center of the disk becomes
 
vG = −R φ̇ cos θ + ψ̇ i (9.82)

Note that for a positive spin rate ψ̇ (which usually is much larger than the precession rate
φ̇), the velocity will be in the −x direction, as can be observed from Figure 9.17a.
As will be derived while studying the dynamics of steady precession in Chapter 11, when
the nutation angle is constant so are the precession and spin rates. Hence, every term in
the above equation is a constant. Neglecting the thickness of the disk, the speed v of the
disk, given by
 
v = R φ̇ cos θ + ψ̇ = const (9.83)

is also constant. Note that the velocity of the center of the disk is vG = −vi. Hence, the
center of the disk follows a circular path. The radius of this path, denoted by ρ, is the
velocity of the center of mass v divided by the precession rate φ̇
v
ρ = − (9.84)
φ̇
We can combine the above two equations and express the spin rate as
 ρ
ψ̇ = −φ̇ cos θ + (9.85)
R
The above equation provides a relationship among four parameters: the precession rate
φ̇, the spin rate ψ̇, the nutation angle θ, and the radius of curvature of the center of the disk
ρ. As will be developed in Chapter 11, there is one more equation relating these parameters
which originates from the dynamics of the problem and it is
4gρ2 cot θ
v2 = (9.86)
6ρ + R cos θ
Steady precession is defined in terms of two parameters. We usually select these param-
eters as v and θ and then calculate the radius of curvature, as well as the precession and
spin rates.
486 Applied Dynamics

Example 9.4
A disk of radius 30 cm is released with a speed of 5 m/s and nutation angle of 60◦ (angle
with vertical is 30◦ ). It rolls without slipping and with steady precession. Calculate the
precession rate and the time it takes for the disk to complete one full circle.
Equation (9.86) can be written as a quadratic equation in terms of the radius ρ as

4g cot θρ2 − 6v 2 ρ − Rv 2 cos θ = 0 [a]

Evaluating the parameters with g = 9.807 m/s2 gives


2
4g cot θ = 22.648 m/s 6v 2 = 150 m2 /s2 Rv 2 cos θ = 3.75 m3 /s2 [b]

Solving Equation [a] for ρ yields two roots, one positive and one negative. The negative
root is discarded as it does not have a physical interpretation. The positive root is ρ =
6.6479 m. From Equation (9.84) the precession rate is obtained as
v 5
φ̇ = − = − = −0.7521 rad/s [c]
ρ 6.6479

The time T to complete a revolution is the time it takes for the precession angle φ = φ̇T
to become −2π radians. This time can be found from
2π 2π
T = = = 8.3540 s [d]
φ̇ 0.7521

The spin rate ψ̇ is found from Equation (9.85) as


 
1 6.6479
ψ̇ = 0.7521 + = 17.04 rad/s [e]
2 0.3

The steady precession of a disk raises some interesting questions, such as what the initial
speed of the disk should be so that there is no slip, and what amount of friction is required
to prevent slip. These questions will be addressed when studying the dynamics of rolling in
Chapter 11.

Example 9.5
The cone in Figure 9.20a is of height L and has apex angle β. It rolls on a flat surface
without slipping. The center B of the base of the cone has a constant speed v. Find the
angular velocity and angular acceleration of the cone.
To find the angular velocity of the cone we can make use of instant centers. Figure 9.20b
shows the geometry. The h1 h2 h3 axes move with the line of contact between the cone and
the horizontal surface, with the h3 axis as the line of contact. The vertical distance between
point B and the line of contact is BA = L sin β. Because of the no-slip condition, A is an
instant center and the velocity of B is equal to the angular velocity times the height. The
total angular velocity of the cone is then
v
ω = [a]
L sin β

The angular velocity in vector form is ω = −v/(L sin β)h3 . To see this, note that vB =
ω × rB/A , with vB = vh1 and rB/A = L sin βh2 .
The motion of the cone can be visualized as the superposition of two angular velocities.
The first angular velocity ω1 is that of a cone which is rotating only in the horizontal
Three-Dimensional Kinematics of Rigid Bodies 487

"$ "$ %$
!" #"
Ȧ+
#&'
Ȧ$ )#&' ȕ
%% (
$ '
'
# *
ȕ ! ȕ
"% !
& #()*ȕ
"%
"+

FIGURE 9.20
Rolling cone.

plane (h1 h2 ). The second angular velocity, ω2 , is the angular velocity of the cone about its
symmetry axis (f3 ). The h1 h2 h3 coordinates are rotating with angular velocity ω1 . Hence,
ω 1 is
ω 1 = ω1 h2 [b]
Considering that the horizontal distance from B to the h2 axis is L cos β leads to an
expression for the angular velocity of the h1 h2 h3 axes as
v v
ω1 = or ω1 = h2 [c]
L cos β L cos β
The angular velocity of the cone with respect to the shape frame f1 f2 f3 is ω 2 = ω2 f3 .
In terms of the h1 h2 h3 axes, this angular velocity is
ω 2 = ω2 f3 = ω2 sin βh2 + ω2 cos βh3 [d]
We can use Equations [b] and [d] to write the angular velocity vector as
v v
ω = ω1 + ω2 = h2 + ω2 sin βh2 + ω2 cos βh3 = − h3 [e]
L cos β L sin β
which can be solved for ω2 as
v ω
ω2 = − = − [f ]
L cos β sin β cos β
Noting that h3 = cos βf3 − sin βf2 , and R = L tan β the angular velocity can also be
expressed as
v v v cos β v v
ω = − h3 = f2 − f3 = f2 − f3 [g]
L sin β L L sin β L R
It is convenient to use the H frame to calculate the angular acceleration. Because v is
ω rel = 0. Also, the angular velocity of the
constant, both ω 1 and ω 2 are also constant, so ω̇
reference frame is ω f = ω 1 . Hence, the angular acceleration becomes
α = ω rel + ω f × ω = ω 1 × (ω
ω1 + ω 2) = ω 1 × ω 2

v2
 
v v
= h2 × − (sin βh2 + cos βh3 ) = − 2 h1 [h]
L cos β L cos β sin β L cos β sin β
The rolling of the cone can also be viewed from an Euler angle point of view, with ω1
denoting the precession rate, constant nutation angle β (i.e., zero nutation rate), and spin
rate of ω2 .
488 Applied Dynamics

9.6 Orientation Change by Successive Rotations


An interesting application of successive coordinate transformations is to apply them in a
reciprocal way so as to change the orientation of a body and make the body rotate to a
desired orientation. In kinematics and dynamics we are generally interested in the result
of successive transformations that arise from the applications of external moments. This
section studies the special case where orientation changes are brought about by applications
of internal moments.
This concept was made famous by the “falling cat” example (Kane) where an explanation
was provided as to how a cat which falls from a height wiggles its feet and flexes its body in
a reciprocating manner so that when the cat lands, its feet hit the ground first. The wiggling
and flexing change the cat’s orientation, while not altering its angular momentum. There
are no external moments acting on the cat. The only external force, neglecting aerodynamic
effects, acting on the cat is its weight, which goes through the center of mass (even though
the center of mass location changes).
A cat does the wiggling maneuver naturally, and there are other examples that can be
found from nature, such as the motion of snakes and certain bacteria, which change their
orientation by moving parts of their body. Figure 9.21 shows the positions of the cat during
a fall.3

FIGURE 9.21
Falling cat (rotate book clockwise to see better).

Consider a body and a set of fixed coordinates XY Z as well as a set of coordinates


xyz attached to the body. The interest is in body-fixed rotations. Rotate first about one
of the axes XY Z by an angle θ1 so that the rotation matrix between the first and second
orientations is [R1 ]. Next, rotate the body about another one of its axes by θ2 , so the
second rotation matrix is [R2 ]. The third rotation is such that it is the opposite of the first,
[R3 ] = [R1 ]T , and the fourth rotation is the opposite of the second, [R4 ] = [R2 ]T . The
combined rotation sequence becomes
T
{x} = [R] {X} = [R2 ]T [R1 ] [R2 ] [R1 ] {X} (9.87)

where {x} = [x y z]T , {X} = [X Y Z]T .


We can show that if the two rotation angles are small,4 then the resulting combined
rotation matrix [R] can be viewed as the approximation of a rotation matrix about the
axis not considered in the rotations by θ1 and θ2 and having the rotation angle θ1 θ2 (in
3 Please do not try this maneuver with your cat. There are cats whose sense of balance is not strong

enough to complete the maneuver successfully.


4 The validity of the smallness approximation will be discussed in the example that follows.
Three-Dimensional Kinematics of Rigid Bodies 489

radians). The right-hand rule is followed in determining the sign of the resulting rotation.
For example, if the first rotation is about the y axis by θy and the second rotation is about
the z axis by θz , the net effect will be a rotation about the x axis of magnitude θy θz . Hence,
if it is not possible to perform rotations about the x axis by means of an external moment,
we can obtain a rotation about the x axis by performing a series of transformations about
the y and z axes. The algebra involved in the proof is complicated and is not given here.
For example, for two rotations, each of magnitude of 0.1 radians (about 5.7◦ ), the net
rotation effect will be about 0.01 radians, which is about 0.57◦ , indeed a small angle. How-
ever, by performing several successive rotations we can accomplish a finite angle change.
Creeping animals and astronauts in space do this when they want to reorient themselves.
Even when the rotation angles are large, the resulting rotation still has a resemblance to a
rotation about the third direction.
"

FIGURE 9.22
Body-fixed coordinate axes for cat.

Let us return to the cat in Figure 9.21 and define a body fixed coordinate system
xyz, with the x axis along the spine of the cat (positive towards the head), z axis as the
perpendicular to the spine of the cat, so that the y axis comes out of the paper, as shown
in Figure 9.22. Comparing the second snapshot, where the cat has begun its free fall, with
the last snapshot, where the cat is about to land, the end maneuver of the cat has to be
a rotation about the local x axis. To effect this, the cat first wiggles itself about the y
axis (third snapshot) and then about the z axis (fourth snapshot). The fifth snapshot is a
rotation about the negative y axis, followed by a reverse rotation about the z axis in the
sixth snapshot. The last snapshot shows the cat safely landing on its four feet.

Example 9.6
This example analyzes the limits of the accuracy of the orientation change maneuver and
evaluates the [R] matrix in Equation (9.87). For θ1 = 0.1 rad (5.7◦ ) about the x axis and
θ2 = 0.2 rad (11.4◦ ) about the y axis we get
 
0.9998 0.0198 0.0010
[R(θ1 = 0.1, θ2 = 0.2)] =  −0.0198 0.9998 0.0020  [a]
−0.0010 −0.0020 1.0000

This matrix very closely resembles a 3 rotation, or a rotation about the z axis. R11 and
R22 are the same and R12 = −R21 . The elements R13 , R23 , R31 , and R32 , which would be
zero in a 3 rotation, are not zero but they are very small, more than 10 times smaller than
R12 . The rotation angle is θ = sin−1 (0.0198) = 0.0198 rad. Comparing this value with the
490 Applied Dynamics

quantity θ1 θ2 = 0.02, the results are in excellent agreement, with less than 1% difference.
If the rotation angle is calculated from R11 we get θ = cos−1 (0.9998) = 0.0200.
To investigate the range of accuracy of the rotation matrix assumption, consider the
same rotation sequence above and select the angles over 20◦ such that the small angles
assumption is almost violated. For θ1 = 0.4 rad (22.9◦ ) and θ2 = 0.5 rad (28.6◦ ), the
resulting rotation matrix is
 
0.9819 0.1849 0.0421
[R(θ1 = 0.4, θ2 = 0.5)] =  −0.1867 0.9814 0.0439  [b]
−0.0332 −0.0510 0.9981

As can be seen, the resulting rotation matrix still looks like a rotation matrix about
the Z axis, but the terms that should be zero for a 3 rotation are no longer negligible.
Calculating the rotation angle from the 12 element of [R] results in

θ ≈ sin−1 R12 = sin−1 (0.1849) = 0.1860 rad [c]

Comparing this with the value of θ1 θ2 = 0.4 × 0.5 = 0.2 rad leads to a difference of about
7%.

9.7 Interconnections
As discussed in Chapter 3, any mechanism or a system of bodies that move together is
composed of components that are interconnected. Components of a mechanism are generally
referred to as links or linkages, and the connection between the links are enforced by joints.
There are several types of joints. Planar joints, such as the pin joint and prismatic joint,
were discussed in Chapter 3. This section extends the study of joints for three-dimensional
motion.
Joining two bodies by a joint has the effect of imposing constraints on two otherwise free
bodies. It is useful to study joints from the viewpoint of the type of motions they permit,
as well as the type of motions they prevent.

9.7.1 Basic Types of Joints

,. /.

$
! "
!
"
! !
Ȧ"
"
#
#$%&'()'*(+,+%(-

FIGURE 9.23
a) Revolute joint, b) prismatic joint.
Three-Dimensional Kinematics of Rigid Bodies 491

Revolute (also known as pin) joints and prismatic joints were studied in Chapter 3.
Here, we review the constraints that they generate. Figure 9.23 depicts these joints. If body
1 has m degrees of freedom and body 2 is connected to body 1 with a revolute or prismatic
joint, the resulting mechanism has m + 1 d.o.f. A body that is unrestrained has six d.o.f.;
therefore, revolute and prismatic joints each impose five constraints.
For the revolute joint, the only allowed motion of body 2 with respect to body 1 is a
rotation about axis n. We see that ω 2 − ω 1 = ω2 n, with n denoting the unit vector along
the rotation axis n. The two rotational constraints then are

ω2 − ω 1) × n = 0
(ω (9.88)

The translational constraints indicate that the two bodies have no relative motion with
respect to each other at the joint, so

vP1 − vP2 = 0 (9.89)

The five constraints of prevention of two rotations and three translations are enforced
by three forces at the joint, as well as two moments about two axes, both perpendicular to
n, and to each other.
A prismatic joint permits motion of body 2 only in the h direction, so vP2 − vP1 = vh,
with h denoting the unit vector along the axis h. The two translational constraints then are

(vP2 − vP1 ) × h = 0 (9.90)

There are three additional constraints that a prismatic joint imposes. All three are rota-
tional:

ω2 − ω1 = 0 (9.91)

The five constraints of prevention of translation in two directions and rotation about two
axes are enforced by two forces, acting in directions perpendicular to the joint, as well as
three moments that act on the joint.
Two commonly encountered combinations of revolute and prismatic joints are the slider,
also known as collar, and the cylindrical pair, shown in Figures 9.24a and 9.24b, respectively.

$( )( \ү

Ȧ%"&'!"##$%

"
! #
#$
Ȧ ]ү
%
&'(&ү

Ȧ!"##$%

FIGURE 9.24
a) Slider, b) cylindrical pair.
492 Applied Dynamics

A slider permits two degrees of freedom and hence imposes four constraints: two trans-
lational and two rotational. A cylindrical pair, on the other hand, permits three types of
motion: translation of the collar in the x direction, rotation of collar about the x axis, and
rotation of the component (or rod) through a pin joint about the z 0 axis, where the x0 y 0 z 0
axes are attached to the collar.
Another commonly encountered joint is the ball-and-socket joint, also known as a spher-
ical pair, which permits three rotational degrees of freedom, as shown in Figure 9.25, and
hence imposes three constraints that prevent translational motion.

FIGURE 9.25
Ball-and-socket joint.

9.7.2 Combined Sliding and Rotation

!" " #" "

$ $
#$
( #$
" %
2ү %
' !
& &
*+ )
#% #%

FIGURE 9.26
Two cylindrical pairs constraining a rod. a) Guide bars not on same plane, b) guide bars
on same plane.

In Figure 9.26a, rod AB is connected to two cylindrical pairs, each pair sliding inside a
guide bar. The rod has six d.o.f. before it is attached to the joints and each cylindrical pair
imposes three constraints. Hence, the rod has zero degrees of freedom and cannot move,
except for certain special geometries that create redundancies. One such redundancy is
shown in Figure 9.26b, when the guide bars lie on the same plane. The angular velocities
of both collars become equal (to zero) and hence a new degree of freedom emerges. What
happens in Figure 9.26b is similar to exceptions to Gruebler’s equation studied in Chapter
3.
When one of the cylindrical pairs, say, joint A, in Figure 9.26a is replaced by a spherical
Three-Dimensional Kinematics of Rigid Bodies 493

joint, as shown in Figure 9.27, the spherical joint at A imposes two constraints, and hence,
the rod AB has one degree of freedom.

'
%
2ү $
!
& "
)* (

FIGURE 9.27
One joint replaced by a ball-and-socket joint.

9.7.3 Universal Joints


In machine dynamics it is often necessary to transfer the rotational motion of a shaft onto
another shaft that rotates in a different direction. A mechanism used to accomplish this is
a universal joint. There are several kinds of universal joints available, such as the Cardan,
Rzeppa, Double Cardan, Thomas, Weiss, and Davos joints. All these joints accomplish the
task of power transmission in different ways; each is useful for specific applications.

z
b

x
B
a Input shaft

Output A
shaft

– x'

FIGURE 9.28
A Cardan joint.

The Cardan joint in Figure 9.28 is a commonly used and low-cost universal joint. The
joint consists of two shafts attached to forks, known as yokes, linked to each other by a
cross. The two shafts are called the input and output shafts. The relative position of the
cross with respect to the yokes can change, as the cross can rotate inside each of the yokes,
494 Applied Dynamics

making it possible to align the shafts in almost any direction. This gives the universal joint
tremendous versatility and its name, as it can operate when shafts connected by the yokes
are not aligned, vibrate, or change orientation.
A universal joint is a single-degree-of-freedom mechanism when the orientations of the
shafts are fixed, because the angular velocity of one of the shafts determines the angular
velocity of the other. To explore the relationship between the two angular velocities, it is
convenient to split the universal joint into two parts, A and B, as shown in Figure 9.29.
The coordinate axes are selected so that both shafts lie on the xy plane.
The input shaft, shaft B, is aligned with the x direction, and its angular velocity is −θ̇i,
with θ measured from the z axis, as shown in Figure 9.29a. Rotating the xy axes about the
z direction by angle β results in the x0 y 0 axes and places the output shaft, shaft A, along
the negative x0 axis. The angular velocity of the output shaft is −φ̇i0 , with φ measured from
the z axis.
Denoting the unit vectors along the sides of the cross as a and b, and noting that
i0 = cos βi + sin βj and j0 = − sin βi + cos βj, we can express the unit vectors a and b as

b = sin θj + cos θk

a = sin φj0 + cos φk = − sin φ sin βi + sin φ cos βj + cos φk (9.92)

a) b)
z z

b
y b
y

x x
a B
a

A

FIGURE 9.29
a) Input shaft B, b) output shaft A.

It follows that the two vectors a and b must be perpendicular to each other. The dot
product between the two gives the constraint relationship between the angles θ and φ as

a · b = sin θ sin φ cos β + cos θ cos φ = 0 (9.93)

It is of interest to compare the angular velocity of the output shaft in terms of the input
shaft. To this end, rewrite the above equation as
sin φ cos θ
tan φ = = − (9.94)
cos φ sin θ cos β
Three-Dimensional Kinematics of Rigid Bodies 495

which, when differentiated, yields

φ̇ θ̇
2
= 2 (9.95)
cos φ sin θ cos β
We can eliminate φ from the above equation by making use of the trigonometric identity
1
cos2 φ = 1+tan2 φ . Combining Equation (9.94) and Equation (9.95) gives the relationship

between the angular velocities as


cos β
φ̇ = θ̇ (9.96)
1 − sin2 θ sin2 β
The angular velocities of the input and output shafts are not linearly related to each
other, except when β = 0, in which case the shafts are aligned. The Cardan joint is not what
is called a constant velocity joint, making it impractical to use in certain applications. This
is because of the sin2 θ term in the denominator in the above equation, which fluctuates
between 0 and 1, even when the input shaft rotates at a constant speed, thus preventing
the output shaft from having constant speed.
The level of nonlinearity depends on the angle β between the input and output shafts.
In the extreme case, when β = 90◦ , the two shafts are perpendicular and no motion trans-
mission is possible. This is known as gimbal lock. However, if two Cardan joints are used
to join two parallel shafts, the angular velocities of the parallel shafts will be the same.
It should also be noted that there are several constant velocity joints available, such as
Rzeppa, Double Cardan, and Thomas. Each of these joints provides a near constant velocity
relationship between the input and output shafts for a range of operation.

Example 9.7
Rod AB in Figure 9.27 is constrained to move between the horizontal and vertical guide
bars. Given vA = 2 ft/sec downward and a = 6 ft, b = 3 ft, c = 4 ft, find the velocity of
point B and the angular velocity of the rod.
The relative velocity equation for points A and B on the rod is

vB = vA + ω × rB/A [a]

where
vA = −2j ft/sec vB = vB i ft/sec rB/A = 6i − 3j + 4k ft [b]
and the angular velocity of the rod can be expressed as ω = ωx i+ωy j+ωz k. The components
of Equation [a] in each direction lead to three equations. There are four unknowns, vB and
the three angular velocity components. An additional relationship is needed to solve for the
unknowns. This fourth relationship comes from the angular velocity of the rod. The angular
velocity of the rod in terms of collar B is

ω rod = ω collar + ω rod/collar = ωcollar i + ω rod/collar [c]

The angular velocity of the rod with respect to the collar B (or with respect to the
pin joint) is perpendicular to the plane of the pin joint. The perpendicular to this plane is
defined as the cross product of the velocity of the collar B and the vector joining A and B.
Hence,
ω rod/collar = ωrod/collar e [d]
where the unit vector e can be obtained from
rB/A × rB/O
e = ± [e]
rB/A × rB/O
496 Applied Dynamics

The sign of the unit vector can be chosen by convenience. Carrying out the cross product
gives
6i × (6i − 3j + 4k) 24j + 18k
e = − = [f ]
M M

where M is the magnitude of the cross product and has the value M = 182 + 242 = 30.
Hence,
4 3
e = j+ k [g]
5 5
and the angular velocity of the rod can be written as
 
4 3
ω rod = ω collar + ω rod/collar = ωcollar i + ωrod/collar j+ k [h]
5 5

The angular velocity of the rod is now expressed in terms of two parameters, thus reduc-
ing the number of unknowns in the problem to three: the two angular velocity components
and vB .
Introducing Equations [b] and [h] into the relative velocity expression in Equation [a]
gives   
4 3
vB i = −2j + ωcollar i + ωrod/collar j+ k × (6i − 3j + 4k)
5 5
1
= −2j + (−4j − 3k) ωcollar +(25i + 18j − 24k) ωrod/collar ft/sec [i]
5
Equating the components of Equation [i] along the x, y, and z axes results in a set of
three equations and three unknowns in the form

i components: vB = 5ωrod/collar [j]

18
j components: 0 = −2 − 4ωcollar + ωrod/collar [k]
5
24
k components: 0 = −3ωcollar − ωrod/collar [l]
5
Solving Equations [j]–[l] for the unknowns gives
1 8
ωrod/collar = rad/sec ωcollar = − rad/sec vB = 1 ft/sec [m]
5 25
The angular velocity of the rod becomes
8 11 1
ω rod = ω collar + ω rod/collar = − i+ (4j + 3k) = (−8i + 4j + 3k) rad/sec [n]
25 55 25

9.8 4×4 Matrix Description of a General Transformation


This chapter and Chapter 2 considered translations and rotations, as well as relative posi-
tion, velocity, and acceleration relationships. This section introduces a compact notation for
stating relative position expressions by means of 4×4 matrices. This notation is frequently
used in mechanisms and robotics.
Consider a vector ri viewed in the XY Z frame, as shown in Figure 9.30. This vector is
Three-Dimensional Kinematics of Rigid Bodies 497
$
%
!
( !"
" &
!‫މ‬
' #
!!
"

FIGURE 9.30
Translation and rotation.

moved to its final position rf by means of a translation and a rotation. The translation is
by vector d resulting in the intermediate vector r0 = ri + d, and the rotation is described
by the rotation matrix [R]. The coordinates of the initial vector in XY Z are the same as
the coordinates of the rotated vector in the rotated frame xyz. The two frames are related
by
   
x X
 y  = [R]  Y  (9.97)
z Z
It follows from the developments of Chapter 2 that, using column vector notation, the
initial and final position vectors are related by
{rf } = [R]T {ri } + {d} (9.98)
Implicit is the understanding that all of the vectors in the above equation are described
in terms of the same reference frame. In robotics and kinematics, it is preferable to use
the coordinates of the initial frame, since we are interested in the final position of a body
relative to its initial position.
Equation (9.98) can be written as a single matrix expression. Define the 4 × 1 vectors
" # " #
{ri } {rf }
{Pi } = {Pf } = (9.99)
1 1
and the 4 × 4 homogeneous transformation matrix
[R]T {d}
" #
[T ] = (9.100)
{0}T 1
Equation (9.98) can be expressed as
{Pf } = [T ] {Pi } (9.101)
Another way describing the homogeneous transformation matrix is [T ] = [T ([A], {d})].
We can think of a combined translation and rotation operation as the end result of two
transformations: a pure translation transformation [Ttr ] and a pure rotation transformation
[Trot ], where
[R]T {0}
" # " #
[1] {d}
[Ttr ] = [Trot ] = (9.102)
{0}T 1 {0}T 1
498 Applied Dynamics

The order of translation and rotation is important because

[Ttr ] [Trot ] 6= [Trot ] [Ttr ] (9.103)

Example 9.8
A robot is mounted on a base that translates in the Y Z plane and rotates about the Y
axis, as shown in Figure 9.31. The robot consists of a telescoping arm in the X direction,
to which an L-shaped arm is attached by a pin joint. Frame xyz is attached to the arm and
is obtained by rotating the XY Z frame by 30◦ about the Z axis. The tip of the telescoping
arm B has moved to rB = 8I + 2J + 3K. The position of point Q, which is the tip of the
L-shaped arm, is rQ/B = 3i−2j. Find the position of the point Q in terms of the coordinates
of the XY Z frame.
y
x
Y

z
O Q
X
B

X
Z

FIGURE 9.31
Robot arm: first link (OB) translates, second link (BQ) rotates.

We can view this system as a two-link robot. The first link translates in three directions,
while the second link rotates about point B. Point Q is fixed on the second link. The rotation
matrix from XY Z to xyz is
 √ 
3
2 0.5 0

[R] =  −0.5 3 [a]
 
02
0 0 1

and the translation vector is {d} = {rB } = [ 8 2 3 ]T . It follows that the 4 × 4 transfor-
mation matrix has the form
 √3 
2 −0.5 0 8

3
 0.5 0 2
 
[T ] =  2  [b]
 0 0 1 3
0 0 0 1

This homogeneous transformation can be viewed of as the product of two homoge-


neous transformation matrices, [T ] = [T1 ][T2 ], where [T1 ] = [T ([I], {d})] and [T2 ] =
T ([[R]T , {0})]. The first transformation, [T1 ] moves the axes to the new location of B,
while the second transformation, [T2 ] rotates the axes. Note that the displacement from Q
to B is in the rotated axes.
Three-Dimensional Kinematics of Rigid Bodies 499

The initial position of point Q (or rQ/B ) is (in the XY Z frame) {Pi } =
[ 3 −2 0 1 ]T . Using Equation (9.101), the position of point Q is obtained in terms
of the XY Z coordinates as
 √3 
2 −0.5 0 8  3  
11.598


3 −2 
 0.5 0 2   =  1.568 
   
{Pf } = [T ]{Pi } =  2  [c]
 0 0 1 3
 0   3 
1 1
0 0 0 1

so the position vector in terms of the XY Z axes is


 
11.598
{rf } =  1.568  [d]
3

Note that for a multilink mechanism, the analysis above needs to be conducted for each
link separately, resulting in a homogeneous transformation matrix for each link.
Next, consider rotation of the base of the robot about the Y axis. We need to define a
set of inertial X 0 Y 0 Z 0 axes, so that the XY Z axes, to which the telescoping rod is attached,
are obtained by rotating X 0 Y 0 Z 0 by φ about the Y 0 axis. The associated rotation matrix is
 
cos φ 0 − sin φ
[R2 ] =  0 1 0  [e]
sin φ 0 cos φ

It follows that the associated homogeneous transformation matrix is [T0 ] = [T [R2 ]T , {0} ].
The combined transformation matrix becomes

[T ] = [T0 ] [T1 ] [T2 ] [f ]

and the final position is obtained by

{Pf } = [T0 ] [T1 ] [T2 ] {Pi } [g]

We can write the above equation also as

{Pf } = [T0 ] [T1 ] [T2 ] [T3 ] {P0 } [h]



where [T3 = [T [I], {rQ/B } ], where rQ/B = [ 3 −2 0 ]T and {P0 } = [ 0 0 0 1 ]T .

9.9 Euler Parameters


We studied the Euler angles in previous sections and related angular velocities to the Euler
angles and their rates. As discussed in Section 9.3, these equations are nonlinear and they
have singularities, which make it difficult to integrate them and to do analytical as well as
numerical work with them.
To alleviate these difficulties, it sometimes is preferable to work with another set of ro-
tation variables called Euler parameters. These parameters increase the number of rotation
variables from three to four, but they eliminate the singularities and are more convenient
to use.
500 Applied Dynamics

Because a set of three Euler angles is expressed in terms of four variables, the use of
Euler angles introduces a redundancy. This implies that there is no unique way of expressing
the Euler parameters, and we can easily come up with different sets, such as the Cayley-
Klein parameters, quaternions, and Rodrigues parameters. Rodrigues parameters will be
studied in Section 9.10. The mathematics of the Euler parameters was first introduced by
Sir William Rowan Hamilton in 1843. The vector formulation of Euler parameters, which
is used with quaternions, was developed by Oliver Heaviside.
Euler parameters are inspired by Euler’s theorem. Introduced in Chapter 2, this theorem
states that any rotation of a rigid body about a point can be accomplished by a single
rotation by an angle Φ (principal angle) about a line passing through the center of rotation
(principal line). Euler parameters are a characterization of this rotation.

!# ĭ
"
ș#
ș!
!!
ș"

!"

FIGURE 9.32
Direction cosines and principal line.

Let us describe the principal line in terms of its direction cosines c1 , c2 , c3 . Consider
Figure 9.32. Denoting by θ1 , θ2 , θ3 the angles the principal line makes with the coordinate
axes, the direction cosines are
c1 = cos θ1 c2 = cos θ2 c3 = cos θ3 (9.104)
The Euler parameters are defined as four parameters e0 , e1 , e2 , and e3 such that
Φ Φ Φ Φ
e0 = cos e1 = c1 sin e2 = c2 sin e3 = c3 sin (9.105)
2 2 2 2
Use of the half angle eliminates ambiguities associated with the rotation angle. From the
above equation we conclude that
e20 + e21 + e22 + e23 = 1 (9.106)
This section develops relationships among the Euler parameters, Euler angles, and angu-
lar velocities. The analysis begins by noting that one of the eigenvalues of the transformation
matrix [R] is 1 (the other two eigenvalues are complex conjugates) and the direction cosines
constitute the eigenvector associated with the eigenvalue λ = 1. Denoting the direction
cosine vector by {n} = [c1 c2 c3 ]T , we can write
[R] {n} = {n} (9.107)
Proof of the above relationship can be found in the text by Baruh. Also, left multiplying
Equation (9.107) gives
[R]T {n} = {n} (9.108)
so [R]T also has an eigenvalue λ = 1 and the associated eigenvector is {n}.
Three-Dimensional Kinematics of Rigid Bodies 501

9.9.1 Transformation Matrix in Terms of Euler Parameters


Consider a set of axes a1 a2 a3 which are rotated to obtain the b1 b2 b3 axes. A vector a1 that
is fixed in the A frame becomes b1 upon rotation. This process is identical to the discussion
in Chapter 2, where a vector whose initial position is qi is rotated about the principal line
n by Φ. The rotated vector is given in Equation (2.84). Hence, Equation (2.84) can be
rewritten by replacing qi with a1 and qf with b1 as

b1 = cos Φa1 + (1 − cos Φ) (a1 · n) n + sin Φn × a1 (9.109)

Using the half-angle formulas


Φ Φ Φ Φ
cos Φ = 2 cos2 −1 1 − cos Φ = 2 sin2 sin Φ = 2 sin cos (9.110)
2 2 2 2
and noting that n = c1 a1 + c2 a2 + c3 a3 , we obtain for the components of Equation (9.109)
 
2 Φ
− 1 a1 = 2e20 − 1 a1

cos Φa1 = 2 cos
2

Φ 2
(1 − cos Φ) (a1 · n) n = 2 sin2 c a1 + c1 c2 a2 + c1 c3 a3 = 2 e21 a1 + e1 e2 a2 + e1 e3 a3
 
2 1

Φ Φ
sin Φn × a1 = 2 sin cos (c3 a2 − c2 a3 ) = 2 (e0 e3 a2 − e0 e2 a3 ) (9.111)
2 2
Combining the terms above, we can express b1 as

b1 = 2e20 − 1 + 2e21 a1 + 2 (e1 e2 + e0 e3 ) a2 + 2 (e1 e3 − e0 e2 ) a3



(9.112)

We can repeat the procedure above to obtain expressions for b2 and b3 in terms of
the Euler parameters. Introducing the column vector {e0 } = [e1 e2 e3 ]T , the relationship
between the unit vectors in the A and B frames becomes
   
b1 a1
2e20 − 1 [1] + 2{e0 }{e0 }T − 2e0 [ẽ0 ]  a2 
 
 b2  = (9.113)
b3 a3

where, as discussed in Chapter 2, [ẽ0 ] is the skew-symmetric matrix associated with the
vector {e0 },
 
0 −e3 e2
[ẽ0 ] =  e3 0 −e1  (9.114)
−e2 e1 0

The transformation between the unit vectors is


   
b1 a1
 b2  = [R]  a2  (9.115)
b3 a3

so the transformation matrix can be expressed in terms of the Euler parameters as

2e20 − 1 [1] + 2 {e0 } {e0 }T − 2e0 [ẽ0 ]


 
[R] = (9.116)
502 Applied Dynamics

or
e20 + e21 − e22 − e23
 
2 (e1 e2 + e0 e3 ) 2 (e1 e3 − e0 e2 )
[R] =  2 (e1 e2 − e0 e3 ) e20 − e21 + e22 − e23 2 (e2 e3 + e0 e1 )  (9.117)
 

2 (e1 e3 + e0 e2 ) 2 (e2 e3 − e0 e1 ) e20 − e21 − e22 + e23

The elements of the transformation matrix [R] (or direction cosine matrix [c] = [R]T ) are
quadratic expressions in terms of the Euler parameters. There are no transcendental terms
and no singularities. For any possible orientation of two coordinate frames there exists, and
it is possible to find, the set of associated Euler parameters. By contrast, every Euler angle
transformation sequence has a singularity at which point we cannot uniquely determine the
values of the Euler angles.

9.9.2 Relating the Euler Parameters to Angular Velocities


In Chapter 2, the angular velocity matrix was obtained in terms of the transformation
matrix [R] as
 
0 −ω3 ω2
[W ] = [ω̃] =  ω3 0 −ω1  = [R] [Ṙ]T (9.118)
−ω2 ω1 0

Introducing Equation (9.117) into the above equation allows us to express the components
of the angular velocity in terms of the Euler parameters and their rates. Note that Equation
(9.118) gives the components of the angular velocity in the rotated frame, that is, the B
frame.
For example, the 32 element of the angular velocity matrix, W32 = ω1 , is obtained by
multiplying the third row of [R], which is

[2 (e1 e3 + e0 e2 ) 2 (e2 e3 − e0 e1 ) e20 − e21 − e22 + e23 ]



(9.119)

with the second column of [Ṙ]T (or second row of [Ṙ]), which is

d 
e20 − e21 + e22 − e23
 
2 (e1 e2 − e0 e3 ) 2 (e2 e3 + e0 e1 )
dt

= 2 [(ė1 e2 + e1 ė2 − ė0 e3 − e0 ė3 ) (e0 ė0 − e1 ė1 + e2 ė2 − e3 ė3 )

(ė2 e3 + e2 ė3 + ė0 e1 + e0 ė1 )] (9.120)

Performing the multiplication results in

W32 = ω1 = 2 (−e1 ė0 + e0 ė1 + e3 ė2 − e2 ė3 ) (9.121)

Similar expressions can be developed for ω2 and ω3 . Introducing the vectors {ω ∗ } =


[0 ω1 ω2 ω3 ]T and {e} = [e0 e1 e2 e3 ]T , we can write
    
0 e0 e1 e2 e3 ė0
 ω1   −e1 e 0 e 3 −e 2
  ė1 
 ω2  = 2  −e2 −e3 (9.122)
    
e0 e1   ė2 
ω3 −e3 e2 −e1 e0 ė3
Three-Dimensional Kinematics of Rigid Bodies 503

or, in compact form

{ω ∗ } = 2 [E ∗ ] {ė} (9.123)

where the notation is obvious. We can show that the coefficient matrix [E ∗ ] is orthonormal;
that is, its inverse is equal to its transpose, [E ∗ ]−1 = [E ∗ ]T . Using this property, the
derivatives of the Euler parameters become
−1 T
{ė} = 0.5 [E ∗ ] {ω ∗ } = 0.5 [E ∗ ] {ω ∗ } (9.124)

which are the kinematic differential equations associated with Euler parameters. They can
be rearranged as
    
ė0 0 −ω1 −ω2 −ω3 e0
 ė1   ω1 0 ω3 −ω2    e1 
 
 ė2  = 0.5  ω2 −ω3 (9.125)
  
0 ω1   e2 
ė3 ω3 ω2 −ω1 0 e3
The equations relating the Euler parameters to their derivatives are linear, and they
contain no singularities, making them easy to deal with analytically and numerically. Also,
the matrix above is skew-symmetric.
Noting that the first row of Equation (9.122) is essentially 0 = 0, we can generate a new
matrix [E] of order 3 × 4 by eliminating the first row of [E ∗ ], so that
 
−e1 e0 e3 −e2
[E] =  −e2 −e3 e0 e1  (9.126)
−e3 e2 −e1 e0

Thus, Equation (9.123) can be expressed as

{ω} = 2 [E] {ė} (9.127)

Below are a few interesting properties of matrix [E]:

[E] [E]T = [1] [ω̃] = 2 [E] [Ė]T [E] {e} = {0} [E]T [E] = [1] − {e} {e}T (9.128)

9.9.3 Relating Euler Parameters to the Transformation Matrix


Given a transformation matrix [R], the task of finding the associated Euler parameters is
the inverse of finding the transformation matrix in terms of the Euler parameters. As it is
an inversion, the task can be accomplished in a number of ways. One way is to note from
Equation (9.117) that

R11 + R22 + R33 = 3e20 − e21 − e22 − e23 = 4e20 − 1 (9.129)

and to solve for e0 as


1p
e0 = R11 + R22 + R33 + 1 (9.130)
2
Another way is to note that the trace (sum of the diagonals) of [R] is invariant for
different representations. In terms of the principal angle Φ, [R] has the form
 
1 0 0
[R] =  0 cos Φ sin Φ  (9.131)
 

0 − sin Φ cos Φ
504 Applied Dynamics

so R11 + R22 + R33 = 1 + 2 cos Φ, from which we calculate cos Φ as


1
cos Φ = (R11 + R22 + R33 − 1) (9.132)
2
After obtaining e0 , we can evaluate the elements of Equation (9.117) to find the remain-
ing Euler parameters. For example,

R23 − R32 = 4e0 e1 (9.133)

which can be solved to find e1 . To find the remaining Euler parameters, we can use
R23 − R32 R31 − R13 R12 − R21
e1 = e2 = e3 = (9.134)
4e0 4e0 4e0
The Euler parameters can be used to find the direction cosines c1 , c2 , c3 . For example,
Φ Φ
R23 − R32 = 4e0 e1 = 4c1 cos sin = 2c1 sin Φ (9.135)
2 2
which can be solved for c1 . Similarly, for the other direction cosines
R23 − R32 R31 − R13 R12 − R21
c1 = c2 = c3 = (9.136)
2 sin Φ 2 sin Φ 2 sin Φ
An alternate way of finding e1 , e2 , and e3 is to obtain the eigensolution of [R] and
ascertain the direction cosines c1 , c2 , c3 as the eigenvector corresponding to the eigenvalue
whose value is 1.

9.9.4 Relating Euler Parameters to the Euler Angles


This task can be accomplished by generating the direction cosine matrix for a specific Euler
angle transformation and using the relationships in the previous subsection. The results
vary from one set of transformations to another. We dispense with the algebra and list in
Table 9.3 the results for common transformations.

TABLE 9.3
Euler parameters in terms of the Euler angles

Sequence 3-1-3 3-2-3 3-2-1


Angles φ, θ, ψ ψ, θ, φ ψ, θ, φ
e0 cos φ+ψ θ
2 cos 2 cos φ+ψ θ
2 cos 2 cos ψ2 cos θ2 cos φ2 + sin ψ2 sin θ2 sin φ2
e1 cos φ−ψ θ
2 sin 2 sin φ−ψ θ
2 sin 2 cos ψ2 cos θ2 sin φ2 − sin ψ2 sin θ2 cos φ2
e2 sin φ−ψ θ
2 sin 2 cos φ−ψ θ
2 sin 2 cos ψ2 sin θ2 cos φ2 + sin ψ2 cos θ2 sin φ2
e3 sin φ+ψ θ
2 cos 2 sin φ+ψ θ
2 cos 2 sin ψ2 cos θ2 cos φ2 − cos ψ2 sin θ2 sin φ2

To determine the Euler angles from the Euler parameters is a more tedious process, and
since it is an inverse operation, it can be accomplished in a number of ways. One way is
to consider the transformation matrix and its entries. Another is to consider Table 9.3 and
Three-Dimensional Kinematics of Rigid Bodies 505

work backwards. For example, for a 3-1-3 sequence, we can manipulate the Euler parameters
as
e3 φ+ψ e2 φ−ψ
= tan = tan (9.137)
e0 2 e1 2
After taking the inverse tangents of the above terms and manipulating, the angles φ and ψ
are found from
       
e3 e2 e3 e2
φ = tan−1 + tan−1 ψ = tan−1 − tan−1 (9.138)
e0 e1 e0 e1

We can then calculate θ using any of the Euler angles in Table 9.3.
To determine the exact values of these angles, that is, the quadrants they lie in, we can
use Table 9.3 or compare the results with the rotation matrix.

Example 9.9
An inertial reference frame is rotated using a 3-2-3 Euler angle transformation with angles
ψ = 15◦ , θ = −28◦ , and φ = 68◦ . Find the direction cosines of the principal line and the
principal rotation Φ. Then, calculate the associated Euler parameters. Then, considering
the rates of the Euler angles ψ̇ = 0.1 rad/s, θ̇ = −0.15 rad/s, and φ̇ = 1.3 rad/s, calculate
the derivatives of the Euler parameters.
We begin by calculating the rotation matrices. They are, from Equation (9.31),
   
0.9659 0.2588 0 0.8829 0 0.4695
[R1 ] =  −0.2588 0.9659 0  [R2 ] =  0 1 0 
0 0 1 −0.4695 0 0.8829
 
0.3746 0.9272 0
[R3 ] =  −0.9272 0.3746 0 [a]
0 0 1
and the combined rotation matrix becomes
 
0.0795 0.9812 0.1759
[R] = [R3 ][R2 ][R1 ] =  −0.8877 0.1500 −0.4353  [b]
−0.4535 −0.1215 0.8829

We solve for the eigenvalues of [R] next. The eigenvector corresponding to the eigenvalue
 T
λ = 1, {n} = [c1 c2 c3 ]T , is found to be {n} = −0.1571 −0.3152 0.9359 , and it gives
the direction cosines of the principal line. Using Equation (9.132), we calculate the principal
angle as  
1
Φ = cos−1 (R11 + R22 + R33 − 1) = 1.5146 rad [c]
2
The Euler parameters are calculated using Equation (9.130) and Equation (9.134) as

1p R23 − R32
e0 = R11 + R22 + R33 + 1 = 0.7267 e1 = = −0.1079
2 4e0

R31 − R13 R12 − R21


e2 = = −0.2165 e3 = = 0.6429 [d]
4e0 4e0
506 Applied Dynamics

The next step is to calculate the derivatives of the Euler angles. To this end, we need to
calculate the angular velocities using Equation (9.39) and the [B] matrix, which gives
 
ω1
  
n o −sθcφ sφ 0 ψ̇
{ω} =  ω2  = [B] φ̇ =  sθsφ cφ 0   θ̇ 
 
ω3 cθ 0 1 φ̇
    
0.1759 0.9272 0 0.1 −0.1215
=  −0.4353 0.3746 0   −0.15  =  −0.0997  [e]
0.8829 0 1 1.3 1.3883
h iT
in which the vector {φ̇} = ψ̇ θ̇ φ̇ .
Invoking Equation (9.125) we find the rates of change of the Euler parameters as
    
0 −ω1 −ω2 −ω3 e0 −0.4636
 ω1 0 ω3 −ω2   e1   −0.1624 
{ė} = 0.5 
 ω2 −ω3   e2  =  −0.0004 
    [f ]
0 ω1
ω3 ω2 −ω1 0 e3 0.4967

9.10 Rodrigues Parameters


The Rodrigues parameters are a variant of the Euler parameters. They are frequently used
with screw algebra and advanced kinematic analysis. They are named after the French
mathematician Olinde Rodrigues (1795–1851), who was known more for his writings on
banking and social reform. His publication of the Rodrigues parameters in 1840 largely
went unnoticed, even though this publication was prior to Hamilton’s 1843 publication on
the Euler parameters.
The Rodrigues parameters bi (i = 1, 2, 3) are defined as
e1 Φ e2 Φ e3 Φ
b1 = = c1 tan b2 = = c2 tan b3 = = c3 tan (9.139)
e0 2 e0 2 e0 2
The Rodrigues parameters encounter singularity problems when tan (Φ/2) is not defined,
as when Φ = 180◦ . This problem can be alleviated by switching to a different set of variables
in the vicinity of the singularity. The singularity associated with Rodrigues parameters is
encountered much less frequently than Euler angle singularities.
The Rodrigues vector {b} is defined as
 
b1  
Φ
{b} =  b2  = {n} tan (9.140)
2
b3
The Rodrigues parameters exhibit some interesting properties. For example,
Φ 1 1
b21 + b22 + b23 + 1 = tan2 +1 = = (9.141)
2 cos2 Φ
2
e20
We can divide and multiply the expression for the transformation matrix [R] by e20 and
express [R] in Equation (9.117) terms of the Rodrigues parameters as
 
1 + b21 − b22 − b23 2 (b1 b2 + b3 ) 2 (b1 b3 − b2 )
1
[R] =  2 (b1 b2 − b3 ) 1 − b21 + b22 − b23 2 (b2 b3 + b1 )  (9.142)
 
2 2 2
1 + b1 + b2 + b3
2 (b1 b3 + b2 ) 2 (b2 b3 − b1 ) 1 − b21 − b22 + b23
Three-Dimensional Kinematics of Rigid Bodies 507

Defining the screw-symmetric matrix associated with the direction cosines as [ñ], we
can show that the rotation matrix [R], which was defined in Equation (2.92), can also be
expressed as

[R] = [1] − sin Φ [ñ] + (1 − cos Φ) [ñ] [ñ] (9.143)

An important relationship in kinematics, known as Cayley’s formula, states that the


transformation matrix [R] can also be expressed in terms of the Rodrigues parameters as
 −1  
[R]T = [c] = [1] − [b̃] [1] + [b̃] (9.144)

where [b̃] is the skew-symmetric matrix associated with the Rodrigues parameters. All the
relationships derived in the previous section involving Euler parameters can also be ex-
pressed in terms of the Rodrigues parameters.
A very useful relationship related to the Rodrigues parameters is the Rodrigues equation.
Consider an initial vector ri , which is rotated by angle Φ about the principal line n. The
final orientation of the vector is rf . The two vectors are related by

{rf } = [R]T {ri } (9.145)

where [R] can be calculated by any of the methods discussed above. Stated here without
derivation, the Rodrigues equation relates the initial and final vectors to the principal line
and rotation angle as
 
Φ
rf − ri = tan n × (rf + ri ) (9.146)
2

Derivation of the Rodrigues equation can be found in advanced texts on kinematics.


Similar to the previous section, we can obtain the relationships between the rates of
the Rodrigues parameters and the angular velocities. Stated here without derivation, these
relationships are
2    
{ω} = [1] − [b̃] {ḃ} {ḃ} = 0.5 [1] + [b̃] + {b}{b}T {ω} (9.147)
1 + {b}T {b}

The second expression above describes the kinematic differential equations associated with
the Rodrigues parameters.
Table 9.4 compares the Euler angles, Euler parameters and Rodrigues parameters re-
garding their complexity, ease of use and singularities.

TABLE 9.4
Comparison of Euler angles, Euler parameters and Rodrigues parameters

Euler Angles Euler Par. Rodrigues Par.


Number of variables 3 4 3
Choice of variables Yes No No
Are variables independent? Yes No Yes
Singularities Relatively frequent None Infrequent
Kinematic diff. equations Nonlinear Linear Nonlinear
508 Applied Dynamics

9.10.1 Screw Algebra and Rodrigues Parameters


Chasles’ theorem, discussed in Section 9.2.3, states that the most general transformation
of a body can be described as the translation of some point on the body plus a rotation
about an axis that goes through that point. While the orientation of the principal line and
amount of rotation are unique, location of the translation point on the body and location
of the rotation axis are not unique.
One way of specifying the most general motion of a body is by means of a screw trans-
formation, where the body is rotated about the principal line, also known as the screw axis,
by the principal angle Φ, and translated by a distance ρ, known as the slide, along the screw
axis. The motion of the body resembles the motion of a screw, where the body gets rotated
about the screw axis and it also moves up or down along the screw axis. The screw axis is
depicted in Figure 9.33.

"
(

!) )
! $
!)ү
'
&

# %

FIGURE 9.33
Screw axis and points Q0 and Q.

Note that in application of Euler’s theorem, that is, the rotation of a body with one
point fixed, the principal line goes through the fixed point. As there is no such fixed point
in general motion, the precise location of the screw axis has to be defined. This is done by
specifying the coordinates of a point along the screw axis.
The following questions are of interest:

• If a vector {ri } is translated by a distance {d} and rotated by [R] so that the rotated
vector is {rf } = [R]T {ri } + {d}, how do we locate the screw axis and calculate the slide?
• Given a reference point on the screw axis, the orientation of the screw axis, the rotation
amount, and the amount of slide, what is the final orientation of a body after the screw
operation?

When addressing the first question, first the direction cosines are found using one of the
relationships in Section 9.9.3, such as the eigenvector associated with the eigenvalue λ = 1
of [R]. Then, the principal angle Φ is calculated.
To calculate the coordinates of a point on the screw axis and the amount of slide, consider
the position of an arbitrary point on the screw axis, say, Q0 . Also consider a reference point
O from which the motion is observed. Without loss of generality, the screw axis goes through
Three-Dimensional Kinematics of Rigid Bodies 509

the xy plane at point (a, b, 0), as shown in Figure 9.33. It follows that we can express the
position of point Q0 as

rQ0 = ai + bj + tn = ai + bj + t (nx i + ny j + nz k) (9.148)

where t is unknown and nx = c1 , ny = c2 , nz = c3 denote the direction cosines of the


screw axis, calculated using one of the approaches in Section 9.9.3. Next, consider the dot
product between the unit vector n along the screw axis and rQ0 , which gives

rQ0 · n = ((a + tnx ) i + (b + tny ) j + tnz k) · (nx i + ny j + nz k) = anx + bny + t(9.149)

The interest is in finding point Q such that rQ · n = 0. Setting anx + bny + t0 = 0, we


solve for t0 as t0 = −anx − bny . This point, of course, is the location of the closest point
from the reference point O to the screw axis, as shown in Figure 9.33.

n
n
P
d

P
d*
P
ri r
rf

O
d
O

FIGURE 9.34
Displacement {d} and its components {d∗ } and ρ{n}.

Figure 9.34 shows the transformation from ri to rf . First, ri is rotated about the prin-
cipal line n going through point O by Φ into r0 . This vector is then translated by d to point
P 00 . In Figure 9.34 the points OO00 P 00 P form a parallelogram. Another way to achieve this
rotation is to move ri by d (both point O and point P are moved) and then to perform the
rotation about an axis aligned with n that goes through O00 . Figure 9.35 depicts the screw
transformation. First, point Q is located. The screw axis goes through this point. Then, the
rotation is carried out. Point O moves to O2 .
It is easier to visualize the rotation of point P about the screw axis by considering point
Q0 on the screw axis, which is the closest point to P . The rotation by Φ moves point P to
P2 . Hence, line OP after rotation has moved to O2 P2 . Then, the slide ρ is added along the
principal line, so that O2 moves to O00 and P2 moves to P 00 .
Point S is on the plane that is perpendicular to the screw axis and point P lies on this
plane. Hence, points Q0 , P , S, P 0 , S 00 , and P2 lie on the same plane. The distance from O
to S is the same as the distance from Q to Q0 , and the lines OQ and SQ0 are parallel and
have the same length. Hence, OQQ0 S form a rectangle.
To find the amount of slide, it is helpful to separate the displacement {d} = {rf } −
510 Applied Dynamics


!
! !
ĭ 3үү
! ȡ
% 6үү
& ĭ
3ү &!

"#
"

'
2үү
ĭ !
"' ȡ
$!
$

"#

"#$%&'()*+

FIGURE 9.35
The screw transformation.

[R]T {ri } into its components along the screw axis {n} and perpendicular to it as

{d} = {d∗ } + ρ{n} (9.150)

where ρ = {n}T {d} and {n}T {d∗ } = 0. Hence, the slide is found.
Next, consider the calculation of point Q. Since the Rodrigues vector is {b} =
{n} tan Φ/2, it follows that {b}T {d∗ } = 0 as well. Because Q is on the screw axis, it
should not move after a rotation is performed. In column vector notation and noting from
{rf } = [R]T {ri } + {d}, the initial and final positions of rQ are the same and

{Q} = [R]T {Q} (9.151)

where {Q} is the column vector representation of rQ , leading to the relationship


[1] − [R]T {Q} = {0}. Since an eigenvalue of [R] (and of [R]T ) is 1, the matrix [1] − [R]T ,
where [1] is the identity matrix, has a zero eigenvalue
 and hence is a singular matrix. This
means that no solution to the equation [1] − [R]T {Q} = {d} should exist for arbitrary
{d}. However, it is possible to find rQ that is the solution of the equation

[1] − [R]T {Q} = {d∗ }



(9.152)

The above equation has a solution because of the special nature of {d∗ }. To see this, left
Three-Dimensional Kinematics of Rigid Bodies 511

multiply Equation (9.152) by {n}T and note that ([1] − [R]) {n} = {0} and that {b}T {d∗ } =
{0}.
Next, introduce Cayley’s formula in Equation (9.144) into Equation (9.152) so that
  −1  
[1] − [1] − [b̃] [1] + [b̃] {Q} = {d∗ } (9.153)

 
Left multiplying by [1] − [b̃] and collecting terms yields
 
−2[b̃]{Q} = [1] − [b̃] {d∗ } (9.154)

Equation (9.154) can be manipulated more easily by switching to algebraic vector nota-
tion

−2b × rQ = d∗ − b × d∗ (9.155)

where the notation is obvious. Left cross-multiplying by the Rodrigues vector b gives

−2b × (b × rQ ) = b × (d∗ − b × d∗ ) (9.156)

Using the identity e × (f × rQ ) = f (e · rQ ) − rQ (e · f ) and noting that here e = f = b, the


left side of the above equation can be written as

b × (b × rQ ) = b(b · rQ ) − rQ (b · b) (9.157)

The first term on the right side of the above equation, b(b·rQ ), is zero, as b = n tan(Φ/2)
and Q was defined such that rQ · n = 0. This result leads to an expression for the point on
the screw axis as
b × (d∗ − b × d∗ )
rQ = (9.158)
2b · b
Since b is parallel to n, we can write

b × d∗ = b × (d∗ + ρn) = b × d (9.159)

so that Equation (9.158) can be written as

b × (d − b × d)
rQ = (9.160)
2b · b
Note that the above equation gives the location of point Q, which is the closest point
to O on the screw axis. Therefore, any other point with coordinates rQ + tn, where t is a
slide parameter, also lies on the screw axis.
To summarize, the procedure to be followed when calculating the screw parameters in
terms of [R] and {d} is as follows:
• Calculate the direction cosines {n} using relationships from Section 9.9.3.
• Calculate the rotation angle Φ using relationships from the same section, and then
calculate the Rodrigues parameters {b} = {n} tan(Φ/2).
• Calculate the slide ρ using the relationship ρ = {n}T {d}, and calculate {d∗ } = {d} −
ρ{n}.
512 Applied Dynamics

• Calculate the coordinates of point Q on the screw axis using Equation (9.158) or Equa-
tion (9.160).
Consider now the second question; that is, given an initial vector {ri }, the screw axis,
rotation angle, and slide, calculate the position of the rotated vector. To this end, we can
use Equation (9.152), from which the displacement {d} is expressed as

{d} = {d∗ } + ρ{n} = [1] − [R]T {Q} + ρ{n}



(9.161)

Considering the transformation equation {rf } = [R]T {ri } + {d} and introducing to it the
expression for {d} from Equation (9.161), the final position becomes

{rf } − {Q} = [R]T ({ri } − {Q}) + ρ{n} (9.162)

Note that the vector {Q} in the above equation can be the coordinate of any point on
the screw axis. The rotation vector [R] is calculated using one of Equation (2.92), Equation
(9.143), or Equation (9.144).

9.10.2 Rotation of a Body


The previous section discussed rotation of a vector measured from a fixed point. Here, we
consider the screw transformation of a body by considering the transformation of two points
on the body. As we considered transformation of P to P 00 earlier, let us consider the screw
transformation of line OP . To this end, it is necessary to rotate point O about the screw
axis and translate it along the screw axis, which leads to point O00 .
The configuration is shown in Figure 9.35. Rotation of O to O2 is achieved by rotating
O about the screw axis by Φ. Then, the slide ρ is added. The resulting point is O00 . The
line OP has thus been transformed into O00 P 00 . Of course, rO00 /O = d. As before, OO00 P 00 P 0
form a parallelogram. Also, OO2 S 00 S form a rectangle and so does QO2 S 00 Q0 .

Example 9.10
Given an XY Z coordinate system, a body is first rotated about the Z axis by an angle of
θ1 = π/10 radians to yield the X 0 Y 0 Z 0 coordinates. Next, the X 0 Y 0 Z 0 axes are rotated by
an angle of π/8 about the Y 0 axis. Then, the body is translated by d = −3I + 5J − K. Find
the direction cosines of the screw axis, the rotation angle Φ, the slide ρ along the screw
axis, and coordinates of point Q on the slide axis, so that rQ · n = 0.
For the first rotation, which is a 3 rotation, the rotation angle is θ1 = 0.3142 rad. The
second rotation is a 2 rotation and angle is θ2 = 0.3927 rad. The corresponding transfor-
mation matrices are
   
0.9511 0.3090 0 0.9239 0 −0.3827
[R1 ] =  −0.3090 0.9511 0  [R2 ] =  0 1 0  [a]
0 0 1 0.3827 0 0.9239
and the combined rotation matrix is
 
0.8787 0.2855 −0.3827
[R] = [R2 ][R1 ] =  −0.3090 0.9511 0  [b]
0.3640 0.1183 0.9239

The eigensolution of [R] yields three eigenvalues. Two are complex conjugates and the
third one is 1. The corresponding eigenvector is the direction cosines of the screw axis
 T
{n} = −0.1230 0.7764 0.6182 [c]
Three-Dimensional Kinematics of Rigid Bodies 513

The rotation angle is obtained using


 
−1 R11 + R22 + R33 − 1
Φ = cos = cos−1 (0.8789) = 0.5016 rad [d]
2
which is equal to 28.73◦ .
The Rodrigues vector {b} is calculated using
Φ T
{b} = [b1 b2 b3 ]T = {n} tan

= −0.0315 0.1989 0.1584 [e]
2
The slide ρ and {d∗ } vectors are
 
−3
ρ = {n}T {d} = −0.1230 0.7764 0.6182  5  = 3.6325
 
[f ]
−1
     
−3 −0.1230 −2.5533
{d∗ } = {d} − ρ{n} =  5  − 3.6325  0.7764  =  2.1799  [g]
−1 0.6182 −3.2455
Finally, the coordinates of point Q on the screw axis are calculated using Equation
(9.158), which yields
 
−8.8237
[b̃]  
{rQ } = {d∗ } − [b̃]{d∗ } =  −2.7692  [h]
2{b}T {b}
1.7726

9.10.3 Finding the Screw Axis and Rotation Angle


The previous sections discussed obtaining the screw axis given the rotation and translation
transformations of an initial vector. Here, we discuss a different problem. Suppose we have
the initial and final positions of a vector. How many such vector pairs are needed to calculate
the principal line, rotation angle, and slide?
The solution is much easier when there is only a rotation transformation. For a simple
rotation transformation, the initial and final positions of two sets of vectors are sufficient
to identify the principal line. Consider two sets of vectors, {ri }, {rf } and {qi }, {qf }. The
initial and final positions are related by

{rf } = [R]T {ri } {qf } = [R]T {qi } (9.163)

Because the final positions of the vectors are obtained by rotating them about the
principal line, the angle that the initial and final vectors make with the principal line
remains unchanged after the rotations, as shown in Figure 9.36. Hence, using algebraic
vector notation, we can write

n · rf = n · ri n · qf = n · qi (9.164)

which can be rearranged as

n · (rf − ri ) = 0 n · (qf − qi ) = 0 (9.165)

The vectors (rf − ri ) and (qf − qi ) lie on planes that are perpendicular to the principal
line. These vectors are not necessarily on the same plane, but the planes they are on are
514 Applied Dynamics

rf - ri

qf - qi rf
ri

qi qf

FIGURE 9.36
Two vectors rotated about principal line.

parallel to each other. The cross product of these two vectors is perpendicular to the planes
and therefore along the principal line. Normalizing the cross product gives the unit vector
n along the principal line as
(rf − ri ) × (qf − qi )
n = (9.166)
|(rf − ri ) × (qf − qi )|

The rotation angle Φ can be found using the Rodrigues equation in Equation (9.146) or
by means of other procedures outlined in Section 9.9.3.
Next, consider transformations that include rotations and translations and the number
of initial and final position vector pairs that are needed to locate the orientation of the
principal line, coordinates of point Q on the screw axis, and the slide ρ. Begin with three
vector pairs, say, {q}, {r}, and {s}. Equation (9.162) can be written for each vector as

{qf } − {Q} = [R]T ({qi } − {Q}) + ρ{n}

{rf } − {Q} = [R]T ({ri } − {Q}) + ρ{n}

{sf } − {Q} = [R]T ({si } − {Q}) + ρ{n} (9.167)

Define two vectors, {x} = {q} − {r} and {y} = {q} − {s}. Relating the initial and final
positions of these two vectors is achieved by substituting from Equation (9.167), which gives

{xf } = [R]T {xi } {yf } = [R]T {yi } (9.168)

It follows that we can use the vectors {xi }, {xf }, {yi }, and {yf } in Equation (9.166) to find
the principal line {n} and Equation (9.146) to calculate the rotation angle Φ. This enables
calculation of the rotation matrix [R].
Three-Dimensional Kinematics of Rigid Bodies 515

Now that the rotation matrix is known, the next steps are to calculate the exact location
of the screw axis and the slide. Recalling that {Q}T {n} = rQ · n = 0, left multiplying any
one of Equation (9.167) by {n}T gives, say, for {r},

{n}T {rf } − [R]T {ri } = ρ



(9.169)

which provides the value of the slide. Using the relation {rf } = [R]T {ri } + {d}, we calculate
{d} as

{d} = {rf } − [R]T {ri } (9.170)

Once {d} is calculated, we invoke Equation (9.160) to calculate rQ .

Example 9.115
A body is transformed from one location to another. Using an xyz coordinate system, the
initial and final coordinates of four points on the body, A, B, C, and D, are given in
Table 9.5. Calculate orientation of the principal line, coordinates of the point where the
principal line crosses the yz plane, the rotation angle Φ, and the slide ρ.

TABLE 9.5
Initial and final coordinates of four points on the body

Point Initial Coordinates Final Position


A (0, 3, 7) (1.90, 11.23, 7.19)
B (2, 7, 10) (3.29, 14.44, 11.29)
C (0, 5, 10) (4.26, 13.41, 8.84)
D (−2, 5, 7) (3.92, 9.83, 8.59)

Based on the developments of the previous section, we need the coordinates of only three
points on the body. Let us select them as A, B, and D, with rA = r, rB = q, and rD = s.
The initial and final values of the difference vectors, x = q − r and y = q − s, become
 T  T
{xi } = 2 4 3 {xf } = 1.39 3.21 4.10

 T  T
{yi } = 4 2 3 {yf } = −0.63 4.61 2.70 [a]
From Equation (9.166), the normalized cross product of ({xf } − {xi }) and ({yf } − {yi })
gives the unit vector along the screw axis as

(xf − xi ) × (yf − yi )
n = = 0.3336i + 0.6683j + 0.6649k [b]
|(xf − xi ) × (yf − yi )|

The next step is to calculate the rotation angle using Equation (9.146). To this end, we
can use x or y. Let us use y. Denoting by z = yf − yi and z0 = n × (yf + yi ), we can write
Equation (9.146) as  
Φ
z = tan z0 [c]
2
5 The data in this example is the same as Examples 6.9 and 6.10 in Advanced Mechanism Design, by

Sandor and Erdman.


516 Applied Dynamics

which indicates that we can solve for Φ three ways, namely by relating the first, second,
or third elements in Equation [c]. For example, z · i = tan (Φ/2) z0 · i. Doing so gives three
values for Φ = 2.8897, 2.8831, and 2.8322 radians.
Ideally, when the initial and final coordinates are known accurately, all three values
obtained for Φ should be the same. These values should also match the values of Φ obtained
by using x (instead of y). We assume here that there is rounding error or other inaccuracy
in the given data and proceed by taking the average of the three values for Φ. The result is
Φ ≈ 2.8683 rad, which corresponds to 164.344◦ .
Using the average value for Φ, we calculate the Rodrigues vector b
 
Φ
b = n tan = 2.4267i + 4.8607j + 4.8365k [d]
2
The rotation matrix [R] can be calculated using Equation (9.142), Equation (9.143), or
Equation (9.144) and has the value
 
−0.7444 0.6171 0.2551
[R] =  0.2582 −0.0863 0.9622  [e]
0.6158 0.7822 −0.0950
The slide is calculated using Equation (9.169). We have the choice of using r or q or
s. This feature can also be used to check the accuracy of the results, by calculating three
values for the slide and comparing them. Ideally, all three values should be identical. Let
us use s, which gives for the slide
 
T T
ρ = {n} {sf } − [R] {si } = 6.2599 [f ]

The vector {d} is calculated from Equation (9.170) (again, we have the three choices of
using r, q, or s, let us use r)
T
{d} = {rf } − [R]T {ri } = −3.1849 6.0138 4.9685

[g]
.
The next step is to find rQ , the coordinates of closest point on the principal line to r.
Using Equation (9.160) we obtain
b × (d − b × d)
rQ = = −2.6833i + 0.6558j + 0.6873k [h]
2b · b
The calculation of rQ completely defines the screw axis. We calculated its orientation in
Equation [b] and the coordinates of a point through which it passes in Equation [h]. The
problem asks us to find the coordinates for the point at which the screw axis crosses the yz
plane, that is, when x = 0. To this end, from Equation (9.149), we can write the position
of any point (denoted by u) on the screw axis as
u = rQ + tn [i]
in which t is a slide parameter along the screw axis. We want the x component of this
equation to be zero, which is calculated by
u · i = 0 = (rQ + tn) · i [j]
Solving for t we obtain t = 8.0431, and the position vector where the screw axis crosses
the yz plane becomes
u = rQ + 8.0431n = 0i + 6.0306j + 6.0354k [k]
Three-Dimensional Kinematics of Rigid Bodies 517

Example 6.9 in Advanced Mechanism Design indicates that the given data for the initial
and final positions correspond to a rotation by 165◦ about a screw axis with unit vector
n = 1/3i + 2/3j + 2/3k, which crosses the yz plane at (0, 6, 6). The slide is ρ = 6.25.
Calculation of the final positions of A, B, C, and D for these parameters is left as an
exercise (see Problem 9.51).
If you solve Problem 9.51, you will see that the errors encountered in this example are
due to roundoff, which indicates that there is numerical sensitivity in the process outlined
above. Such sensitivity is encountered in a large class of inverse kinematics problems.

9.11 Bibliography
Altmann, S.L., Rotations, Quaternions and Double Groups, Dover, 2005.
Baruh, H., Analytical Dynamics, McGraw-Hill, 1998.
Greenberg, M.D., Advanced Engineering Mathematics, 2nd Edition, Prentice-Hall, 1998.
Kane, T.R., Likins, P.W., and Levinson, D.A., Spacecraft Dynamics, McGraw-Hill, 1984.
Kane, T.R. and Scher, M.P., “A Dynamical Explanation of the Falling Cat Phenomenon,”
International Journal of Solids and Structures, Vol. 5, No. 5, pp. 663–670, 1969.
McCarthy, J.M. and Soh, G.M., Geometric Design of Linkages, 2nd Edition, Springer, 2011.
Sandor, G.N. and Erdman, A.G., Advanced Mechanism Design, Vol. 2. Prentice-Hall, 1984.

9.12 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 9.2—Basic Kinematics of Rigid Bodies

A
1

a
X

B
2 v
b
Z s C

FIGURE 9.37
Figure for Problems 9.1 and 9.2.
518 Applied Dynamics

9.1 (E) Collar A in Figure 9.37 rotates around a stationary thin rod with angular velocity
ω1 . Collar B rotates about a rod attached to A with ω2 . Calculate the angular velocity and
angular acceleration of collar B at the position shown (ω1 and ω2 are not constant).
9.2 (M) Consider Figure 9.37 and the parameters a = 11 in., b = 6 in., s = 8 in., ω1 = 0.5
rad/sec and constant, ω2 = −2 rad/sec and constant. A thin rod protrudes out of collar C
at the constant rate of ṡ = 3 in./sec. Calculate the velocity and acceleration of point C.
9.3 (M) Consider the rotating rod in Figure 9.3. The collar is executing a screw motion
on the rod, advancing by a distance L every time the collar rotates by one revolution. At
the given instant, ω1 = 0.2 rad/sec, ω3 = 0.4 rad/sec, β = 30◦ and constant, and y = 4L.
Calculate the velocity of the collar.
9.4 (M) Consider the rotating rod in Figure 9.3. The collar rotates around the rod. At
the given instant, ω1 = 0.2 rad/sec constant and ω3 = 0.4 rad/sec and it also is constant.
β = 30◦ and increasing with β̇ = 0.1 rad/s and β̈ = −0.3 rad/sec2 . Calculate the angular
acceleration of the collar.

Section 9.3—Euler Angles


9.5 (E) The body-fixed axes xyz of a body are obtained after a 3-2-1 transformation (ψ, θ, φ).
At a certain instant, the following measurements are taken: ψ = 30◦ , θ = 10◦ , φ = −25◦ ,
ωx = 0.3 rad/s, ωy = −0.2 rad/s, ωz = 1.5 rad/s. Calculate the derivatives of the Euler
angles at this instant.
9.6 (E) A coordinate system XY Z is rotated using a 2-1-3 Euler angle sequence into axes
xyz by the angles φ, θ, ψ. Calculate the resulting rotation matrix [R]. Then, obtain the
expression for the angular velocity.
9.7 (E) A coordinate system a1 a2 a3 is rotated using a 1-3-1 Euler angle sequence into axes
b1 b2 b3 by the angles φ, θ, ψ. Calculate the resulting rotation matrix [R]. Then, obtain the
expression for the angular velocity in terms of the Euler angles.

b3 Top view

r b2 r
b2
m G m
G R

b1 b1

FIGURE 9.38
Figure for Problem 9.8.

9.8 (M) The body-fixed axes of the spacecraft in Figure 9.38 are obtained after a 3-1-3 Euler
angle transformation (φ, θ, ψ). Inside the spacecraft is a tube, and a mass slides inside the
tube. Calculate the velocity of the sliding mass in the inertial a1 a2 a3 as well as body-fixed
b1 b2 b3 frames.
9.9 (M) Consider a 3-1-3 transformation (φ, θ, ψ). At a certain instant φ = 15◦ , θ = 30◦ , and
ψ = −45◦ . The body angular velocities are ω1 = 0.1 rad/s, ω2 = 0.2 rad/s, and ω3 = −0.3
rad/s. Calculate the values of the time derivatives of the Euler angles.
Three-Dimensional Kinematics of Rigid Bodies 519

9.10 (C) Consider a 3-2-3 transformation with angles ψ, θ, and φ. At a certain instant
ψ = 15◦ , θ = 2◦ , and φ = −60◦ . The body angular velocities are ω1 = 0.1 rad/s, ω2 = 0.2
rad/s, and ω3 known approximately, as ω3 = −0.3 ± 0.05 rad/s. Calculate the values of
the time derivatives of the Euler angles and plot them for the different possible values of
ω3 , that is, in the range −0.35 ≤ ω3 ≤ −0.25 rad/s. Repeat the same analysis for when
θ = 45◦ . Comment on the variation of the values that you get.

X
A
Y B x
Z

y L

FIGURE 9.39
Figure for Problem 9.11.

9.11 (M) The car in Figure 9.39 is moving with the following kinematical information:
Velocity of point A, vA = 66i ft/sec, angular velocity ω = 0.1i + 0.05k rad/sec, heading
angle ψ = 10◦ , attitude θ = 2◦ , and roll angle φ = 5◦ . We are using a 3-2-1 Euler angle
transformation with the inertial XY Z axes selected such that the Z axis is pointing down
and the X axis is pointing forward. Find the velocity of point B. Hint: Take advantage of
the smallness of the Euler angles.
9.12 (C) The body-fixed axes xyz of a body are obtained after a 3-2-1 transformation
(ψ, θ, φ). The body is initially at rest, when forces and moments are applied to it and the
ensuing angular velocities have the following time histories: ωx (t) = 0.1t rad/s, ωy (t) =
−0.05t rad/s, ωz (t) = 0.2 sin 2t rad/s. Derive the kinematic differential equations6 for the
3-2-1 Euler angle sequence and numerically integrate them to calculate the time histories
of the Euler angles.

Section 9.4—Axisymmetric Bodies


9.13 (E) As discussed in Chapter 4, a wheel with caster is a very useful way of moving an
object. Consider the caster in Figure 9.40a and obtain the angular acceleration of the wheel
when ω1 = 0.2 rad/s and ω2 = 10 rad/s. Also, calculate the velocity of point C on the
caster with respect to point A.
9.14 (M) Consider the spinning top in Example 9.3. Obtain expressions for the angular
velocity, angular acceleration, and acceleration of the center of mass using a 3-2-3 transfor-
mation. Show, by comparing magnitudes of the velocity and acceleration terms, that the
results you obtain are the same as the results in Example 9.3.
9.15 (M) Consider the frisbee in Figure 9.40b and obtain expressions for the angular velocity
and angular acceleration using a 3-1-3 Euler angle transformation (rotation angles φ, θ, and
ψ) and the shape frame.
6 Equation (9.40) gives the kinematic differential equations for a 3-2-3 Euler angle sequence.
520 Applied Dynamics

a) b)
f3

f3 a3, a3
1

A R=12 cm

f2
f1 f2

2
15 cm
C
B f1
3 cm

FIGURE 9.40
Figures for a) Problems 9.13 and 9.16, b) Problem 9.15.

Section 9.5—Rolling
9.16 (E) The caster in problem 9.13 is now rolling on a surface with no slip. Calculate the
velocity of point A and the acceleration of point A. Both angular velocities ω1 and ω2 are
constant.
9.17 (E) A disk of radius 12 in. rolls with steady precession with nutation angle with the
vertical of 15◦ and it completes a circle every 10 seconds. Calculate the radius of curvature
and speed of the center of the disk.
9.18 (M) The small cone in Figure 9.41a rolls over the large stationary cone and makes a
trip every five seconds. Find the angular velocity and angular acceleration of the small cone.
Both cones are right-angled.

a) b)

R
3R

r=R/3
4R
R

FIGURE 9.41
Figures for a) Problem 9.18, b) Problem 9.19.

9.19 (M) The small cone in Figure 9.41b rolls inside the large stationary cone with angular
velocity ω1 = 0.6 rad/s about its own axis. Find the angular acceleration of the small cone
Three-Dimensional Kinematics of Rigid Bodies 521

and the amount of time it takes to complete a circle. Both cones are right-angled and have
height 4R.

z Z
a) b)

G
4R
2.5R

2R
D

C C
Side view Front view

FIGURE 9.42
Figure for Problems 9.20 and 9.21. a) Side view, b) front view.

9.20 (D) The wheel of the unicycle in Figure 9.42 is of radius R and the rider is of height
4R. The orientation of the rider is the same as the shape frame. Using the Euler angle
transformation in Section 9.5.2, calculate the velocity and acceleration of the head of the
rider. The rider does not move with respect to the shape frame.
9.21 (M) Consider the unicycle in the previous problem with R = 2 ft. The unicycle is rolling
with steady precession with speed 8 ft/sec at a lean angle of θ = 10◦ . Using the formulation
that begins with the wheel upright as in Section 9.5.2, calculate the acceleration of the head
of the rider, assuming the rider does not move with respect to the shape frame.
9.22 (M) A disk of radius R in Figure 11.27b rotates freely about a rod of length L. As the
rod rotates about O with Ω, the disk rolls without slipping with ω. Obtain the relationship
between Ω and ω and calculate the angular velocity of the disk for when θ = 30◦ .
9.23 (C) Consider a disk of radius 12 in. that rolls with steady precession. Write a computer
program that solves for the nutation angle θ as a function of the speed v and radius of
curvature ρ. Generate a three-dimensional plot of Equation (9.86) with v and ρ as the
horizontal axes and θ as the vertical axis. Consider the ranges of 0 < v < 25 ft/sec and
0 < ρ < 30 ft.
9.24 (M) Consider the disk rolling without slipping in Figure 9.18. The 3-1-2 coordinate
transformation sequence is used, which begins with the disk upright. The following values
are given: heading angle ψ = 30◦ , ψ̇ = 0.1 rad/s, lean angle θ = 15◦ = constant, spin rate
φ̇ = 2 rad/sec, and radius of disk 7 in. Find the velocity of the center of the disk. Then, for
ψ̈ = θ̈ = 0, φ̈ = 0.2 rad/s2 , calculate the angular acceleration of the disk.
9.25 (M) The disk of mass m and radius R in Figure 9.43 rolls freely and without slip about
the end of a rod of length b on the horizontal plane. The rod rotates with constant angular
velocity Ω. Calculate the angular acceleration of the disk.
9.26 (M) The disk of radius R in Figure 9.44 rolls freely and without slip at the end of a rod
of length L. The rod rotates with constant angular velocity Ω about the Z axis. Calculate
the angular velocity and angular acceleration of the disk for when L/R = 6.
522 Applied Dynamics

!"# ' '


]ү $"# ]ү ș
ș

ȍ
$ [ү
" # ș &
# &
! %(
(
%
%

FIGURE 9.43
Figure for Problem 9.25. a) General view, b) side view.

" (
!"#$%
ȍ # Ȧ &"$'

(
! $
! $

' & ' &


% %

FIGURE 9.44
Figure for Problem 9.26.

Section 9.6—Orientation Change by Successive Rotations


9.27 (M) Beginning with XY Z axes, conduct an orientation change maneuver for the angles
θ1 = 35◦ about the Z axis, leading to the X 0 Y 0 Z 0 axes and then θ2 = −25◦ about the Y 0
axis. Discuss the resemblance to rotation about the X axis.
9.28 (C) Design an orientation change sequence that resembles a rotation of 20◦ about the Y
axis of an XY Z frame. Design the rotation angles that will enable you to do that. Conduct
a sensitivity analysis for different possible values of the rotation angles.

Section 9.7—Interconnections
9.29 (M) Door ABCD in Figure 9.45a, when perfectly aligned, rotates about a vertical axis.
Due to a defect, the hinge axis is no longer along the Z axis, but makes an angle φ with
the vertical, as shown in Figure 9.45b. Calculate the difference in distance between points
C and C 0 , when the door swings open by θ = 25◦ for when L = 6 ft, h = 3 ft, and φ = 5◦ .
9.30 (C) Consider the previous problem and plot the distance between point B and B 0 as
a function of the hinge angle θ (0 ≤ θ ≤ 45◦ ) and distortion of hinge axis φ (0 ≤ φ ≤ 15◦ ).
At what value of θ will the difference in the positions of B and B 0 will be the greatest?
Three-Dimensional Kinematics of Rigid Bodies 523

!" #"
( )
) $ү
ș
' %ү
‫׋‬

! "
ș

&
%
# % #
$ &ү

FIGURE 9.45
Figure for Problems 9.29 and 9.30. a) Aligned door, b) misalignment of hinge axis by φ.

9.31 (M) Rod AB in Figure 9.27 is constrained to move between the horizontal and vertical
guide bars. The velocity of point A is 2 m/s upward and A = 6 m, b = 3 m, and c = 4 m.
Find the velocity of point B and angular velocity of the rod.

a) b) z
y 10 cm
A B
O
vB
C 24 cm
6 ft
3 ft
y
2 ft B
15 cm
x A

z vA
18 cm
x

FIGURE 9.46
Figures for a) Problems 9.32 and 9.33, b) Problem 9.34.

9.32 (M) Guide bar OA in Figure 9.46a is along the x direction and guide bar BC lies on
the yz plane, and it makes an angle of 30◦ with the y axis. There is a ball-and-socket joint
at C and slider at point A. At the instant shown, collar C is moving up with velocity of 0.2
ft/sec. Calculate the velocity of point A and angular velocity of rod AC.
9.33 (M) Rod AC in Figure 9.46a is at rest when a force is applied to point A giving it an
acceleration of aA = 0.5i ft/sec2 . Calculate the acceleration of point C at this instant.
9.34 (M) The telescoping link AB in Figure 9.46b is attached to collars that move along
parallel guide bars. At the instant shown, vB = 4 cm/s and vA = 2 cm/s. Calculate the
angular velocity of the telescoping link when i) joint B is a ball-and-socket joint and A is a
slider, and ii) joint A is a ball-and-socket joint and B is a slider.
524 Applied Dynamics

Section 9.8—4×4 Matrix Description of a General Transformation


9.35 (C) A vector ri = 2J − 5K is rotated using a 3-2 transformation by angles ψ = 30◦ and
θ = −45◦ and then translated by 3 in the X direction and 8 in the Z direction. Find the
coordinates of the final position of the vector, rf , in i) XY Z frame, ii) xyz frame, which is
obtained after the 3-2 transformation.
9.36 (M) Express the coordinates of the mass m in Figure 5.18 using the matrices [Ttr ] and
[Trot ]. Identify the entries of all the matrices.
9.37 (D) Express the coordinates of point P in Figure 8.28 using the matrices [Ttr ] and
[Trot ]. Identify the entries of all the matrices.
9.38 (M) A vector ri in the XY Z frame is rotated by 30◦ about the Z axis (resulting in
the xyz axes) and then translated by 12I − 2J + 7K. If the final position of the vector is
rf = 5i + 9j − 6k, what is ri in terms of the XY Z coordinates?

Z
y B

3L/2
O Y

L
A x
X

FIGURE 9.47
Figure for Problem 9.39.

9.39 (M) The double link OAB in Figure 9.47 is brought into position by first rotating link
OA about the Z axis by θ and then by rotating link AB about the x axis by φ. Using a 4 × 4
matrix formulation, calculate the position of B in the XY Z frame. Hint: First calculate the
position of A.

Section 9.9—Euler Parameters


9.40 (E) The coordinate system a1 a2 a3 is transformed into the b1 b2 b3 system by first rotating
by an angle φ = 45◦ about the a2 axis and by rotating the resulting a01 a02 a03 frame about
the a03 axis by θ = 30◦ . Find the orientation of the principal line and the rotation angle Φ.
9.41 (E) The principal line makes angles of 60◦ , −60◦ , and 45◦ with the inertial X, Y , and
Z axes, respectively. A body is rotated about this axis by angle 36◦ . Calculate the Euler
parameters and the resulting rotation matrix [R].
9.42 (E) The principal line goes through the point (0.5, −0.7, 0.4). The rotation angle is
Φ = 75◦ . Calculate the Euler parameters and associated rotation matrix [R].
9.43 (M) The coordinate system a1 a2 a3 is transformed into the b1 b2 b3 system by first
rotating by an angle φ = 30◦ about the a2 axis and then by rotating the resulting a01 a02 a03
frame about the a03 axis by θ = 25◦ . Find the orientation of the principal line, the rotation
angle Φ, and associated Euler parameters.
9.44 (C) The orientation of a body is described by a 3-2-3 Euler angle rotation, with the
rotation angles ψ, θ, and φ. At a certain instant, the values of the Euler angles and their
Three-Dimensional Kinematics of Rigid Bodies 525

rates are ψ = 4◦ , θ = 30◦ , φ = 30◦ , ψ̇ = 0.2 rad/s, θ̇ = 0.1 rad/s, φ̇ = 1 rad/s. Find the
values of the Euler parameters and their derivatives at this instant.
9.45 (M) In Problem 9.42 you are also given that the angular velocities associated with the
body-fixed frame are ω1 = 0.1 rad/s, ω2 = 0.4 rad/s and ω3 = −2.3 rad/s. Calculate the
time derivatives of the Euler parameters ėi .

Section 9.10—Rodrigues Parameters


9.46 (E) Calculate the Rodrigues parameters associated with Problem 9.42.
9.47 (E) Calculate the Rodrigues parameters associated with Problem 9.41.
9.48 (M) Find the screw parameters associated with the transformation in Problem 9.35.
9.49 (M) A vector ri = 3I − 2J + 7K is rotated by an angle of Φ = 45◦ about an axis that
goes through the point (4, −3, 0) and whose unit vector is n = (−2I + 3J + 6K)/7. The
body then is translated by 2.7 along the screw axis. Find the resulting vector rf .
9.50 (M) A vector ri = I + 2J + 3K is rotated by an angle of Φ = −30◦ about an axis that
goes through the point (3, 4, 0) and whose unit vector is n = (2I − 3J + 6K)/7. The body
then is translated by ρ = 2.2 along the screw axis. Find the resulting vector rf .
9.51 (M) Using an xyz coordinate system, a body is rotated by Φ = 165◦ about a screw axis
with unit vector n = 1/3i + 2/3j + 2/3k. The screw axis crosses the yz plane at (0, 6, 6).
The value of the slide is ρ = 6.25. Calculate the final positions of the points A, B, C, and
D, whose initial positions are given in Example 9.11.
10
Mass Moments of Inertia

10.1 Introduction
Chapter 4 discussed mass moment of inertia for plane motion. We learned that three pieces
of information uniquely quantify the properties of a rigid body: the mass, the center of
mass, and the mass moment of inertia. This chapter extends the concept of mass moment
of inertia to three dimensions. The inertia properties are described in terms of the inertia
matrix. The translation and rotation of coordinates are considered and the parallel axis
theorem is extended to three dimensions. Principal moments of inertia are discussed.
The study of mass moments of inertia is needed for the analysis of three-dimensional
kinetics, as description of the angular momentum and kinetic energy is in terms of the
inertia matrix.

10.2 Center of Mass

!"
ȡ
#

!
!#
!"#$%&
'$($'$)*$&+,")-

FIGURE 10.1
Center of mass and differential element.

Consider the rigid body in Figure 10.1. The vector r, which is measured from a fixed
point, denotes the position of a differential mass element dm. The center of mass G is
denoted by rG and its location is defined as
Z
1
rG = rdm (10.1)
m body
R
where m = body dm is the mass of the body. It follows that because r = rG + ρ , where ρ

527
528 Applied Dynamics

is the position vector from the center of mass to the differential mass element,
Z
ρ dm = 0 (10.2)
body

We can calculate the center of mass of a body by subdividing it into parts whose in-
dividual centers of mass can be calculated more easily. If the body is separated into N
components, and the mass and center of mass of each component is denoted by mi and
rGi (i = 1, 2, . . . , N ), respectively, the center of mass can be found from
N
1 X
rG = mi rGi (10.3)
m i=1
P
where m = mi is the total mass. The mass of a body describes the amount of matter
contained in the body and it represents the resistance of the body to translational motion.

10.3 Mass Moment of Inertia


While the center of mass provides valuable information and simplifies the analysis of trans-
lational motion, it gives no measure of the way the mass is distributed on the body. It is
necessary to develop a quantity that describes the resistance of a body to rotational motion.
Such a quantity is dependent on how the mass is distributed on the body.
"

%&

$!

" !

!
#

FIGURE 10.2
Differential mass element and distance from x axis.

The coordinate system xyz in Figure 10.2 is fixed to the body and describes the config-
uration of a differential mass element by the vector r = xi + yj + zk, also written as the
column vector {r} = [x y z]T . The primary interest is in two quantities: the distribution of
mass with respect to a certain axis and the distribution of mass with respect to a certain
plane. Consider the x axis first. From Figure
p 10.2, the perpendicular distance of a differen-
tial element dm from the x axis is Rx = y 2 + z 2 . The mass moment of inertia about the
x axis is defined as
Z Z
2
y 2 + z 2 dm

Ixx = Rx dm = (10.4)
body body
Mass Moments of Inertia 529

In a similar fashion, the mass moments of inertia about the y and z axes are defined as
Z Z Z Z
2 2 2 2
x2 + y 2 dm (10.5)
 
Iyy = Ry dm = x + z dm Izz = Rz dm =
body body body body

The mass moment of inertia of a body about a certain axis becomes larger if the axis
is selected further away from the body. To analyze the distribution of mass with respect to
the xy, yz, and xz planes, introduce the products of inertia
Z Z Z
Ixy = xy dm Iyz = yz dm Ixz = xz dm (10.6)
body body body

It is clear that Ixy = Iyx and so forth. In general, the products of inertia do not contribute
too much to the physical description of the mass distribution, unless there are certain
symmetry properties. If the body has a plane of symmetry, the products of inertia associated
with coordinates normal to that plane vanish.

x x
y y
x

FIGURE 10.3
A body symmetric about the yz plane.

The concept is illustrated in Figure 10.3. When there is symmetry with respect to the
yz plane, for each point defined by the coordinates (x, y, z), there corresponds the point
(−x, y, z). The products of inertia Ixy and Ixz vanish because both these terms require
integration over x and the contribution of the differential element on the left side of the yz
plane is countered by the element on the right side. Table 10.1 summarizes the cases when
the products of inertia vanish.

TABLE 10.1
Cases when products of inertia vanish

Symmetry with Respect to Vanishing Inertia Terms


yz plane Ixy = Ixz = 0
xz plane Ixy = Iyz = 0
xy plane Ixz = Iyz = 0

If a rigid body is symmetric about an axis, then it must have symmetry about at least
530 Applied Dynamics

two planes. Thus, for a body that has an axis of symmetry, all products of inertia vanish
when one of the coordinate axes is the symmetry axis. It should be noted that a body need
not have planes of symmetry for the products of inertia to vanish. A proper orientation of
the xyz axes leads to the same result, as will be shown later.
The moments and products of inertia form the inertia matrix, denoted by [I] and defined
as
 
Ixx −Ixy −Ixz
[I] =  −Ixy Iyy −Iyz  (10.7)
 
−Ixz −Iyz Izz

The inertia matrix is symmetric and positive definite. Positive definiteness, discussed in
Chapter 7, can be verified by evaluating the minor determinants of [I]. In some cases, such
as slender rods or thin plates, components of the inertia matrix about certain axes will be
very small compared to inertia components about other axes. We can then assume that
these very small inertia components are negligible.
The units of mass moment of inertia are mass times length squared. The moments of
inertia of a body are usually written as the mass of the body multiplied by a length squared
parameter and an appropriate constant. For example, for a slender rod of mass m and length
1
L, the centroidal mass moment of inertia about a transverse axis is IG = 12 mL2 .
The mass, center of mass, and inertia matrix of a rigid body specify what are called the
internal properties of the body. When calculating the inertia matrix of a body, one must
specify the point about which the inertia matrix is calculated, as well as the orientation of
the coordinate axes used.
A useful quantity when dealing with mass moments of inertia is the radius of gyration.
The radius of gyration about the x, y, and z axes is defined as
r r r
Ixx Iyy Izz
κxx = κyy = κzz = (10.8)
m m m
and it represents the distance that a point mass m would have to be placed from the axis so
that it would have the same mass moment of inertia. Mass moments of inertia are expressed
as Ixx = mκ2xx , . . . . The radius of gyration is used when tabulating the inertia properties
of commonly used components such as I-beams and brackets. The Appendix gives the mass
moments of inertia of bodies with common shapes.

10.3.1 Calculation of the Mass Moments and Products of Inertia


We next focus on calculating entries of the inertia matrix. There are several possible ways,
depending on the choice of the differential element. Two are outlined below:
• Select the differential element as a differential mass and perform a triple integration.
For Cartesian coordinates xyz the differential element is dm = ρdxdydz, with ρ being
the density. Polar coordinates rθz come in handy for bodies with circular cross sections.
The differential area element is dA = rdrdθ, as shown in Figure 10.4, and the differential
mass is dm = ρrdrdθdz.
• Select the differential element so that one or two sides of the differential element have a
finite size. Here, it is customary to select the differential element as a thin rod or as a
thin plate, as shown in Figure 10.5. The differential element has a finite mass moment
of inertia about one or two axes. The expression for mass moment of inertia is written
as (say, about the z axis)
Mass Moments of Inertia 531
y

dr
r
rsin
x
rcos

FIGURE 10.4
Differential element in polar coordinates.

!" #"
"

! # !$ #$

!
#
"
%# "
"
%"
!
# %!

FIGURE 10.5
Differential element as a) a rod, b) a thin plate.

Z
Izz = dIzz (10.9)

in which dIzz is the mass moment of inertia of the differential element about the z axis.
The advantage of using such elements is that the number of integrations is reduced.
When the differential element is a rod, we can perform a double integration, and when
the differential element is a plate, we can evaluate the mass moment of inertia by a
single integral.
Consider a thin flat plate of thickness t, shown in Figure 10.6, and a differential element
dm = ρtdxdy. The mass moment of inertia of the thin plate about the z axis becomes
Z Z
2 2
x2 + y 2 dxdy = ρtJz
 
Izz = x + y dm = ρt (10.10)
body area

where Jz is recognized as the area polar moment of inertia about the z axis. Next, consider
the mass moments of inertia about the x and y axes. Noting that the thickness t is small,
for the x axis the mass moment of inertia becomes
Z Z
y 2 + z 2 dxdy = ρt y 2 + t2 dxdy
 
Ixx = ρt
area area
532 Applied Dynamics
#
$%
&
"

#
&
!

'

FIGURE 10.6
Thin plate differential element.

Z
≈ ρt y 2 dxdy = ρtIx (10.11)
area

in which Ix is the area moment of inertia of the plate about the x axis. In a similar fashion,
we can show that Iyy = ρtIy . From area moments of inertia, Jz = Ix + Iy . It follows that
for a thin plate oriented as in Figure 10.6, the mass moment of inertia about the z axis is

Izz = Ixx + Iyy (10.12)

Now, consider the differential element in Figure 10.5b in the form of a plate with
infinitesimal thickness dz. Denoting by x∗ y ∗ z ∗ the set of coordinate axes that are at-
tached to the differential element, the differential element has a mass moment of inertia
of dIzz = dIx∗ x∗ + dIy∗ y∗ . We then find Izz using the single integral
Z Z
Izz = dIzz = (dIx∗ x∗ + dIy∗ y∗ ) (10.13)

If there is an axis of symmetry in the body, we can select the z axis to exploit this symmetry
and have dIx∗ x∗ = dIy∗ y∗ , such that dIzz = 2dIx∗ x∗ .
To calculate Ixx and Iyy , the parallel axis theorem is invoked. Stated first in Chapter
4 for bodies in plane motion, the parallel axis theorem relates the moments of inertia of a
body about two axes that are parallel to each other. The general form of the parallel axis
theorem will be derived in the next section. Here, it is stated between the x and x∗ axes,
which are separated by a distance z, as shown in Figure 10.5b. If the z axis is the symmetry
axis, dIxx can be expressed as

dIxx = dIx∗ x∗ + z 2 dm (10.14)

For a differential element in the form of a circular plate, dIx∗ x∗ = dIzz /2 and

dIzz
dIxx = + z 2 dm (10.15)
2
It is clear that the approach of using thin plates as differential elements is most appealing
when there is a symmetry axis, as dIx∗ x∗ = dIzz /2 and the products of inertia vanish.
Mass Moments of Inertia 533
& &"

%
' *ү
$
$

! !"

FIGURE 10.7
Circular rod.

Example 10.1
Calculate the inertia matrix of a circular rod in Figure 10.7 of mass m, length L, and cross
section radius of R, about its center of mass.
Let us calculate the inertia elements by direct integration, as well as by means of the thin
plate approach. Consider direct integration first. It is preferable to use polar coordinates in
this case, and the differential element, shown in Figure 10.4, has the form
dm = ρrdrdθdz [a]
The first task is to calculate the mass, which is m = ρπR2 L. For illustrative purposes,
let us calculate it by triple integration. The integral has the form
Z L2 Z 2π Z R Z L2 Z 2π Z R
m = dm = ρr dr dθ dz [b]
−L
2 0 0 −L
2 0 0

The three variables r, θ, and z are independent of each other, so that the integrals can be
evaluated separately. Doing so gives
Z L2 Z 2π Z R
R2
m = ρ dz dθ r dr = ρL2π = ρπR2 L [c]
−L 2 0 0 2

The moments of inertia are calculated next. Note from Figure 10.4 that x = r cos θ and
y = r sin θ. Beginning with Izz (as z is the symmetry axis)
Z Z
x2 + y 2 dm = r2 dm

Izz = [d]
body body

Evaluating Equation [d] results in


Z L2 Z 2π Z R
R4 1 1
r3 dr dθ dz = ρL2π ρπR2 L R2 = mR2

Izz = ρ = [e]
−L 2 0 0 4 2 2

From symmetry, Iyy = Ixx , which is


Z Z Z
2 2 2
z 2 dm

Iyy = x + z dm = x dm + [f ]
body body body
534 Applied Dynamics

It is convenient to evaluate the two integrals separately. Recalling that cos2 θ dθ =


R
1
2 (θ + sin θ cos θ), the first integral is
L
2π R
R4
Z Z Z Z
2
2
2 1
x dm = ρ dz cos θ dθ r3 dr = ρLπ = mR2 [g]
body −L
2 0 0 4 4

The second integral becomes


L
Z Z Z 2π Z R
2
2 1 1
z 2 dz ρπR2 L L2 = mL2

z dm = ρ dθ r dr = [h]
body −L
2 0 0 12 12

so that the mass moments of inertia about the x and y axes become
1 1
Ixx = Iyy = mR2 + mL2 [i]
4 12
For a thin disk, the length L is much smaller than the radius R, and eliminating L we
get
1 1
For a disk: mR2 Izz =Ixx = Iyy ≈ mR2 [j]
2 4
For a slender rod, the radius is much smaller than the length. Eliminating R from the
inertia expressions results in
1
For a slender rod: Izz ≈ 0 Ixx = Iyy ≈ mL2 [k]
12
Because of the axisymmetry, all products of inertia vanish. Next, let us calculate the
mass moments of inertia using the thin plate approach. Taking the differential element as
a disk of radius R and thickness dz, the differential mass and differential mass moment of
inertia about the axis of symmetry become
1 2 1
dm = ρπR2 dz dIzz = R dm = ρR4 dz [l]
2 2
It follows that the mass of the rod is m = ρπR2 L. The mass moment of inertia about
the symmetry axis, Izz , becomes
L
ρπR2
Z Z 2 1 1
Izz = dIzz = dz = ρπR4 L = mR2 [m]
2 −L
2
2 2

To calculate the other mass moments of inertia consider the x∗ y ∗ z axes in Figure 10.7
and note that
1
dIx∗ x∗ = dIy∗ y∗ = dIzz [n]
2
Using the parallel axis theorem to calculate dIxx results in
1
dIxx = dIx∗ x∗ + z 2 dm =
dIzz + z 2 dm [o]
2
The first term in the right side of Equation [o] becomes 12 dIzz = 12 Izz = 1
mR2 . The
R
4
second term integrates to
Z Z L2  3
1 L 1
z 2 dm = ρπR2 z 2 dz = ρπR2 2 = mL2 [p]
−L 2
3 2 12

Adding the two integrated terms gives Equation [i].


Mass Moments of Inertia 535

a) y b)
y

b
y
z

O a O
x x
b

dm
x
c a
x
z z b – bx/a

FIGURE 10.8
a) Triangular wedge, b) differential element.

Example 10.2
Find the mass moments of inertia and products of inertia of the right triangular prism of
uniform density ρ shown in Figure 10.8a about point O.
The mass of the prism is m = ρV , where the volume is V = abc/2, so that the
mass Ris m = ρabc/2. ToR find the moments of inertia we need to calculate the expres-
sions x2 dm, y 2 dm, z 2 dm. Using the rectangular differential element in Figure 10.8b
R

the differential element is


dm = ρdxdydz [a]
The integrals become
b
a b− a x c a
ρa3 bc ma2
Z Z Z Z Z  
2 2 2 b
x dm = ρx dz dy dx = ρc x b − x dx = = [b]
0 0 0 0 a 12 6
b 3
a b− a x c a
ρab3 c mb2
Z Z Z Z Z 
2 2 ρc b
y dm = ρy dz dy dx = b − x dx = = [c]
0 0 0 3 0 a 12 6
b
a b− a x c a
ρc3 ρabc3 mc2
Z Z Z Z Z  
2 2 b
z dm = ρz dz dy dx = b − x dx = = [d]
0 0 0 3 0 a 6 3
The mass moments of inertia then become
 2
c2
 2
c2 a2 b2
   
b a
Ixx = m + Iyy = m + Izz = m + [e]
6 3 6 3 6 6

Note the similarity between Ixx and Iyy . The products of inertia are
b 2
Z Z a Z b− a x Z c Z a 
ρc b mab
Ixy = xy dm = ρxy dz dy dx = x b− x dx = [f ]
0 0 0 2 0 a 12
b
a b− a x c a
ρc2
Z Z Z Z Z  
b mac
Ixz = xz dm = ρxz dz dy dx = x b − x dx = [g]
0 0 0 2 0 a 6
536 Applied Dynamics

b 2
a b− a x c a
ρc2
Z Z Z Z Z 
b mbc
Iyz = yz dm = ρyz dz dy dx = b − x dx = [h]
0 0 0 4 0 a 6
The inertia matrix becomes
 
2b2 + 4c2 −ab −2ac
m
[I] =  −ab 2a2 + 4c2 −2bc  [i]

12
−2ac −2bc 2a2 + 2b2

10.4 Transformation Properties of the Inertia Matrix


Given the inertia properties of a body about a point and in terms of a particular set of
axes, it is desirable to relate these properties to the inertia matrix about a different point
or a different orientation of the coordinate axes. This permits us to find the mass moments
of inertia once and to then use the transformation equations to find the moments of inertia
about other points and about other axes, without having to perform the integrations again.
We will first study how the inertia properties change as the coordinate system is translated
and then as the coordinate system is rotated.
The inertia matrix can be expressed in column vector format and in terms of the position
vector {r}, which connects the point about which the inertia matrix is calculated to the
differential element, as
Z   Z  
T T T
[I] = {r} {r} [1] − {r} {r} dm = r2 [1] − {r} {r} dm (10.16)
body body

in which r2 = {r}T {r} = r · r and [1] is the identity matrix. The inertia matrix can also be
expressed as
Z
T
[I] = [r̃] [r̃]dm (10.17)
body

where [r̃] is the skew-symmetric matrix formed from the elements of {r} = [x y z]T . The
above definitions of the inertia matrix will be of use when discussing the dynamics of rigid
bodies in the next chapter.

10.4.1 Translation of Coordinates


Let O be the point about which the inertia matrix is calculated and consider a coordinate
system xyz, as shown in Figure 10.9. The associated inertia matrix is [IO ] and the inertia
components are IOxx , IOxy , . . . . The position vector from point O to a differential element
is r = xi + yj + zk. Next, consider another point, D. The vector from point O to D is

rD = rD/O = dx i + dy j + dz k (10.18)

The vector r00 from point D to the differential element then is

r00 = r − rD/O = (x − dx ) i + (y − dy ) j + (z − dz ) k (10.19)

To find the moments and products of inertia about D, replace x with x − dx , y with
Mass Moments of Inertia 537
'

&

!‫މމ‬ #'
#$

! !&
(
%
#!

!" #(

FIGURE 10.9
Translation of coordinates.

y − dy , and z with z − dz in Equations (10.4)–(10.5). For example, for the moment of inertia
about the x axis
Z  
2 2
IDxx = (y − dy ) + (z − dz ) dm
body

Z Z Z Z
y 2 + z 2 dm + d2y + d2z dm − 2dy
 
= y dm − 2dz z dm (10.20)

The first term on the right


 side of Equation (10.20) is recognized as IOxx . The second
term reduces to m d2y + d2z . The third and fourth terms are basically the first moments
of the mass distribution. If point O is selected as the center of mass, O = G, these terms
vanish. This is yet another advantage of dealing with the center of mass, and it leads to the
parallel axis theorem, which states that

IDxx = IGxx + m d2y + d2z



(10.21)

Mass moments of inertia have their lowest values when calculated about the center of
mass. The remaining moments and products of inertia associated with D are calculated in a
similar fashion. For example, for the product of inertia term IDxy and considering the case
when point O is the center of mass G,
Z
IDxy = (x − dx ) (y − dy ) dm
body

Z Z Z Z
= xy dm + dx dy dm − dy x dm − dx y dm = IGxy + mdx dy (10.22)

The parallel axis theorem is expressed in matrix form as


 2
dy + d2z −dx dy

−dx dz
[ID ] = [IG ] + m  −dx dy d2x + d2z −dy dz  (10.23)
 

−dx dz −dy dz d2x + d2y


538 Applied Dynamics

Using column vector notation, denoting by {r00 } the position vector associated with a
point observed from point D, such that {r00 } = {r} − {d}, in which {d} = [dx dy dz ]T , {r}
is measured from the center of mass G with r002 = {r00 }T {r00 }, the parallel axis theorem has
the form
Z  
T
[ID ] = r002 [1] − {r00 } {r00 } dm
body

Z h   i
T T T T
= r2 + d2 − 2 {r} {d} [1] − {r} {r}T + {d} {d} − {r} {d} − {d} {r} dm
body

Z  
T
= [IG ] + d2 [1] − {d} {d} dm (10.24)
body

where d2 = {d}T {d}, [1] is the identity matrix, and we note that {r}dm = {0}.
R

The above equation is significant because it permits calculation of the moments of inertia
about any set of axes parallel to the centroidal axes, as long as the inertia matrix about the
centroidal axes is known. For bodies with complex shapes, we split the body into smaller
parts whose centroidal moments of inertia can be calculated more easily or looked up in
a handbook. The total moment of inertia of the body is found by using the parallel axis
theorem and by adding up the individual moments of inertia. The Appendix gives the mass
moments of inertia of bodies with commonly found shapes.

10.4.2 Rotation of Coordinate Axes


Consider a set of axes x0 y 0 z 0 , obtained by applying a set of rotations to the original coordi-
nate system xyz. As previously, denote by {r} the position vector of a differential element
in terms of xyz coordinates. The same position vector, in terms of the x0 y 0 z 0 coordinates,
is written as {r0 }. From Chapter 2, using the direction cosine matrix [c] or rotation matrix
[R], we can relate the vectors {r} and {r0 } by

{r0 } = [R] {r} or {r0 } = [c]T {r} (10.25)


T
The magnitudes of the two vectors are the same, or r02 = {r0 }T {r0 } = {r}T [R] [R]{r}
= {r}T {r} = r2 . The inertia matrix about the primed axes is defined as
Z   Z  
T T
[I 0 ] = r02 [1] − {r0 } {r0 } dm = r2 [1] − {r0 } {r0 } dm (10.26)
body body

Substituting Equation (10.25) into this expression and noting that r2 [1] = [R]r2 [1][R]T
give the mass moment inertia about the rotated axes as
Z  
T T
[I 0 ] = [R] r2 [1] [R] − [R] {r} {r}T [R] dm
body

Z   
T
= [R] r2 [1] − {r} {r} dm [R]T (10.27)
body

Recognition of the middle term as the mass moment of inertia [I] about the xyz axes leads
to the result

[I 0 ] = [R] [I] [R]T or [I 0 ] = [c]T [I] [c] (10.28)


Mass Moments of Inertia 539

Note that for a rotational transformation it is not necessary to begin with a centroidal
set of axes. The relationship above is for an inertia matrix about any point. In general, we
select the xyz axes such that the calculation of the moments of inertia is simplified (e.g.,
using symmetry axes or symmetry planes). We then use a coordinate transformation to
obtain the moments of inertia about the desired axes.
A very important coordinate transformation is one that yields a diagonal inertia matrix.
The next chapter will discuss the significance of dealing with diagonal inertia matrices. The
question arises whether a rotation transformation can be found so that the resulting inertia
matrix is diagonal. As will be shown in the next section, such a transformation can be found
by solving an eigenvalue problem.
When the coordinate axes are both translated and rotated, the order of these operations
does not affect the final result.

Example 10.3
Calculate the center of mass and inertia matrix of the L-shaped uniform rod of mass m in
Figure 10.10 about point O.

2a
1 z
G1
G
O
G2 a

FIGURE 10.10
Rotating L-shaped rod.

It is convenient to separate the rod into two parts: part 1, of length 2a along the z axis,
and which has mass 2m/3, and part 2, of length a along the x axis, with mass m/3. The
center of mass is found using the relation

m1 r̄1 + m2 r̄2 = mrG [a]

where m1 = 2m/3, m2 = m/3, r̄1 = ak, r̄2 = a/2i + 2ak. The overbar denotes centers of
mass of the individual parts. Applying the center of mass formula gives
 
2 1 a  1 4
mrG = mak + m i + 2ak =⇒ rG = a i+ k [b]
3 3 2 6 3

The mass moments of inertia are calculated next. For part 1, since the z axis is the
symmetry axis, all products of inertia vanish and, using the relationship I = mL2 /3, the
nonzero moments of inertia about O are
1 2m 2 8
I1xx = I1yy = (2a) = ma2 [c]
3 3 9
540 Applied Dynamics

with all other inertia components being zero. For part 2, the centroidal mass moments of
inertia are
1 m 2 1
I¯2yy = I¯2zz = a = ma2 [d]
12 3 36
with all other inertia components being zero.
The parallel axis theorem in Equation (10.23) needs to be invoked for the second part.
Noting that r̄2 = a/2i + 2ak, we conclude that dx = a/2, dy = 0, dz = 2a. The added terms
in the parallel axis theorem are
 2
dy + d2z −dx dy 4a2 −a2
  
−dx dz 0
m m
 −dx dy d2x + d2z −dy dz  =  0 17a2 /4 0  [e]
 
3 2 2
3 2 2
−dx dz −dy dz dx + dy −a 0 a /4

Combining Equations [c], [d], and [e], the inertia matrix about point O is obtained as
 
20 0 −3
1
[IO ] = ma2  0 21 0 [f ]
9
−3 0 1

Example 10.4—Inertia Matrix for Unbalanced Tire


Consider a tire modeled as a thick disk of mass M , radius R, and thickness L. There is
an imbalance on the tire in the form of a concentrated mass m = M , where  is a small
number, as shown in Figure 10.11. Calculate the inertia matrix about the center of the disk.
Then, use static balancing to balance the tire and analyze the effect of static balancing on
the inertia matrix.

# #

$
!! "
' "
& (
'
!" '

#$%&'()* )&
! !
*+(,-.'*
)#

FIGURE 10.11
Portion of tire as a disk with an imbalance at z = a.

The mass moments of inertia of the disk without the imbalance were calculated in
Example 10.1 as
1 1 1
ICzz = M R2 ICxx = ICyy = M R2 + M L2 [a]
2 4 12
and all products of inertia vanish due to symmetry. The parallel axis theorem is used to
Mass Moments of Inertia 541

calculate the inertia matrix associated with the imbalance. The inertia matrix of a point
mass about itself is zero. Noting that the position vector from point C to the imbalance is

r1 = bj + ak [b]

so that dx = 0, dy = b, dz = a, we can calculate the contribution of the point mass as


 2
dy + d2z
 2
b + a2
 
0 0 0 0
[ICmass ] = [0] + m  0 d2z −dy dz  = m  0 a2 −ba  [c]
   

0 −dy dz d2y 0 −ba b2

The total inertia matrix is [IC ] = [ICdisk ] + [ICmass ], so


 1 2 1 2 2 2
 
4 R + 12 L +  a + b 0 0
1 2 1 2 2
[IC ] = M  0 R + 12 L + a −ba [d]
 
4 
1 2 2
0 −ba 2R + b

The xyz axes are no longer principal axes, as there are nonzero off-diagonal elements in
the inertia matrix due to the imbalance.1 An imbalance that is on the symmetry axis of the
disk (b = 0, a 6= 0), or on the plane of point C (a = 0, b 6= 0), changes the center of mass
location, but the inertia matrix remains diagonal. An imbalance of this form is called static
imbalance, as it can be corrected by means of static balancing. By contrast, when both
a 6= 0 and b 6= 0, there exists a dynamic imbalance, which results in off-diagonal elements
in the inertia matrix.
The objective in static balancing of tires is to ensure that the center of mass lies on the
axis of rotation. To this end, a counterweight of mass m is placed on the disk. Denoting the
location of the counterbalance by r2 = ex i + ey j + ez k, the position of the center of mass of
the entire assembly (disk, imbalance, counterbalance) becomes
1
rG = (M 0 + mr1 + mr2 )
M + 2m
1
= (mex i + m (b + ey ) j + m (a + ez ) k) [e]
M + 2m
Because the objective is to have the center of mass at the center of the disk, the desired
coordinates of the center of mass are x = y = z = 0. For this to happen, the counterweight
location is selected as ex = 0, ey = −b, ez = −a. This answer is relatively straightforward.
However, in real life, the location of the imbalance is not known. Further, the counter-
weight can be placed only at specific locations, namely at the rim of the wheel, limiting the
effectiveness of static balancing.
Let us examine the contribution of the counterbalance to the overall inertia matrix.
Considering Equation [c] and replacing dy with ey (or b with −b) and dz with ez (or a with
−a) the contribution of the counterbalance to the inertia matrix becomes
   
e2y + e2z 0 0 a2 + b2 0 0
[ICcounterbalance ] = m  0 e2z −ey ez  = m  0 a2 −ba  [f ]
   
0 −ey ez e2y 0 −ba b2

Once the counterbalance is added, the inertia matrix of the entire system (mass, imbal-
ance, counterbalance) has even larger off-diagonal terms. These off-diagonal terms lead to
1 Even if we calculate the center of mass of the disk with the imbalance and associated inertia matrix

using the xyz coordinates, the centroidal inertia matrix will not be diagonal.
542 Applied Dynamics

bearing forces, as will be discussed in the next chapter. What appears intuitively to be a
solution for the static case is not a solution for the dynamic case.
One way of minimizing the increase in off-diagonal terms is to take the counterbalance
mass, divide it into two, and put the two counterbalances on either side of the rim. The
coordinates of the two imbalances are then (ex , ey , c) and (ex , ey , −c), where 2c is the
rim width. This way, we correct the imbalance in the xy plane and the off-diagonal terms
in the inertia matrix remain unchanged.
Static balancing does not eliminate the off-diagonal elements of the inertia matrix. Dy-
namic balancing, which will be discussed in the next chapter, is necessary to diagonalize
the inertia matrix.

10.5 Principal Moments of Inertia


As will become abundantly clear in the next chapter, it is preferable to work with an inertia
matrix that is diagonal. The equations for rotational motion become considerably simpler.
When the coordinate axes are selected such that the products of inertia vanish, the axes are
called principal axes, and the moments of inertia are called principal moments of inertia.
There are two ways of finding the principal moments of inertia. The first is by visual
inspection. For example, if the coordinate axes are selected such that they lie on a plane of
symmetry, then at least two of the products of inertia vanish. If one of the axes is an axis
of symmetry, all products of inertia vanish.
The second method of finding principal moments of inertia is by solving the eigenvalue
problem associated with the inertia matrix. Since [I] is symmetric and positive definite, all
of its eigenvalues are real and positive. Its eigenvectors are real, as well.
From Equation (10.28), the transformation equation for the inertia matrix represents
an orthogonal transformation. Consider the case when [I 0 ] in Equation (10.28) is diagonal.
Because the transformation matrix [R] and direction cosine matrix [c] = [R]T are orthogonal
matrices, we can write Equation (10.28) as

[I 0 ] = [c]T [I] [c] [c]T [c] = [1] (10.29)

These equations are essentially the mathematical description of the eigenvalue problem
associated with the matrix [I]. The diagonal matrix [I 0 ] contains the eigenvalues of [I], and
[c] is the normalized eigenvector matrix with the eigenvectors as the columns of [c]. To verify
this, consider the eigenvalue matrix associated with the inertia matrix [I],

[I] {u} = λ {u} or ([I] − λ[1]) {u} = {0} (10.30)

where λ denotes the eigenvalues and {u} the eigenvectors. A solution to the above equation
exists if

det ([I] − λ[1]) = 0 (10.31)

This is called the characteristic equation, and it is a polynomial in λ of order three.


Solution of the characteristic equation yields three eigenvalues λi (i = 1, 2, 3) and associated
eigenvectors {ui }. The eigenvectors are orthogonal to each other, such that

{ui }T {uj } = 0 {ui }T [I] {uj } = 0 i 6= j; i, j = 1, 2, 3 (10.32)

and, as discussed in Chapter 7, they can be normalized so that {ui }T {ui } = 1 (i = 1, 2, 3).
Mass Moments of Inertia 543

It follows that {ui }T [I] {ui } = λi . We can then form the eigenvector matrix [U ] =
[{u1 } {u2 } {u3 }]. The eigenvector matrix is orthogonal (unitary), so [U ]T [U ] = [1]. The
orthogonality relationships with respect to the inertia matrix can be written as

[U ]T [I] [U ] = [λ] = [I 0 ] [U ]T [U ] = [1] (10.33)

where [λ] is a diagonal matrix containing the eigenvalues, which are the principal moments
of inertia. We recognize that [U ] = [c] and [λ] = [I 0 ].
Two issues are of significance when calculating the principal moments of inertia and the
eigenvector matrix:

• Construction of the direction cosine matrix. Chapter 2 discussed that for [c] (or
[R]) to describe a proper rotation, its determinant must be equal to 1. The determinant of
the normalized eigenvector matrix [U ] is ±1, with the choice of plus or minus depending
on the way the eigenvectors are normalized. This is because if {ui } is an eigenvector of
[I], so is −{ui }. For [U ] to represent a proper transformation matrix, the eigenvectors
should be normalized such that det [U ] = 1.
• Ordering of the principal moments of inertia. While not of mathematical sig-
nificance, ordering the principal moments of inertia helps us understand the inertia
properties better. In general, we write the eigenvalues in ascending (or descending) or-
der and construct the eigenvector matrix accordingly, as discussed in Chapter 7. In
vibration problems, the lower eigenvalues are usually of more interest than the higher
eigenvalues, as they contribute more to the motion amplitudes. When dealing with the
inertia matrix, there is no such need. We can order the eigenvalues arbitrarily.
One guideline is to order the principal moments of inertia such that they are close to
the order of the diagonal elements of [I]. Such an approach usually leads to the smallest
rotation angles between the original coordinates and principal axes.

10.5.1 Special Case: Repeated Principal Moments of Inertia


An interesting special case in eigenvalue analysis is that of systems possessing repeated
eigenvalues. In mathematics, a matrix that has repeated eigenvalues is referred to as degen-
erate. Inertia matrices having repeated eigenvalues have an elegant physical interpretation.
Consider a symmetric matrix that has two repeated eigenvalues λi = λj , with cor-
responding eigenvectors {ui } and {uj }. Because the eigenvalues are repeated, there is a
certain amount of arbitrariness in {ui } and {uj }. Any linear combination of these two
eigenvectors is also an eigenvector belonging to λi . That is, β1 {ui } + β2 {uj } is also an
eigenvector for any values of β1 and β2 . There is no unique eigenvector {ui } or {uj }. We
select these two eigenvectors so that {ui }T {uj } = 0.
Consider now a rigid body that has two of its principal moments of inertia the same, say,
I1 and I2 . Such a body is referred to as inertially symmetric. The most common example
of this is axisymmetric bodies, such as disks, circular rods, a rugby ball, or an American
football. Attach a set of axes b1 b2 b3 to the body, with the b3 axis as the symmetry axis, as
shown in Figure 10.12. Because the body possesses axial symmetry, all products of inertia
vanish and the symmetry axis is a principal axis.
In Figure 10.12, the b1 b2 plane is referred to as the principal plane. If the coordinate
axes are rotated about b3 to b01 b02 , these axes also constitute a set of principal axes. There
is no unique way of defining the principal axes, just as in the eigenvectors associated with
repeated eigenvalues.
Earlier in this chapter, the shape frame associated with axisymmetric bodies was intro-
duced. From the discussion above, using the shape frame does not alter the inertia matrix
544 Applied Dynamics

"!‫ މ‬# "!#


$%&'(&)*+,
ș )+*'-

!
"#

ș
""#
""‫މ‬

FIGURE 10.12
Axisymmetric body and principal axes.

of an axisymmetric body. This property will come in handy in the next chapter. The next
chapter will also show that axisymmetric bodies have better stability properties than ar-
bitrary bodies. Bodies that spin are usually designed to be axisymmetric, with the spin
axis and symmetry axis coinciding. A similar statement can be made about the stability of
vehicles, which are usually designed so that they have a symmetry plane.

Example 10.5

G y

FIGURE 10.13
Vehicle that is symmetric about the xz plane.

Given a rigid body, whose inertia matrix [IG ] about the xyz axes is
 
400 0 −125
[IG ] =  0 350 0  kg · m2
−125 0 100
find the principal moments of inertia, as well as the direction cosine matrix and rotation
angle that diagonalize the inertia matrix.
Mass Moments of Inertia 545

Two of the products of inertia are zero, IGxy = IGyz = 0, leading to the conclusion that
the body probably has symmetry about the xz plane, such as the vehicle shown in Figure
10.13. The eigenvalue problem is defined in Equation (10.31) as

([IG ] − λ[1]) {u} = {0} [a]

which requires that the determinant det ([IG ] − λ[1]) = 0, or


 
400 − λ 0 −125
det  0 350 − λ 0  = 0 [b]
−125 0 100 − λ

The associated characteristic equation is

(400 − λ)(350 − λ)(100 − λ) − (350 − λ) × 1252

= (350 − λ) ((400 − λ)(100 − λ) − 15, 625) = 0 [c]


We recognize right away that one of the eigenvalues is λ = 350 kg · m2 . Solution of the
characteristic equation yields three eigenvalues, which are the principal moments of inertia:

λ1 = I1 = 445.26, λ2 = I2 = 350, λ3 = I3 = 54.74 kg · m2 [d]

Note that the eigenvalues are ordered so that they follow the order of the diago-
nal elements of the inertia matrix (descending order). To find the corresponding eigen-
vectors, we use Equation [a] in conjunction with the eigenvalues. Using the notation
{ui } = [u1i u2i u3i ]T , for the first eigenvalue, λ1 = 445.26 kg·m2 , we write
  
400 − λ1 0 −125 u11
det  0 350 − λ1 0   u21  = {0} [e]
−125 0 100 − λ1 u31

which leads to the simultaneous equations

−45.26u11 − 125u31 = 0 u21 = 0 − 125u11 − 345.26u31 = 0 [f ]

From the middle equation above, u21 = 0. As discussed in Chapter 7, the other two
equations in Equation [f] are not independent. The equation on the right can be obtained
by multiplying the equation on the left by 125/45.26 = 2.7618. Solving the equation on the
left yields u31 = −45.26/125u11 = −0.3621u11 . The first eigenvector is
 T
{u1 } = d1 1 0 −0.3621 [g]

where d1 is an arbitrary constant. As discussed earlier, it is convenient to normalize the


eigenvector so that its magnitude is one. Doing so yields

{u1 }T {u1 } = d21 1 + 0.36212 = 1.131d21



[h]

so that d1 = ±1/ 1.131 = ±0.9403.
The second eigenvalue is 350 kg·m2 , unchanged from IGyy . We can observe this from the
 T
symmetry about the xz plane. The second eigenvector is thus {u2 } = 0 1 0 . This
 T
choice (and not {u2 } = 0 −1 0 ) implies that the y axis is not changed after the
rotation. The y axis is a principal axis. It follows that the coordinate transformation that
leads to the principal axes is a rotation about the y axis, also known as a 2 rotation.
546 Applied Dynamics

Using the same procedure as the first eigenvalue, the normalized third eigenvector be-
 T
comes {u3 } = 0.3404 0 0.9403 . The next step is to construct the eigenvector matrix,
keeping in mind that its determinant has to be equal to 1 (and not −1). Considering what
a 2 rotation looks like, the eigenvector matrix is written as
 
0.9403 0 0.3404
[U ] = [c] = [R]T =  0 1 0  [i]
−0.3404 0 0.9403

Comparing [U ] with the transformation matrix for a 2 rotation, denoting the rotation
angle by θ, cos θ = 0.9403 and sin θ = 0.3404. Solving for the rotation angle gives θ = 19.90◦ .
The rotation is shown in Figure 10.14 where the x0 z 0 axes are the principal axes.
"

Û

! #
Û

!"#$%#&'()'*+,

FIGURE 10.14
Principal axes x0 z 0 of vehicle.

10.6 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1998.
Greenberg, M.D., Advanced Engineering Mathematics, 2nd Edition, Prentice-Hall, 1998.
Kane, T.R., Likins, P.W., and Levinson, D.A., Spacecraft Dynamics, McGraw-Hill, 1984.

10.7 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 10.2—Center of Mass


10.1 (E) Find the center of mass of the truck in Figure 10.15a. Neglect the contribution of
the wheels.
Mass Moments of Inertia 547

a) b) y
y

3
d2 d1
2 r2 
2 a x

2 d3 r1
8
x r3
O

FIGURE 10.15
Figures for a) Problems 10.1 and 10.8, b) Problem 10.2.

10.2 (E) Calculate the center of mass of the plate with holes in Figure 10.15b for when
d1 = d2 = d3 = a/2, r1 = a/5, r2 = r3 = a/8.

.& /&
%
!
"$ !
!$ &%"&'('&$')'%"*+,-&$&
&$

# "
#

#$ !&$
%

"

FIGURE 10.16
Figures for a) Problems 10.3 and 10.9, b) Problem 10.4.

10.3 (M) Calculate the center of mass of the composite body in Figure 10.16a consisting of
a rod of length 5L along the x axis and a rod of length 7L bent into an L-shape in the yz
plane. The rods have the same mass density.

Section 10.3—Mass Moment of Inertia


10.4 (M) Calculate the mass moments of inertia about O of the axisymmetric shape in
Figure 10.16b.
548 Applied Dynamics

a) b)
y
y y=x
y=x2/4
B x
1
R
O x

z L x
O 1 2

FIGURE 10.17
Figures for a) Problem 10.5, b) Problem 10.6.

10.5 (M) Calculate the mass moments of inertia of the right circular cone in Figure 10.17a
about point O and using the xyz axes.
10.6 (M) Calculate the mass moments of inertia about the origin O of the body of revolution
generated by rotating about the y axis the area between the curves y = x2 /4 and y = x,
where 0 ≤ x ≤ 2, 0 ≤ y ≤ 1, as shown in Figure 10.17b.
10.7 (M) Calculate the mass moment of inertia of the torus in Figure 10.18a of mass m,
mean radius R, and core radius of the inner section a about the generating axis.

a) b) y

m, L
m, L

G m, L

R m, L
O x
a

m, L

m, L
z

FIGURE 10.18
Figures for a) Problem 10.7, b) Problem 10.10.

Section 10.4—Transformation Properties of the Inertia Matrix


10.8 (M) Obtain the inertia matrix of the truck in Figure 10.15a about point O and about
the center of mass G.
10.9 (M) Obtain the inertia matrix of the shape in Figure 10.16a about point O.
Mass Moments of Inertia 549

10.10 (M) Obtain the inertia matrix of the bent rod in Figure 10.18b about point O.

&

$
!
)*+,
)*+,+
)*+,+ '
% #
)*+,
"
(

FIGURE 10.19
Figure for Problem 10.11.

10.11 (M) Obtain the inertia matrix of the shape in Figure 10.19 about point O. Each
segment is a thin rod of mass m and length L.

a) y, y b)

y
y
x
5 x
D B

R
x O x
4 O 2
z
1 1

z, z L
7
z x

FIGURE 10.20
Figures for a) Problem 10.12, b) Problem 10.13.

10.12 (M) Consider the box in Figure 10.20a and determine the inertia matrix about point
O, using the x00 y 00 z 00 axes. The x00 axis goes through points O and D. The x00 y 00 z 00 axes are
obtained by first rotating the xyz axes about the −y axis by θ1 and rotating the resulting
frame x0 y 0 z 0 about the z 0 axis by θ2 .
10.13 (M) Consider the right circular cone in Figure 10.20b. Given R = 0.2041L, find the
elements of the inertia axis about point O using the axes x0 y 0 z 0 , which are obtained by
rotating the xyz axes about the z axis, so that the x0 axis goes through point B.
10.14 (M) Consider the right triangular prism of Example 10.2. Given that m = 1, a =
2, b = 1, c = 3, find the elements of the inertia matrix associated with a set of coordinates
obtained by rotating the xyz axes about the z axis such that the x0 axis is parallel with the
incline, as shown in Figure 10.21a.
10.15 (M) Consider the circular rod of Example 10.1. Find the elements of the inertia matrix
associated with a set of coordinates obtained by rotating the xyz axes about the y axis such
that the z 0 axis is along the line joining A and B, as shown in Figure 10.21b.
550 Applied Dynamics

a) b) x
z
y
y B
b=1
 R
G z


O x
 a=2 A L

y
x

FIGURE 10.21
Figures for a) Problem 10.14, b) Problem 10.15.

z
z m

L G m/4

h
O L/2 y
b
x

FIGURE 10.22
Figure for Problem 10.16.

10.16 (M) The shape in Figure 10.22 consists of a triangular prism and a concentrated mass
m/4. Calculate the center of mass, mass moment of inertia about O and z axis (IOzz ), and
(IGz0 z0 ). You can make use of the results in Example 10.2.

Section 10.5—Principal Moments of Inertia


10.17 (M) Calculate the principal moments of inertia of the mass with an imbalance in
Example 10.4 for when m = 0.05M, L = R/4, a = L/2, and b = 2R/3, and the angles
the principal axes make with the original axes. Next, redo the analysis for m = 0.1M and
compare the angles that the principal axes make with the original axes.
10.18 (C) Consider the right triangular prism of Example 10.2. Given that m = 1, a =
2, b = 1, c = 3 find the principal moments of inertia. Sketch the principal axes.
10.19 (C) Calculate the principal moments of inertia of the L-shaped rod in Example 10.3.
10.20 (C) Calculate the principal moments of inertia of the body whose inertia matrix [I]
is given below for the following values of a: a = 10, 100, −200. Compare the principal
moments of inertia and the direction cosines (eigenvectors) with each other and with the
Mass Moments of Inertia 551

case when a = 0.  
460 0 100
[I] =  0 500 a 
100 a 575
10.21 (M) Calculate the value of a and the angle by which the coordinate axes need to
be rotated if two of the principal moments of inertia of the body are to be the same. The
inertia matrix [I] is given below.
 
100 0 a
[I] =  0 325 0 
a 0 300
552 Applied Dynamics


11
Dynamics of Three-Dimensional Rigid Body
Motion

11.1 Introduction
This chapter extends the developments and derivations in previous chapters associated
with rigid body dynamics to three-dimensional motion. Linear and angular momentum
expressions are developed and used in conjunction with the translational and rotational
laws of motion. Analytical methods that make use of Lagrange’s and Kane’s equations are
used for obtaining the describing equations.
When obtaining the describing equations for rigid bodies, we have a number of choices
in selecting the motion variables. We will make use of these choices in this chapter and
develop several forms of the describing equations. The use of a reference frame attached to
the body continues to be the most common case for rigid bodies, because the elements of
the inertia matrix remain constant when using a set of body-fixed axes. For axisymmetric
bodies, use of the shape frame is attractive, as it leads to simpler expressions for angular
velocity and acceleration.
This chapter also develops tools to analyze the qualitative behavior of three-dimensional
rigid body motion, such as stability, motion integrals, work-energy principles, and impulse-
momentum principles.

11.2 Linear and Angular Momentum


Consider a rigid body, such as the one in Figure 11.1, with center of mass G. Also consider
an arbitrary point B, located on or off the body, and a differential element dm. As discussed
in Chapter 4, the linear momentum of the body, denoted by p, is defined as

Z Z
dr
p = v dm = dm (11.1)
body body dt

The position of the differential element is

r = rG + ρ (11.2)

whereR ρ connects the center of mass with the differential element. Recall from Chapter 4
that body ρ dm = 0. Using relative velocity relationships, the velocity of the differential
element is expressed as

v = vG + ω × ρ (11.3)

553
554 Applied Dynamics

dm
v
s
r G
rG/B B
rG
O
rB


FIGURE 11.1
A rigid body.

where ω is the angular velocity of the body. Substituting the above equation into the
expression for the linear momentum in Equation (11.1), we obtain
Z Z Z
p = (vG + ω × ρ ) dm = vG dm + ω × ρ dm = mvG (11.4)
body body body

The linear momentum of a rigid body is equal to its mass multiplied by the velocity
of the center of mass. Hence, the translational motion of a rigid body can be treated the
same way as the motion of a particle having the same mass as the rigid body and where
the entire mass of the body is concentrated at the center of mass of the body.
The angular momentum, or moment of the linear momentum, of the body about the
center of mass, denoted by HG , is defined as
Z Z
HG = ρ × v dm = ρ × (vG + ω × ρ ) dm (11.5)
body body
R
Since body ρ dm = 0, the expression in Equation (11.5) for the angular momentum about
the center of mass reduces to
Z
HG = ρ × (ω
ω × ρ ) dm (11.6)
body

It was shown in Chapter 4 that the angular momentum expression takes a simple form
for plane motion. This is not the case for three-dimensional motion. Consider a body-fixed
xyz coordinate system, and write the angular velocity ω and position vector ρ as

ω = ωx i + ωy j + ωz k ρ = xi + yj + zk (11.7)

Using the vector identity

ρ × (ω
ω × ρ ) = (ρρ · ρ ) ω − (ρρ · ω ) ρ (11.8)

expressing the angular momentum HG in terms of its components as

HG = HGx i + HGy j + HGz k (11.9)

and carrying out the algebra give the components of the angular momentum as
Z
 2
y + z 2 ωx − xyωy − xzωz dm
 
HGx =
body
Dynamics of Three-Dimensional Rigid Body Motion 555
Z
 2
x + z 2 ωy − xyωx − yzωz dm
 
HG y =
body
Z
x2 + y 2 ωz − xzωx − yzωy dm
  
HGz = (11.10)
body

Recalling the definitions for mass moments of inertia derived in the previous chapter,
the angular momentum components can be expressed as
HGx = IGxx ωx − IGxy ωy − IGxz ωz HGy = IGyy ωy − IGxy ωx − IGyz ωz

HGz = IGzz ωz − IGxz ωx − IGyz ωy (11.11)


or, in column vector format
{HG } = [IG ] {ω} (11.12)
where
 T T
{HG } = HGx HGy HGz {ω} = [ωx ωy ωz ] (11.13)
are the angular momentum and angular velocity vectors. The inertia matrix [IG ] is defined
in Equation (10.7).
The expression for the angular momentum about the center of mass can also be derived
in column vector format. Writing Equation (11.6) as
Z Z
HG = ρ × (ω
ω × ρ ) dm = −ρρ × (ρρ × ω ) dm (11.14)
body body

and using the column vector notation from Chapter 2, we can express Equation (11.14) as
Z
{HG } = − [ρ̃] [ρ̃] {ω} dm (11.15)
body

Noting that [ρ̃] is a skew-symmetric matrix, so that [ρ̃]T = −[ρ̃], and considering Equation
(10.17) we can write
Z
T
[ρ̃] [ρ̃] dm = [IG ] (11.16)
body

which leads to Equation (11.12).


The expression for angular momentum is simplified considerably when the coordinate
axes coincide with the principal axes or, for axisymmetric bodies, the shape frame. The
products of inertia vanish, and the angular momentum expression becomes
HG = IGxx ωx i + IGyy ωy j + IGzz ωz k (11.17)
or, using b1 b2 b3 as the coordinate frame,
HG = IG1 ω1 b1 + IG2 ω2 b2 + IG3 ω3 b3 (11.18)
where the body-fixed axes xyz or b1 b2 b3 are aligned with the principal axes.
When calculating the angular momentum, we need to decide on three issues: i) selection
of the point about which the angular momentum is written, ii) the reference frame to use,
and iii) the approach to use, algebraic vector or column vector.
The first and second issues will be addressed in subsequent sections. Regarding the third
issue, there is no general answer except for experience and personal preference. Many times,
it may be preferable to use a mixed description, as outlined in the example that follows. You
are encouraged to develop familiarity with the several different ways to calculate angular
momentum and, as we will shortly see, the describing equations.
556 Applied Dynamics

Example 11.1
A disk of weight 25 lb and radius R = 10 in. is mounted on a cart, as shown in Figure 11.2.
The disk rotates freely on the rod (the rod is not moving with respect to the cart) with
the spin rate of ω = 400 rpm. The cart is moving on a circular path with radius ρ = 120 ft
with a constant speed of v = 25 mph. Calculate the angular momentum of the disk about
its center of mass.

y
x
b
b

v
G
 z


FIGURE 11.2
Spinning disk on a cart.

The xyz coordinates are attached to the vehicle and constitute the shape frame for
the disk and they are also principal coordinates. The angular velocity of the disk has two
components due to i) the rotation of the cart and ii) the spin of the disk about its symmetry
axis. The angular velocity due to the motion of the cart is about the negative y axis

v 25 × 88
60
ωy = − = − = −0.3056 rad/sec [a]
ρ 120
and the angular velocity component due to the spin is
400 × 2π
ωz = − = −41.888 rad/sec [b]
60
so the angular velocity of the disk is the sum of the angular velocity components ω =
ωy j + ωz k = −0.3056j − 41.888k rad/sec.
The angular momentum of the disk about its center is

HG = IGyy ωy j + IGzz ωx k [c]


1 1
where IGzz = 2 mR2 and IGyy = 4 mR2 . The mR2 term comes out to be
 2
W 2 25 10
mR 2
= R = = 0.5397 slug · ft2 [d]
g 32.17 12
so the angular momentum becomes
 
1 1 rad
HG = 0.5397 × − 0.3056j − 41.888k = −0.04123j − 11.303k slug · ft2 · [e]
4 2 sec
Dynamics of Three-Dimensional Rigid Body Motion 557

One component of the angular velocity (and of the angular momentum) is much larger
than the other. This smaller component may be ignored for the purpose of calculating
the magnitude of the angular momentum or the magnitude of the total angular velocity.
However, when calculating the moment that needs to be generated by the supports of
the disk to hold the disk in place, this small component of the angular velocity plays an
important part, as we will soon see.

11.3 Transformation Properties of Angular Momentum


This section discusses translational and rotational transformations of angular momentum.
These transformation properties are very similar to the transformation properties associated
with mass moments of inertia.

11.3.1 Translation of Coordinates


Consider angular momentum expressions about an arbitrary point B that is fixed on the
body (or at an extension of the body that is moving with the body). From Figure 11.1, the
distance from B to the differential element is s, so that
s = rG/B + ρ (11.19)
Introducing the above equation to the definition of the angular momentum about B
results in
Z Z

HB = s × v dm = rG/B + ρ × (vG + ω × ρ ) dm (11.20)
body body

which, when expanded, yields


Z

HB = rG/B × vG + ρ × vG + rG/B × (ω
ω × ρ ) + ρ × (ω
ω × ρ ) dm (11.21)
body

The second and third terms onR the right side of Equation (11.21) vanish due to the
definition of the center of mass, ρ dm = 0. The last term is recognized as the angular
momentum about the center of mass HG , and the first term integrates to mrG/B × vG . The
angular momentum about point B can then be written as
HB = HG + mrG/B × vG (11.22)
If the body is rotating about a fixed point C that is on the body (or that can be viewed
as an extension of the body), we can express the velocity of the differential mass element
as v = ω × r, where r is the distance vector connecting the point C and the differential
element. The angular momentum expression about C then becomes
Z Z
HC = r × v dm = r × (ωω × r) dm (11.23)
body body

Comparing the above equation with Equation (11.14) we conclude that the angular mo-
mentum for a body rotating about a fixed point C can be written as
{HC } = [IC ] {ω} (11.24)
Caution: A body may be connected to a fixed point but it may not necessarily rotating
about that point. In such cases, Equation (11.24) does not apply. See Problems 11.4 and
11.5.
558 Applied Dynamics

11.3.2 Rotation of Coordinates


While the translation relationships for angular momentum are in terms of the center of
mass, the rotation relationships are valid for angular momentum calculated about any point.
Consider an arbitrary point B and a set of coordinates xyz. Given the angular momentum
about point B in terms of the xyz axes, the goal is to find the angular momentum in terms of
a set of rotated coordinates x0 y 0 z 0 . The derivation is conveniently carried out using column
vector notation.
The relationship between the two coordinates xyz and x0 y 0 z 0 is

{r0 } = [R] {r} (11.25)

where {r} is a vector in the xyz coordinate system and {r0 } is the same vector expressed
in the x0 y 0 z 0 coordinates. The angular momentum in terms of the xyz axes is {HB } and in
terms of the x0 y 0 z 0 axes it is {HB
0
}. These two vectors are related by
0
{HB } = [R] {HB } (11.26)

An interesting result arises when point B happens to be the center of mass or if B is


fixed and the body is rotating about B. For the center of mass, {HG } = [IG ] {ω}. From the
previous chapter, the following relationships are in effect:
0 T
{HG } = [R] {HG } {ω 0 } = [R] {ω} 0
[IG ] = [R] [IG ] [R] (11.27)

By manipulation of the above equations, we obtain the angular momentum in the x0 y 0 z 0


frame as
0
{HG } = [R] {HG } = [R] [IG ] {ω} = [R] [IG ] [RT ]{ω 0 } = [IG
0
] {ω 0 } (11.28)

It follows that the angular momentum about the center of mass in terms of the rotated
axes becomes
0 0
{HG } = [IG ] {ω 0 } (11.29)

and, for rotation about a fixed point C,

{HC0 } = [IC0 ] {ω 0 } (11.30)

where [IC0 ] = [R] [IC ] [RT ].


Considering the complicated nature of the moment of inertia and angular momentum
expressions, the advantages of using the principal axes becomes obvious. When calculating
the angular momentum of a body about a set of axes not coinciding with the principal axes,
it is often preferable to find the angular momentum using the principal axes and then use
the rotation relationship in Equation (11.26).

Example 11.2
Consider the spinning top in Figure 9.14 and calculate its angular momentum about the
pivot point O.
The top rotates about the fixed point O. Consider a 3-1-3 Euler angle transformation.
From Equation (9.48), the angular velocity vector of the body, expressed in the shape frame
f1 f2 f3 , is  
ω = ω1 f1 + ω2 f2 + ω3 f3 = θ̇f1 + φ̇ sin θf2 + φ̇ cos θ + ψ̇ f3 [a]
Dynamics of Three-Dimensional Rigid Body Motion 559

Denoting the centroidal mass moments of inertia by IG1 about the f1 and f2 axes and
IG3 about the f3 axis, the mass moments of inertia about point O are

IO1 = IO2 = IG1 + md2 IO3 = IG3 [b]

The angular momentum about O is

IG1 + md2 (ω1 f1 + ω2 f2 ) + IG3 ω3 f3



HO = [c]

Example 11.3

–x'
–x
–x'
 –x
z 
d
b
P1 C
c h b P2 y
x c

FIGURE 11.3
Spacecraft with solar panels.

The spacecraft in Figure 11.3 has a hub and two identical solar panels. It is rotating with
angular velocity Ω about the z axis, where the xyz coordinate system is attached to the
hub. The panels are being rotated by angle θ about the y axis. Find the angular momentum
of the spacecraft. Neglect the translational motion of the spacecraft.
The xyz axes are the principal axes for the hub, but not for the panels. For the panels,
it is more desirable to calculate the angular momentum about principal axes of the panels
and then to transfer the angular momentum to the xyz axes. The x0 y 0 z 0 axes are attached
to the panels. This reference frame is obtained by rotating the panels by θ about the y axis.
Figure 11.4 illustrates the relationship between the xyz and x0 y 0 z 0 axes.

z' z

x –x'

x'

–z'

FIGURE 11.4
The xyz and x0 y 0 z coordinates.
560 Applied Dynamics

The angular momentum of the hub is

RS2 h2
 
HS = mS + Ωk [a]
2 4

where mS and RS are the mass and radius of the hub.


To find the angular momentum of the panels, we can calculate the angular momentum
of each panel about its center of mass and then use Equation (11.22), or we can treat each
panel as rotating about the fixed point C (center of the hub). We will use both approaches.

Centers of Mass of Panels Approach


Using the equations for a rectangular prism, and considering the thickness of the panels to
be small, the centroidal moments of inertia for each panel are
1 1 1
mP c2 m P b2 mP b2 + c2

Ix0 x0 = Iy0 y0 = Iz0 z0 = [b]
12 12 12
where b and c are the dimensions of the panel along the x0 and y 0 axes, respectively, and mP
is the mass of each panel. All products of inertia are zero. Noting that the angular velocity
of each panel is
ω P = Ωk + θ̇j = −Ω sin θi0 + θ̇j0 + Ω cos θk0 [c]
the angular momentum of each panel about its center of mass can be written as
1  
mP −c2 Ω sin θi0 + b2 θ̇j0 + b2 + c2 Ω cos θk0

HP = [d]
12
The x0 y 0 z 0 coordinates are obtained by rotating the xyz coordinates by θ about the y
axis, so that the angular momenta in the two coordinate systems are related by

{HP0 } = [R] {HP } [e]

where the rotation matrix is


 
cos θ 0 − sin θ
[R] =  0 1 0  [f ]
sin θ 0 cos θ

Hence, the angular momentum of each panel about the xyz coordinate system becomes
1
 mP c2 Ω sin θ


cos θ 0 sin θ − 12
T
{HP } = [R] {HP0 } =  0 1 0  1 2
12 mP b θ̇
 

− sin θ 0 cos θ 1 2 2

m b + c Ω cos θ 12 P

b2 Ω sin θ cos θ
 
1
= mP  b2 θ̇ [g]
 
12

c2 Ω + b2 Ω cos2 θ
The angular momentum of each panel about its center of mass is the same.
The next step is to find the velocities of the center of mass of the two panels and to
invoke the parallel axis theorem. The velocity of the center of the first panel is
 
vP1 = ω × rP1 /C = Ωk + θ̇j × −dj = dΩi [h]
Dynamics of Three-Dimensional Rigid Body Motion 561

and
mP rP1 /C × vP1 = −mP dj × dΩi = mP d2 Ωk [i]
Similarly, for the second panel
 
vP2 = ω × rP2 /C = Ωk + θ̇j × dj = −dΩi [j]

mP rP2 /C × vP2 = mP dj × −dΩi = mP d2 Ωk [k]


Considering Equations [a], [g], [i], and [k], the angular momentum of the entire system
becomes
HC = HS + HP1 + mrP1 /C × vP1 + HP2 + mrP2 /C × vP2 [l]
which can be written in column vector format as
1 2
 
6 mP b Ω sin θ cos θ
1 2
{HC } =  6 mP b θ̇
[m]
 

1 2 1 2 1
 2 2 2
 2
mS 2 RS + 4 h Ω + 6 mP c Ω + b Ω cos θ + 2mP d Ω

Rotation about C Approach


Since each panel rotates about point C, the inertia matrix about C is calculated by means
of Equation (10.24). Consider the panel on the right. The vector connecting C and P2 is
rP2 /C = dj, so that dx0 = 0, dy0 = d, dz0 = 0. The inertia matrix of each panel about point
C and using the x0 y 0 z 0 frame becomes
c2
 
12 + d2 0 0
[IC ] = [IP ] + mP d2 [1] − {d}{d}T b2

= mP  0 0 [n]
 
12 
b2 +c2 2
0 0 12 +d

The angular momentum of each panel about point C becomes


  2  
c
−mP 12 + d2 Ω sin θ
 
1
{HC } = [IC ]{ω} = 
 mP b2 θ̇ 
[o]
 2 122

  
mP b 12+c
+ d2 Ω cos θ

T
Left multiplying the above equation by [R] , as was done in Equation [g], multiplying by
two (to account for the two panels), and adding Equation [a] gives the angular momentum
of the spacecraft.

Example 11.4
Find the angular momentum of the rolling cone in Figure 9.20 about the pivot point O.
The center of the base of the cone B has speed v. Use the h1 h2 h3 coordinates to express
the angular momentum.
The cone rotates about the fixed point O. We can solve this problem using Equation
(11.26) or Equation (11.30). Using the shape frame F where the coordinate axes are f1 f2 f3
and are related to the h1 h2 h3 axes by

h1 = f1 h2 = cos βf2 + sin βf3 h3 = − sin βf2 + cos βf3 [a]


562 Applied Dynamics

so that the rotation matrix from the F frame to the H frame has the form
 
1 0 0
[R] =  0 cos β sin β  [b]
0 − sin β cos β
Because the cone is axisymmetric, all products of inertia vanish. The center of mass
of the cone is 3L/4 away from point O and the centroidal mass moments of inertia are
IG3 = 3mR2 /10 and IG1 = IG2 = 3mR2 /20 + 3mL2 /80. Using the parallel axis theorem,
the mass moments of inertia in the F frame about O become
 2
3L 3 3
m 4R2 + 16L2 mR2

I1 = I2 = IG1 + m = I3 = [c]
4 80 10
From Example 9.5, using an instant center analysis and noting that the contact point
A is the instant center, the angular velocity of the cone is
v v v v
ω = − h3 = − (− sin βf2 + cos βf3 ) = f2 − f3 [d]
L sin β L sin β L L tan β
Using the shape frame, the angular momentum about O becomes
v v
HO = I1 f2 − I3 f3 [e]
L L tan β
so that, using Equation (11.26) and column vector notation, the angular momentum in the
H frame becomes
  
1 0 0 0
{H HO } = [R] {F HO } =  0 cos β sin β   I1 v/L 
0 − sin β cos β −I3 v/L tan β
 
0
=  (I1 − I3 ) v cos β/L [f ]
 

2
−I1 sin β/L − I3 v cos β/L sin β
where the subscript on the left denotes the frame in which the angular momentum vector
is resolved.
Using Equation (11.30), we first calculate the inertia matrix associated with the H
frame. Noting that [F IO ] = diag (I1 , I1 , I3 ), where I1 and I3 are defined in Equation [c],
the inertia matrix in terms of the H frame becomes
 
I1 0 0
T
[H IO ] = [R] [F IO ] [R] =  0 I1 cos2 β + I3 sin2 β (I3 − I1 ) sin β cos β  [g]
 

0 (I3 − I1 ) sin β cos β I1 sin2 β + I3 cos2 β


Right multiplying this matrix with the angular velocity vector, which in the H frame has
the form {H ω} = [0 0 − v/L sin β]T , so that {H HO } = [H IO ]{H ω}, yields Equation [f].
You should compare the two solutions and determine which approach you prefer.

11.4 General Describing Equations


This section develops the governing equations of dynamics, namely force and moment bal-
ances and Newton’s Second Law and Euler’s Law of rotation. The general forms of these
Dynamics of Three-Dimensional Rigid Body Motion 563

equations were stated in Chapter 4. We revisit these equations here, because for three-
dimensional motion there are different choices for expressing the linear and angular mo-
mentum, as well as the applied moments. These choices include
• Velocity, angular velocity, linear momentum, and angular momentum expressed in terms
of inertial, body-fixed, or other coordinates (such as the shape frame);

• Algebraic vector or column vector formulation; and


• Moment balance about the center of mass or about another point.
Moreover, there are several different forms of the force and moment balances. Experience
allows us to select the most convenient and appropriate form of the describing equations.
We begin this section by revisiting the concept of resultant force and moment and
continue with the force and moment balances.

11.4.1 Resultant Force and Moment

  


 
 
  
  
    

 

FIGURE 11.5
a) Forces and moments, b) resultant force and moment, c) resultant about B.

As discussed in Chapter 4, given a body to which a series of forces Fi and moments Mi


are applied, as shown in Figure 11.5a, the net effect of these forces and moments can be
expressed as a resultant force F going through the center of mass, where
X
F = Fi (11.31)

and a resultant moment MG about the center of mass, where


X X
MG = r i × Fi + Mi (11.32)

in which ri connects the center of mass to the point through which Fi is applied. The
resultant force and moment are depicted in Figure 11.5b.
It is also possible to express the net effect of the forces and moments acting on a body
as a resultant force F applied through an arbitrary point, say, B, and a moment MB about
B, as illustrated in Figure 11.5c. Consider Figure 11.5b and sum moments about B. The
moment about B has the form

MB = MG + rG/B × F (11.33)
564 Applied Dynamics

11.4.2 Force and Moment Balances


It was shown in Chapter 4 that for the translational motion of a rigid body, the force
balance, or Newton’s Second Law, officially stated in 1687, is expressed as
d d
p = mvG = maG = F (11.34)
dt dt
Defining the inertia force acting on the body as −maG , Newton’s Second Law can be
described as the inertia force being equal and opposite to the applied force.
The law governing the rotational motion, formally stated by Euler in 1765, states that
the rate of change of the angular momentum about the center of mass is equal to the sum
of all applied moments about the center of mass, or
d
HG = MG (11.35)
dt
For most rigid body problems, the linear and angular momenta are expressed in terms of
a moving reference frame. Denoting the angular velocity of this frame by ω f and applying
the transport theorem, the above two equations can be written as
d d d d
p = prel + ω f × p = F HG = HGrel + ω f × HG = MG (11.36)
dt dt dt dt
The term ω f × HG is known as the gyroscopic moment, and it describes the change in
direction in the angular momentum as the body rotates.
Equations (11.34) and (11.35) provide a complete description of the governing equations
of a rigid body. They constitute the basis of Newtonian mechanics, also known as the
Newton-Euler formulation.

11.4.3 Moment Balance about an Arbitrary Different Point


The translational equations of motion are not written about (or do not reference) a partic-
ular point, as linear momentum is an absolute quantity. To obtain the rotational equations
about a point other than the center of mass, consider the angular momentum balance about
an arbitrary point, say, B, that is fixed on the body (or at an extension of the body that
moves with the body).1 We can obtain the moment balance about B in two ways. In the
first approach, we sum moments about B and invoke Equation (11.33), with the result
d
MB = MG + rG/B × F = HG + mrG/B × aG (11.37)
dt
The above relationship is very useful, especially when B is selected as a point through
which reaction forces go. On the other hand, Equation (11.37) requires that the rates of
change of both the angular momentum and the linear momentum be evaluated.
In the second approach, Equation (11.22) is differentiated, with the result
d d d  d
HB = HG + m rG/B × vG = HG + mrG/B × aG + mvG/B × vG (11.38)
dt dt dt dt
The sum of moments about B is calculated in Equation (11.37) and is equal to the first
two terms on the right side of Equation (11.38). The last term on the right of Equation
(11.38) can also be expressed as
mvG/B × vG = m (vG − vB ) × vG = mvG × vB (11.39)
1 Make sure that the point B is fixed to the body and not to the reference frame. This is especially

important when dealing with the shape frame.


Dynamics of Three-Dimensional Rigid Body Motion 565

so introduction of Equation (11.37) and Equation (11.39) into Equation (11.38) results in

d
HB = MB + mvG × vB (11.40)
dt
Equations (11.37) and (11.40) are the most general forms of the moment balance equa-
tions. The moment balance equation about a fixed point C is
d
HC = MC (11.41)
dt
Example 11.5
Consider Example 11.1. Calculate the gyroscopic moment.
The xyz frame attached to the cart in Figure 11.2 is the shape frame for the disk. It
follows that the angular velocity of the shape frame is
v
ω f = ωy j = − j = −0.3056j rad/sec [a]
ρ
The angular momentum was obtained as

HG = IGyy ωy j + IGzz ωz k [b]

The gyroscopic moment is



ω f × HG = ωy j × IGyy ωy j + IGzz ωz k = IGzz ωy ωz i [c]

Note that this term involves both ωy and ωz . Evaluating it we obtain the gyroscopic moment
as
1
ω f × HG = IGzz ωy ωz i = × 0.5397 × (−0.3056) × (−41.888) = 3.4543k lb · ft [d]
2
Gyroscopic moment becomes a significant quantity for rapidly spinning bodies.

11.5 Description in Terms of Body-Fixed Coordinates


For three-dimensional motion, using inertial coordinates as motion variables does not give
too much physical insight and usually leads to lengthy expressions. It is preferable to write
the force and moment balances using the coordinates of a moving reference frame. The most
commonly used reference frame is the body-fixed frame.
Consider a body-fixed xyz frame, where the angular velocity of the frame is ω = ωx i +
ωy j + ωz k. The angular acceleration for a body-fixed coordinate system is

α = αx i + αy j + αz k = ω̇ ω rel + ω × ω = ω̇x i + ω̇y j + ω̇z k


ω = ω̇ (11.42)

so that the angular acceleration has a simple form.


The rate of change of the linear momentum becomes
d d
ω × vG = F
p = m vG = maGrel + mω (11.43)
dt dt
566 Applied Dynamics

In terms of the A and B frame notation,

A d B d
m vG = mvG + mAω B ×A vG = F (11.44)
dt dt
The derivative of the angular momentum is obtained in a similar way, with the result
d
HG = ḢGrel + ω × HG = MG (11.45)
dt
or, in the A and B frame notation,

A d B d A
HG = HG + ω B × HG = MG (11.46)
dt dt
The term ω × HG is the gyroscopic moment, and it describes the change in direction in
the angular momentum as the body rotates.
Let us expand the force and moment balances in terms of the body-fixed frame. The
velocity of the center of mass is written as vG = vx i + vy j + vz k, so aGrel = v̇x i + v̇y j + v̇z k.
Carrying out the cross product mω ω ×vG in Equation (11.43), we obtain for the translational
equations

m (v̇x + ωy vz − ωz vy ) = Fx m (v̇y + ωz vx − ωx vz ) = Fy

m (v̇z + ωx vy − ωy vx ) = Fz (11.47)

where the components of the resultant force are F = Fx i + Fy j + Fz k. In column vector


form, the translational equations are written as

m {v̇Grel } + m [ω̃] {vG } = {F } (11.48)

where [ω̃] is the skew-symmetric angular velocity matrix.


The rotational equations are obtained in a similar fashion. Writing the angular momen-
tum as {HG } = [IG ] {ω} and resultant moment about the center of mass as {MG } =
[MGx MGy MGz ]T , the rate of change of the angular momentum in Equation (11.45) be-
comes

{ḢG } = {ḢGrel } + [ω̃]{HG } = [IG ] {α} + [ω̃] [IG ] {ω} (11.49)

It follows that the angular momentum balance is

{ḢG } = [IG ] {α} + [ω̃] [IG ] {ω} = {MG } (11.50)

Expanding the equation above gives



IGxx αx − IGxy (αy − ωx ωz ) − IGxz (αz + ωx ωy ) − IGyy − IGzz ωy ωz

−IGyz ωy2 − ωz2



= MG x

IGyy αy − IGyz (αz − ωx ωy ) − IGxy (αx + ωy ωz ) − (IGzz − IGxx ) ωx ωz

−IGxz ωz2 − ωx2



= MG y
Dynamics of Three-Dimensional Rigid Body Motion 567

IGzz αz − IGxz (αx − ωy ωz ) − IGyz (αy + ωx ωz ) − IGxx − IGyy ωx ωy

−IGxy ωx2 − ωy2



= MG z (11.51)

with the reminder that αx = ω̇x , αy = ω̇y , αz = ω̇z .


The rotational equations are quite lengthy, even though there are no translational veloc-
ity terms in them. The complicated terms arise because both the angular momentum and
the angular velocity change directions for general three-dimensional rotation.
We can achieve significant simplification of the rotational equations by selecting the
body axes xyz as principal axes. In that case, all products of inertia vanish. Denoting the
principal axes by I1 , I2 , and I3 and the angular velocity components by ω1 , ω2 , and ω3 ,
the angular momentum balance equations become

I1 ω̇1 − (I2 − I3 ) ω2 ω3 = M1 I2 ω̇2 − (I3 − I1 ) ω1 ω3 = M2

I3 ω̇3 − (I1 − I2 ) ω1 ω2 = M3 (11.52)

Equations (11.52) are known as Euler’s equations. Note the order of the mass moments of
inertia in Equation (11.52), which lends itself to a convenient way of remembering these
equations.
The rotational equations can be described as resolving the angular momentum balance
into components along body-fixed axes. Writing the unit vectors associated with the coor-
dinate system as b1 , b2 , and b3 , Euler’s equations are expressed as
     
ḢG − MG · b 1 = 0 ḢG − MG · b 2 = 0 ḢG − MG · b 3 = 0 (11.53)

A rigid body that is unrestrained in rotation has three rotational degrees of freedom and
Euler’s equations become the equations of motion. When the motion is constrained and the
body has less than three rotational d.o.f., Euler’s equations yield the equation(s) of motion
as well as expressions for the reactions.
Among common uses of the rotational equations is instantaneous analysis, such as finding
angular accelerations, finding the gyroscopic moment, or calculating reactions at supports.
These equations can also be integrated qualitatively to come up with integrals of the motion
as well as for conducting a stability analysis. However, we cannot integrate just these equa-
tions to find the orientation of a body at a given instant. For that purpose, the rotational
equations need to be combined with kinematic differential equations.
The frequently encountered case of axisymmetric bodies and use of the shape frame in
conjunction with Euler’s equations will be discussed in the next section.

Example 11.6
A slender rod of mass m and length L is attached to a rotating support by a pin joint. The
support spins with constant angular speed Ω, as shown in Figure 11.6a. Find the equation
of motion of the rod.
This problem has one degree of freedom, and we select the motion variable as the angle
θ that the rod makes with the vertical. A set of body-fixed coordinates b1 b2 b3 are selected,
with the b1 axis along the rod. The free-body diagram is shown in Figure 11.6b. The rod
swings in the b1 b2 plane. There are three reaction forces and two reaction moments at point
C. Because the motion is rotation about a fixed point, moment equations about the fixed
point C are preferred in this problem.
568 Applied Dynamics

  



  






  



 

FIGURE 11.6
a) Slender rod, b) free-body diagram.

Noting that K = − cos θb1 + sin θb2 , the angular velocity of the rod is

ω = ΩK + θ̇b3 = −Ω cos θb1 + Ω sin θb2 + θ̇b3 [a]

Because a body-fixed frame is used, the angular acceleration can be obtained by simple
differentiation
α = Ωθ̇ sin θb1 + Ωθ̇ cos θb2 + θ̈b3 [b]
The mass moments of inertia about C are
1
I1 ≈ 0 I2 = I3 = mL2 [c]
3
so combining Equations [a] and [c], the angular momentum about C becomes
1  
HC = mL2 Ω sin θb2 + θ̇b3 [d]
3
The next step is to differentiate the angular momentum

ḢC = ḢCrel + ω × HC [e]

The relative rate of change of angular momentum about point C is


1  
ḢCrel = mL2 Ωθ̇ cos θb2 + θ̈b3 [f ]
3
and the transport term is
1  
ω × HC = mL2 Ωθ̇ cos θb2 − Ω2 sin θ cos θb3 [g]
3
The moment of all forces acting about the pivot C is
L 1
MC = M1 b1 + M2 b2 − b1 × −mgK = M1 b1 + M2 b2 − mgL sin θb3 [h]
2 2
We now invoke Euler’s equations, which yield

For b1 : 0 = M1 [i]
Dynamics of Three-Dimensional Rigid Body Motion 569

For b2 : 2I2 Ωθ̇ cos θ = M2 [j]


1
For b3 : I3 θ̈ − I2 Ω2 sin θ cos θ = − mgL sin θ [k]
2
Equation [k] is the equation of motion, while Equations [i] and [j] give expressions for
the reactions. Equation [k] can be simplified to
3g
θ̈ − Ω2 sin θ cos θ + sin θ = 0 [l]
2L
The reaction forces can be obtained from a force balance analysis.

11.6 Angular Momentum Balance for Axisymmetric Bodies


As discussed in the previous chapter, when dealing with an axisymmetric body, it is more
convenient to make use of the shape frame, also known as the F frame. Even though the
reference frame is not attached to the body, the inertia matrix remains constant. In addition,
we deal with a simpler form for the angular velocity.
We can use the shape frame in conjunction with Euler’s equation in two ways: i) we can
perform the differentiation of the angular velocity and angular momentum using the shape
frame, or ii) we can express the Euler’s equations in terms of the components of the shape
frame. We will see both approaches in sequence. Before that, we point to an interesting
characteristic of the moment balance equations written about an arbitrary point. Derived
in Equation (11.40), the moment balance equations about an arbitrary point B are given
by
d
HB = MB + mvG × vB (11.54)
dt
In the above equation, point B is fixed on the body (or at an extension of the body
that is moving with the body). When using the shape frame, because the shape frame is
not attached to the body, point B is not necessarily fixed in the shape frame. Denoting the
point on the shape frame that coincides with point B as B 0 , so that B 0 moves in the shape
frame and
A F
vB 6= vB 0 or vB 6= vB (11.55)

the moment balance about an arbitrary point is written by replacing vB with vB 0 as


d
HB = MB + mvG × vB 0 (11.56)
dt

11.6.1 Modified Euler’s Equations


The Modified Euler’s equations are primarily used for axisymmetric bodies. The equations
are based on writing the angular momentum balance in terms of the shape frame. They
sometimes are used with bodies of arbitrary shape (see Problem 11.18).
Consider the rotational motion equations about the center of mass
d
HG = MG (11.57)
dt
570 Applied Dynamics

Taking the derivative from a moving reference frame, say, the shape frame F , and denoting
the angular velocity of the shape frame by Aω F = ω f , we can write the Modified Euler’s
equations as

A d F d A
HG = HG + ω F × HG = MG (11.58)
dt dt
The matrix representation of the angular velocity is [ω̃f ]. The Modified Euler equations
can be written in column vector format as

[IG ] {ω̇}rel + [ω̃f ] [IG ] {ω} = {MG } (11.59)

Consider an axisymmetric body, such as the spinning top in Figure 9.14, and describe the
orientation of the body by means of a 3-1-3 coordinate transformation. In the inertial a1 a2 a3
coordinate system, the symmetry axis of the top is aligned with the a3 axis, the vertical.
Begin with the precession angle φ and rotation about the a3 axis, leading to the a01 a02 a03
coordinates. The second rotation is by the nutation angle θ about the a01 axis, resulting in
the shape frame. The spin, by ψ̇, is about the f3 axis.
The angular velocities of the shape frame and body are
A
ω F = ω f = φ̇a3 + θ̇a01 = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 (11.60)

 
A
ω B = ω b = φ̇a3 + θ̇a01 + ψ̇f3 = θ̇f1 + φ̇ sin θf2 + φ̇ cos θ + ψ̇ f3 (11.61)

and the spin angular velocity is


F
ω B = ω s = ω b − ω f = ψ̇f3 (11.62)

Denote the angular velocity components of the body and of the shape frame as ωi and
ωfi (i = 1, 2, 3), respectively. The angular velocity components are related to each other by

ω1 = ωf1 = θ̇ ω2 = ωf2 = φ̇ sin θ ω3 = φ̇ cos θ + ψ̇

ω2
ωf3 = φ̇ cos θ = (11.63)
tan θ
The inertia matrix is [IG ] = diag (I1 I1 I3 ). Carrying out the algebra, the Modified
Euler’s equations become

I1 ω̇1 + ω2 (I3 ω3 − I1 ωf3 ) = M1 I2 ω̇2 − ω1 (I3 ω3 − I1 ωf3 ) = M2

I3 ω̇3 = M3 (11.64)

in which the derivatives of the angular velocities are

ω̇1 = θ̈ ω̇2 = φ̈ sin θ + φ̇θ̇ cos θ ω̇3 = φ̈ cos θ − φ̇θ̇ sin θ + ψ̈ (11.65)

Note that these derivatives are considerably shorter than angular acceleration components
expressed in terms of a body-fixed frame. The first two angular momentum balances can
also be written as
   
I1 ω2 I1 ω2
I1 ω̇1 + ω2 I3 ω3 − = M1 I2 ω̇2 − ω1 I3 ω3 − = M2 (11.66)
tan θ tan θ
Dynamics of Three-Dimensional Rigid Body Motion 571

11.6.2 Expressing the Rotational Equations Using Components of


Shape Frame
In this approach, the original rotational equations are resolved in terms of the shape frame
components. Because the angular velocity vector is resolved in the shape frame (and not
using the angular velocity of the shape frame), the notations {F ω} and {F α} are used
to denote that these vectors are expressed in terms of the F frame. Hence, the rotational
equations are written as

{F ḢG } = [IG ] {F α} + [F ω̃] [IG ] {F ω} = {F MG } (11.67)

Note that all angular velocities in the above equation are body angular velocities. The
above equation requires evaluation of the angular acceleration term. Note that, because a
body-fixed reference frame is not being used, the expression for angular acceleration needs
to be obtained by applying the transport theorem to the angular velocity, with the result

{F α} = {F ω̇} + [F ω̃f ] {F ω} = {F ω̇} + [F ω̃f ] {F ωs } (11.68)

Considering the spinning top  discussedabove, where the angular velocity is ω = ω1 f1 +


ω2 f2 + ω3 f3 = θ̇f1 + φ̇ sin θf2 + φ̇ cos θ + ψ̇ f3 , and the angular velocity of the shape frame
is ω f = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 , the angular acceleration terms become
    

 θ̈ 
 0 −φ̇ cos θ φ̇ sin θ  θ̇ 

{F α} = φ̈ sin θ + φ̇θ̇ cos θ +  φ̇ cos θ 0 −θ̇  φ̇ sin θ
 
   
φ̈ cos θ − φ̇θ̇ sin θ + ψ̈ −φ̇ sin θ θ̇ 0 φ̇ cos θ + ψ̇
   

ω22
   

 θ̈ + φ̇ψ̇ sin θ 
 ω̇1 − tan θ + ω2 ω3
= φ̈ sin θ + θ̇φ̇ cos θ − θ̇ψ̇ =  ω̇2 + ω1 ω2 (11.69)
− ω1 ω3 
 
  tan θ
φ̈ cos θ − φ̇θ̇ sin θ + ψ̈
 
ω̇3

and the second term in Euler’s equations become (note that the angular velocity of the
shape frame is not involved in this expression)

[F ω̃] [IG ] {F ω}

     
0 −ω3 ω2 I1 0 0 ω1 (I3 − I1 )ω2 ω3
=  ω3 0 −ω1   0 I1 0   ω2  =  (I1 − I3 )ω1 ω3  (11.70)
−ω2 ω1 0 0 0 I3 ω3 0

Left multiplying Equation (11.69) with the inertia matrix and adding it to the above
equations results in the same equations as the Modified Euler’s equations for the top in
Equation (11.64).

Example 11.7
Consider Example 11.1. Calculate the reactions at the point where the disk is connected to
the cart. The speed of the cart and the angular velocity of the disk are constant.
The xyz frame attached to the cart in Figure 11.2 is the shape frame, whose angular
velocity is
v
ω f = ωy j = − j = −0.3056j rad/sec [a]
ρ
572 Applied Dynamics

The angular momentum was obtained as

HG = IGyy ωy j + IGzz ωz k [b]

Using the shape frame, the derivative of the angular momentum is


d d
HG = HGrel + ω f × HG [c]
dt dt
where
d
HGrel = IGyy ω̇y j + IGzz ω̇z k = 0 [d]
dt
and 
ω f × HG = ωy j × IGyy ωy j + IGzz ωz k = IGzz ωy ωz i [e]
is the gyroscopic moment.

y
x
FA
FB
a
a
A G
z
B
mg

FIGURE 11.7
Free-body diagram of disk.

The gyroscopic moment is balanced by the moment generated by the reactions at the
supports. Denoting the distance between the ends of the supports by 2a = 9 in., and noting
that the disk axis is in the z direction, it follows that the reaction forces need to act in the y
direction to produce a moment in the x direction. Figure 11.7 shows the free-body diagram
of the disk. Denoting the forces on each end of the rod (assumed to be massless) supporting
the disk by FA and FB , the sum of moments about the center of mass become

MG = (aFA − aFB ) i = IGzz ωy ωz i = MG i [f ]

where

MG = IGzz ωy ωz = 0.5 × 0.5397 × (−0.3056) × (−41.888) = 3.4543 lb · ft [g]

To solve for the two forces FA and FB , an additional equation is needed, which comes
from the vertical force balance
X
Fy = 0 = FA + FB − mg [h]

The solution to Equations [f] and [h] becomes

MG 1 MG 1
FA = + mg = 17.106 lb FB = − + mg = 7.8943 lb [i]
2a 2 2a 2
Dynamics of Three-Dimensional Rigid Body Motion 573

Let us analyze the numbers. The gyroscopic moment creates a 9.2 lb (17.1−7.9) difference
between the two supports. For a disk of the same weight but twice the radius, the mass
moment of inertia would be four times as much and so would the gyroscopic moment,
reaching a value of MG = 3.45 × 4 = 13.8 lb·ft. The forces at the supports would become
FA = 30.9 lb and FB = −5.9 lb, resulting in a force difference of 36.8 lb, exactly four times
the previous force difference. The force at B is now downwards and the gyroscopic moment
dominates the weight.
You can observe this phenomenon by lifting the front wheels of a bicycle using the
handlebars. If you the spin the front wheel rapidly, you’ll notice that it becomes difficult to
rotate the handlebars.

Example 11.8

Z
y

z
Z

Y G
X R mg

C F x
Y
FX
FZ

FIGURE 11.8
Free-body diagram of rolling disk.

Obtain the equations of motion of the disk in Figure 9.16, which is rolling without
slipping and sliding, by summing moments about the contact point. Use a 3-1-3 Euler angle
transformation and the shape frame.
The free-body diagram is shown in Figure 11.8. It is preferable to write the moment
balance using Equation (11.56) and about the point of contact C. There are three unknown
contact forces at C, and we do not wish to calculate their magnitudes, at least for the sake
of this problem.2 Note that, while the contact point C has zero velocity for roll without
slip, the second term on the right side of Equation (11.40) does not disappear, as the shape
frame is being used and C is replaced by C 0 in Equation (11.56). The describing equation
is
d
HC = MC + mvG × vC 0 [a]
dt
From Equation (9.62) and Equation (9.63) the angular velocities of the disk and of the
shape frame are
 
ω = ωx i + ωy j + ωz k = θ̇i + φ̇ sin θj + φ̇ cos θ + ψ̇ k [b]

2 In an actual application, we would want to know the magnitudes of these contact forces, which prevent

sliding and slipping.


574 Applied Dynamics

ωy
ω f = ωx i + ωy j + k [c]
tan θ
The mass moments of inertia of the disk are
1 1
Ixx = Iyy = mR2 Izz = mR2 [d]
4 2
so that the angular momentum is
1
HG = Ixx ωx i + Iyy ωy j + Izz ωz k = mR2 (ωx i + ωy j + 2ωz k) [e]
4
Note that the angular momentum in the above equation is expressed in terms of the com-
ponents of the angular velocity ωx , ωy , ωz . This approach is simpler algebraically than
writing the angular momentum in terms of the Euler angles. This same approach will be
used later in this chapter, when obtaining the equations of motion of the rolling disk by
means of Kane’s equations.
The next step is to find the angular momentum of the disk about the contact point C.
Use of Equation (11.22) gives

HC = HG + mrG/C × vG [f ]

To evaluate the second term on the right in Equation [f], it is necessary to calculate the
velocity of the center of the disk, which is

vG = ω × rG/C = (ωx i + ωy j + ωz k) × Rj = −Rωz i + Rωx k [g]

so
mrG/C × vG = mRj × (−Rωz i + Rωx k) = mR2 ωx i + mR2 ωz k [h]
The angular momentum about C thus becomes
1
HC = HG + mrG/C × vG = mR2 (5ωx i + ωy j + 6ωz k) [i]
4
Next, the angular momentum is differentiated. The transport theorem gives

ḢC = ḢCrel + ω f × HC [j]

with the individual terms as


1
ḢCrel = mR2 (5ω̇x i + ω̇y j + 6ω̇z k)
4
 ωy  1
ω f × HC = k × mR2 (5ωx i + ωy j + 6ωz k)
ωx i + ωy j +
tan θ 4
" ! #
1 2
ωy2  ω ω
x y

= mR 6ωy ωz − i+ 5 − 6ωx ωz j − 4ωx ωy k [k]
4 tan θ tan θ
The moment about C is due to gravity, which has the form

F = −mgK = −mg cos θk − mg sin θj [l]

so the moment about C becomes

MC = rG/C × F = Rj × mg (− cos θk − sin θj) = −mgR cos θi [m]


Dynamics of Three-Dimensional Rigid Body Motion 575

Next, evaluate the last term in Equation [a], mvG × vC 0 . The velocity of the contact
point in the shape frame, vC 0 , is obtained by
 ωy 
vC 0 = vG + ω f × rC/G = −Rωz i + Rωx k + ωx i + ωy j + k × −Rj
tan θ
 ωy 
= −R ωz − i = −Rψ̇i [n]
tan θ
This result can be explained by noting that if we look from afar at a disk that is rolling,
and the disk is smooth with no distinctive marks on it, we get the impression that the
contact point of the disk is moving. The eye is fooled by the change in location of the
contact point.
The last term in Equation [a] becomes
 ωy   ωx ωy 
mvG × vC 0 = m (−Rωz i + Rωx k) × −R ωz − i = −mR2 ωx ωz − j [o]
tan θ tan θ
Using Equations [k], [m], and [o], the equations of motion are obtained as
!
1 2
ωy2
In x direction: mR 5ω̇x + 6ωy ωz − + mgR cos θ = 0
4 tan θ

1  ωx ωy 
In y direction: mR2 ω̇y + − 2ωx ωz = 0
4 tan θ
1
In z direction: mR2 (6ω̇z − 4ωx ωy ) = 0 [p]
4
Using Equation [b], we introduce to Equation [p] the expressions for ωx , ωy , and ωz
in terms of the Euler angles which leads to the equations of motion in terms of the Euler
angles
1  
mR2 5θ̈ + 6φ̇ sin θ(ψ̇ + φ̇ cos θ) − φ̇2 sin θ cos θ + mgR cos θ = 0
4
1   1  
mR2 φ̈ sin θ − 2θ̇ψ̇ = 0 mR2 6ψ̈ + 6φ̈ cos θ − 10φ̇θ̇ sin θ = 0 [q]
4 4
Example 11.9
Obtain the equation of motion of the spinning top in Figure 9.14 by means of the Modified
Euler’s equations. Then, calculate the reactions at O and calculate the amount of friction
that is necessary to keep the top from sliding, when point O is not fixed.
The free-body diagram is given in Figure 11.9. The motion of the top is rotation about
the fixed point O. We will use a 3-1-3 Euler angle transformation and the shape frame.
From Example 11.2, the angular velocity vector of the body and of the reference frame are
 
ω = ω1 f1 + ω2 f2 + ω3 f3 = θ̇f1 + φ̇ sin θf2 + φ̇ cos θ + ψ̇ f3

ω2
ω f = θ̇f1 + φ̇ sin θf2 + φ̇ cos θf3 = ω1 f1 + ω2 f2 + f3 [a]
tan θ
Denoting the centroidal mass moments of inertia by IG1 about the f1 and f2 axes and
IG3 about the f3 axis, using the parallel axis theorem, the mass moments of inertia about
point O become
I1 = IG1 + mL2 I3 = IG3 [b]
576 Applied Dynamics

f3 a3 , a3

M 

 f2
L
G
F3

F2
mg a2
O
F1


f1, a1

FIGURE 11.9
Free-body diagram of spinning top.

The gravity force is F = −mga3 = −mg (cos θf3 + sin θf2 ). The sum of moments about O
is due to the weight and has the form

MO = M f3 + rG/O × F = M f3 + Lf3 × −mg (cos θf3 + sin θf2 ) = M f3 + mgL sin θf1 [c]

The angular momentum about O is

HO = I1 ω1 f1 + I1 ω2 f2 + I3 ω3 f3 [d]

Differentiating the above equation yields

ḢO = ḢOrel + ω f × HO [e]

where
ḢOrel = I1 ω̇1 f1 + I1 ω̇2 f2 + I3 ω̇3 f3 [f ]
and  ω2 
ω f × HO = f3 × (I1 ω1 f1 + I1 ω2 f2 + I3 ω3 f3 )
ω1 f1 + ω2 f2 +
tan θ

ω2
   ω ω 
1 2
= −I1 2 + I3 ω2 ω3 f1 + I1 − I3 ω1 ω3 f2 [g]
tan θ tan θ
Combining Equations [c], [f], and [g], the equations of motion are obtained as

ω22
I1 ω̇1 − I1 + I3 ω2 ω3 = mgL sin θ
tan θ

ω1 ω2
I1 ω̇2 + I1 − I3 ω1 ω3 = 0 I3 ω̇3 = M [h]
tan θ
To find the reactions
P at the point of contact, we need to examine the translational
equations, maG = F. The velocity of the center of mass is

vG = ω × rG/C = (ω1 f1 + ω2 f2 + ω3 f3 ) × Lf3 = Lω2 f1 − Lω1 f2 [i]


Dynamics of Three-Dimensional Rigid Body Motion 577

Differentiating the above equation, the acceleration is obtained as


ω2
aG = Lω̇2 f1 − Lω̇1 f2 + (ω1 f1 + ω2 f2 + f3 ) × (Lω2 f1 − Lω1 f2 )
tan θ

ω22
 
 ω1 ω2 
f2 − L ω12 + ω22 f3

= L ω̇2 + f1 + L −ω̇1 + [j]
tan θ tan θ
The forces that act on the top are

F = F1 f1 + F2 f2 + F3 f3 − mg (cos θf3 + sin θf2 ) [k]


P
so from maG = F, the values of the reaction forces are

ω22
 
 ω1 ω2 
F1 = mL ω̇2 + F2 = mL −ω̇1 + + mg sin θ
tan θ tan θ

F3 = −mL ω12 + ω22 + mg cos θ



[l]
The forces in the above equation are along the axes of the shape frame. In order to relate
them to the normal and friction forces, it is necessary to express these forces in terms of
the inertial frame, or in terms of the a01 a02 a03 axes, which are obtained after the precession
rotation. Because the shape frame is obtained by rotating the a01 a02 a03 axes about a01 by θ,
we can express the relationship between the a01 a02 a03 and f1 f2 f3 axes as

a01 = f1 a02 = cos θf2 − sin θf3 a03 = sin θf2 + cos θf3

f1 = a01 f2 = cos θa02 + sin θa03 f3 = − sin θa02 + cos θa03 [m]
Considering Equations [m] and [l], the forces acting along the a01 a02 a03 axes are

ω22
 
0 0
cos θ + mL ω12 + ω22 sin θ

F1 = F1 F2 = F2 cos θ − F3 sin θ = mL −ω̇1 +
tan θ

ω2
 
F30 = F3 cos θ + F2 sin θ = −mL ω12 + ω22 cos θ + mL −ω̇1 + 2

sin θ + mg [n]
tan θ
The assumption of rotation about a fixed point O can be valid only if the resultant of
the friction forces F10 and F20 is less than the maximum friction force µs F30 , or
p
F102 + F202
< µs [o]
F30

where µs is the coefficient of static friction.

11.7 Stability Analysis of Rotational Motion


We may wish to ascertain the nature of the motion of a body without explicitly solving
the equations of motion. Such qualitative analysis can be carried out by linearizing the
equations of motion about an operating position or by evaluating integrals of the motion.
578 Applied Dynamics

This section discusses linearizing the rotational equations of a rigid body about an operating
point.
Consider a rigid body and a set of body-fixed principal axes b1 b2 b3 . The body has an
initial motion in the form of rotation about one of the axes, say, b1 , so ω1 = Ω, ω2 = ω3 = 0.
Is this motion stable or not? Stated otherwise, if we apply a small moment to the body,
how will the body behave and what will the ensuing motion look like?
The angular velocity after the application of the moment is

ω 1 = Ω + 1 ω2 = 2 ω3 = 3 (11.71)

where i (i = 1, 2, 3) are small angular velocities. We wish to determine the evolution of


these perturbed angular velocities in time. Introducing the angular velocities into Euler’s
equations and noting that no further moments are applied gives
 
I1 Ω̇ + ˙1 − (I2 − I3 ) 2 3 = 0 I2 ˙2 − (I3 − I1 ) (Ω + 1 ) 3 = 0

I3 ˙3 − (I1 − I2 ) (Ω + 1 ) 2 = 0 (11.72)

Because the change in the angular velocities i is small with respect to Ω, the above
equations can be linearized by eliminating quadratic and higher-order terms in i (i =
1, 2, 3). Noting that Ω is constant, the linearized equations become

I1 ˙1 = 0 I2 ˙2 + (I1 − I3 ) Ω3 = 0 I3 ˙3 + (I2 − I1 ) Ω2 = 0 (11.73)

The first of the above equations indicates that 1 remains constant. To understand
the behavior of the remaining angular velocities, an eigenvalue analysis provides insight.
Introducing the expansions

2 (t) = E2 eλt 3 (t) = E3 eλt (11.74)

into Equation (11.73) gives the matrix eigenvalue problem

(I1 − I3 ) Ω
    
I2 λ E2 0
eλt = (11.75)
(I2 − I1 ) Ω I3 λ E3 0

As discussed in Chapters 5 and 7, for the above equation to have a nontrivial solution, the
determinant of the coefficient matrix must be zero, which yields the characteristic equation

I2 I3 λ2 − (I2 − I1 ) (I1 − I3 ) Ω2 = 0 (11.76)

The solution is
s
(I2 − I1 ) (I1 − I3 )
λ = ±Ω (11.77)
I2 I3

Two types of solutions are possible. If I2 > I1 and I3 > I1 , that is, if I1 is the smallest
moment of inertia, or if I2 < I1 and I3 < I1 , that is, if I1 is the largest moment of inertia,
the roots of the characteristic equation λj (j = 1, 2) are pure imaginary, and the system
is critically stable. A higher-level stability analysis indicates that rotation about the axes
representing the minimum or maximum moments of inertia are indeed stable.
On the other hand, if I2 > I1 and I3 < I1 , or if I2 < I1 and I3 > I1 , that is, if axis b1
is the intermediate moment of inertia axis, the roots of the characteristic equation λj are
Dynamics of Three-Dimensional Rigid Body Motion 579

real, with one root negative and the other positive. The conclusion is that rotation about
the intermediate axis of inertia is unstable.
The results for when all three principal moments of inertia are distinct is shown in Figure
11.10. Any initial rotation about the minimum and maximum moments of inertia axes is
stable, and the rotation will continue after the application of small disturbances. Initial
rotation about an intermediate axis is unstable and turns into wobbly motion. You can
demonstrate these results by taking a book (or other object with three distinct moments of
inertia) and spinning it about its principal axes, as illustrated in Figure 11.10.

b2
2 (stable)

G b1
1
(stable)
3
b3 (unstable)

FIGURE 11.10
A rigid body with distinct principal moments of inertia.

For an axisymmetric body, two of the mass moments of inertia are the same. There are
two possibilities. When b1 is the symmetry axis and I2 = I3 , the body is rotating about
the symmetry axis, as in a properly thrown football or a frisbee. From Equation (11.77),
all roots λj are pure imaginary, so the outcome is critical stability.
In the second case, b1 is a transverse axis and I1 = I2 or I1 = I3 . For both possibilities,
Equation (11.77) indicates that the roots of the characteristic equation are zero. The motion
can be explained as the continuation of tumbling motion and transfer of energy between
the two transverse axes.
Table 11.1 summarizes the stability results. It turns out that in real-life applications,
rotation is stable only about the axis of maximum moment of inertia. This result holds
whether the body is axisymmetric or not. Energy loss and energy transfer due to flexibility
and friction are the sources of instability. In space mechanics, nutational instabilities were
observed after the launch of the Explorer 3 satellite (Figure 1.3). This instability is discussed
in Section 11.13.4. Ideally, the shape of a space vehicle should be like disk or a sphere.
Because of practical considerations, spacecraft are designed as slender bodies.

11.8 Steady Precession of a Rolling Disk


An interesting case of rolling is steady precession, where a disk follows a circular path with-
out any change in the nutation angle θ. The kinematics of steady precession was discussed
in Section 9.5.3. This section analyzes the kinetics of steady precession.
Figure 11.11 shows the geometry and free-body diagram of the disk. The velocity of the
580 Applied Dynamics

TABLE 11.1
Stability results for a body with initial angular velocity

Type of Body Initial Rotation about Ensuing Motion


Minimum moment of inertia axis Stable
Arbitrary shape Maximum moment of inertia axis Stable
Intermediate axis Unstable, wobbly
Symmetry axis Stable
Axisymmetric
Transverse axis Not unstable, tumbling

Z
. y
. 

z


G
mg
.
C 
F x

FIGURE 11.11
Free-body diagram of disk.

center of a rolling disk, using a 3-1-3 Euler angle transformation, is


 
vG = ω × rG/C = Rθ̇k − R φ̇ cos θ + ψ̇ i (11.78)

The equations of motion of a rolling disk were derived in Example 11.8. They are repeated
here as
1    
mR2 5θ̈ + 6φ̇ sin θ ψ̇ + φ̇ cos θ − φ̇2 sin θ cos θ + mgR cos θ = 0 (11.79)
4

1  
mR2 φ̈ sin θ − 2θ̇ψ̇ = 0 (11.80)
4

1  
mR2 6ψ̈ + 6φ̈ cos θ − 10φ̇θ̇ sin θ = 0 (11.81)
4
Setting θ̇ = 0 for constant nutation rate in Equation (11.80) leads to

φ̈ sin θ = 0 (11.82)

from which we conclude that φ̈ = 0, so the precession rate φ̇ is constant (hence, the name
“steady precession”). Introducing this result into Equation (11.81) leads to ψ̈ = 0, which
Dynamics of Three-Dimensional Rigid Body Motion 581

indicates that the spin rate ψ̇ is also constant. Assume, without loss of generality, that
(φ̇ cos θ + ψ̇) > 0. The velocity of the disk can then be written
 
vG = −R φ̇ cos θ + ψ̇ j = −vi (11.83)

All the terms in the above equation are constant, so the speed of the center of the disk
v = R (φ̇ cos θ + ψ̇) is also constant. Furthermore, the radius of the path of the center of the
disk, denoted here by ρ, is also constant and it can be written in terms of the precession
rate as vG = −vi = ρφ̇i, from which the relationship between the radius of curvature to the
speed becomes
v
ρ = − (11.84)
φ̇
Combining Equation (11.83) and Equation (11.84), we can express the spin rate as
 ρ
ψ̇ = −φ̇ cos θ + (11.85)
R
The precession and spin rates have negative signs, indicating that they are in different
directions. This type of motion is called retrograde precession and is typical of the motion
of flat axisymmetric bodies.3
The steady motion of a disk can be described in terms of two parameters, the precession
and spin rates φ̇ and ψ̇, or in terms of the speed v and nutation angle θ. To explore
the relationship among v, ρ, and θ, introduce Equation (11.84) and Equation (11.85) into
Equation (11.79) with θ̈ = 0, with the result
 2 !
1 2 v v v
mR −6 sin θ − sin θ cos θ + mgR cos θ = 0 (11.86)
4 ρ R ρ

Solving for v 2 results in

4gρ2 cot θ
v2 = (11.87)
6ρ + R cos θ

It is instructive to analyze the no-slip assumption in more detail. To this end, use the
X 0 Y 0 Z 0 frame, which is obtained after the first Euler angle rotation. Figure 11.12 shows
the free-body diagram. The disk rolls in the X 0 Y 0 plane. Hence, the normal force N is in
the Z = Z 0 direction and the friction forces F1 and F2 are along the X 0 and Y 0 directions,
respectively. These two friction forces prevent two types of slip.
Slip along the X 0 = x direction is tangent to the path of the contact point. This point,
which coincides with the contact point C at all times, is denoted by C 0 . From Example 11.8,
the velocity of C 0 for no slip is calculated as
 ωy 
vC 0 = −R ωz − i = −Rψ̇i = −Rψ̇I0 (11.88)
tan θ
When slip occurs in the X 0 direction, part of the motion of the disk is to spin in place.
The disk may still travel forward. This type of slip is referred to as spin slip. An example
is someone moving a car from rest by depressing the accelerator fully, especially when on
a wet or icy road. If the friction between the car and road is not sufficient to prevent slip,
3 By contrast, the steady motion of a slender body, such as the spinning top, is in the form of direct

precession, where the precession and spin rates are in the same direction.
582 Applied Dynamics

Z, Z'
. y

z 

G Y'

F1 mg
F2
C Path of C'

N
X', x

FIGURE 11.12
Path of disk and free-body diagram.

the tires spin in place, as the vehicle gains speed very slowly. For spin slip to occur, a large
force or moment must be applied to the disk.
The second type of slip is along the Y 0 axis, and is very similar to the sliding of a body
on a surface. This type of slip is referred to as sliding slip or sliding. For a freely rolling
disk slipping will be of this type.
To examine whether slipping is occurring during steady motion, let us write the force
balance for the disk. Using the X 0 Y 0 Z 0 axes, the force balance is
maG = −F1 I0 + F2 J0 + (N − mg)K0 (11.89)
The expression for acceleration of the center of the disk is given in Equation (9.71) as
 ωy ωz 
aG = R (−ω̇z + ωx ωy ) i + R −ωx2 − j + R (ω̇x + ωy ωz ) k (11.90)
tan θ
The unit vectors in the shape frame (xyz) and in the X 0 Y 0 Z 0 frame are related by
i = I0 j = cos θJ0 + sin θK0 k = cos θK0 − sin θJ0 (11.91)
Introducing this relationship into Equation (11.90) and using the force balance, we can
obtain the values for the normal and friction forces. The resulting expressions are quite
lengthy, so this section considers the steady precession case. Noting that the velocity of the
center of mass for steady precession is vG = −vi = −vI0 , the acceleration is obtained as
aG = v̇Grel + ω f × vG = ω f × vG = φ̇K0 × −vI0 = −φ̇vJ0 (11.92)
v
From Equation 11.84, for steady motion −φ̇ = ρ so the acceleration of the center of the
disk for steady motion is simply
v2 0
aG = J (11.93)
ρ
Introducing the above equation into Equation (11.89), the external forces for steady
precession become
mv 2
F1 = 0 F2 = N = mg (11.94)
ρ
Dynamics of Three-Dimensional Rigid Body Motion 583

The above results are expected. In general, once rolling motion without spin slip begins,
the friction force F1 becomes zero. Because the nutation angle is not changing, the only force
that is encountered in the vertical direction is gravity, and the friction force encountered
in the lateral direction, F2 , has the same amplitude as the centrifugal force. The critical
condition for sliding is when the friction force attains its maximum value,

mv 2
F2 = µN = µmg = (11.95)
ρ

Using Equation (11.87), the friction coefficient needed to prevent sliding is related to
the radius of curvature and nutation angle by

v2 4ρ cot θ
µ = = (11.96)
ρg 6ρ + R cos θ
The expression for the friction coefficient in terms of v and θ becomes a quadratic
expression. Given v and θ, we first calculate ρ and plug in the above expression.
For the majority of rolling bodies, the radius of curvature ρ is much larger than the
radius R of the disk. We can then approximate the denominator of Equation (11.87) as
6ρ + R cos θ ≈ 6ρ, and use this approximation in Equation (11.87) and Equation (11.96)
with the result
4gρ2 cot θ 2 4ρ cot θ 2
v2 = ≈ ρg cot θ µ = ≈ cot θ (11.97)
6ρ + R cos θ 3 6ρ + R cos θ 3
We reiterate that the results obtained here are for the special case of steady precession,
also known as steady motion, and that the expressions for the friction forces are more
complicated for the general case of rolling.

Example 11.10
Consider the rolling disk in Example 9.4 with R = 12 cm, v = 5 m/s and θ = 60◦ , and
calculate the amount of friction necessary to prevent slip for steady precession.
The value of the radius of curvature in Example 9.4 was calculated as ρ = 6.648 m.
Plugging this into Equation (11.96) gives

v2 25
µ = = = 0.3835 [a]
ρg 6.648 × 9.807
As expected, the radius of curvature is much larger than the radius of the disk. Using
this property and Equation (11.97) to approximate the radius of curvature and the friction
coefficient gives the approximate values of

3v 2 tan θ 2 cot θ
ρ ≈ µ ≈ [b]
2g 3
Considering the parameters of this example, we obtain

3 × 25 × 3 2
ρ ≈ = 6.623 m µ ≈ √ = 0.3849 [c]
2 × 9.807 3 3
Both quantities are very close to their actual values.
584 Applied Dynamics

11.9 Rotation about a Fixed Axis


An important special case is rotation about a fixed axis. Such motion has a single degree
of freedom, requiring only one angular velocity component. Shafts transmitting power is
a typical example and so is a wheel on an axle. For an axisymmetric body, the motion
is smooth when the rotation axis coincides with the symmetry axis. However, a dynamic
imbalance will be present under the following circumstances:
• There is a mass imbalance in the body, or

• The body is axisymmetric, but is rotating about an axis that is not the symmetry axis.
The resulting motion is not smooth, with wobbling of the rotating body and of the
body to which it is connected, as well as fluctuations in the angular velocity. Such motion
is undesirable when transmitting motion or when moving objects, as loads on supports
increase and become uneven, resulting in vibration, noise, unnecessary wear, and damage.
In addition to the problems mentioned earlier, a car tire that is not balanced also leads to
reduced grip of the road surface, resulting in lateral and braking instabilities.
Consider the case where the rotation axis is not a principal axis, as in an unbalanced
shaft. Selecting the rotation axis, say, z, as one of the body axes, the products of inertia
Ixz and Iyz do not vanish. Writing components of the angular velocity as

ωx = ωy = 0 ωz = ω (11.98)

and substituting this into the rotational equations of motion, Equations (11.51), the moment
balances become

Mx = −Ixz ω̇ + Iyz ω 2 My = −Iyz ω̇ − Ixz ω 2 (11.99)

Mz = Izz ω̇ (11.100)

Equation (11.100) is the equation of motion with Mz as the externally applied moment.
On the other hand, Mx and My in Equation (11.99) are moments generated as a result of
the imbalance. These resultant moments are countered by reaction forces at the bearings
supporting the shaft, thus adding to the loads acting on the bearings. The resultant moments
Mx and My are affected by the rate of change of the angular velocity, as well as the square
of the angular velocity. Even when the angular velocity is constant, these moments are
not zero. Further, the magnitudes of these moments increase by the square of the angular
velocity, becoming more critical at high speeds.

Example 11.11
Consider the L-shaped rod in Figure 10.10, which is rotating with constant angular velocity
ω about a fixed base O. Using an xyz coordinate system attached to the rod, with the z
axis as the fixed axis of rotation, calculate the reactions at the base O. Then, calculate
these reactions in terms of a set of inertial axes XY Z. The xyz axes are obtained from the
inertial axes by rotating about the Z axis by θ, so that ω = θ̇.
From Example 10.3, the center of mass location is
 
1 4
rG = a i+ k [a]
6 3
Dynamics of Three-Dimensional Rigid Body Motion 585

and the inertia matrix about O is


 
20 0 −3
1
[IO ] = ma2  0 21 0 [b]
9
−3 0 1

Using Equations (11.99) and (11.100), and noting that the angular velocity is constant
and that Ixy = Iyz = 0, the moment equations become

1
Mx = Mz = 0 My = −Ixz ω 2 = − ma2 ω 2 [c]
3

–X
y y My y
My
2a
z  Y
O –Y

G Ox
Ox
a 

X x

FIGURE 11.13
Free-body diagram and orientation of xyz and XY Z axes.

The free-body diagram is shown in Figure 11.13. The acceleration of the center of mass
is due to the normal acceleration, as the angular velocity is constant, and it is
  
1 4 1
aG = ω × (ωω × rG ) = ωk × ωk × a i+ k = − aω 2 i [d]
6 3 6

so that, neglecting gravity, the force in the x direction becomes


X 1
Fx = max =⇒ Ox = − maω 2 [e]
6
To obtain the reaction forces and moments in the inertial frame, we note that the xyz
and XY Z coordinate systems are related by
    
x cos θ sin θ 0 X
 y  =  − sin θ cos θ 0  Y  [f ]
z 0 0 1 Z

and, considering that θ = ωt, the external forces and moments become

OX = Ox cos ωt OY = Ox sin ωt MX = −My sin ωt MY = My cos ωt [g]

with other components of external force and moment being zero.


586 Applied Dynamics

Y
y

z' B
A  G
Z, z
R
 x
X

FIGURE 11.14
A disk (tire) whose symmetry axis is not aligned with the rotation axis.

Example 11.12—Dynamic Tire Balancing


A purely static imbalance is in the form of an added (or reduced) mass on the symmetry
plane of a tire, and it moves the center of mass away from the center of the tire. The
resultant inertia matrix, as we saw in Example 10.4, is still diagonal. A dynamic imbalance,
on the other hand, introduces off-diagonal terms to the inertia matrix.
A commonly used dynamic imbalance model is a symmetric disk (tire) mounted on a
shaft so that the symmetry axis of the disk and the shaft are not aligned, with angle γ
between them, as shown in Figure 11.14. Assuming that the shaft is rotating with constant
angular velocity Ω, let us calculate the reactions at the bearings. In an actual application,
magnitudes of the bearing reactions are used to balance the tire.

My
y'
y


z'
 R
z
A G B

x, x' Mx
L/2 L/2

FIGURE 11.15
Side view of unaligned disk and acting moments.

Three coordinate systems describe the orientation. The XY Z coordinates are fixed, with
the Z axis along the axis of the shaft and the Y axis as vertical. The xyz axes are attached
to the shaft so that the angle θ between the X and x (and also between Y and y) axes is
θ = Ωt, as shown in Figure 11.14. The x0 y 0 z 0 axes are attached to the disk and are obtained
Dynamics of Three-Dimensional Rigid Body Motion 587

by rotating the xyz axes by γ about the −x axis, as shown in side view in Figure 11.15.
The moment about the z axis, Mz = 0, as we assume that the angular velocity Ω constant.
It is customary to use the xyz axes to write the moment balance equations, which
necessitates calculation of the inertia matrix components for the xyz frame. The yz plane
is the plane of symmetry, so Ixz = Ixy = 0. To find Iyz , we use the definition of the
product of inertia. Note that Ix0 x0 = Iy0 y0 = mR2 /4, Iz0 z0 = mR2 /2. Expressing the y and
z coordinates in terms of the primed axes as

y = y 0 cos γ + z 0 sin γ z = z 0 cos γ − y 0 sin γ [a]

the product of inertia Iyz is obtained as


Z Z
Iyz = yz dm = (y 0 cos γ + z 0 sin γ) (z 0 cos γ − y 0 sin γ) dm

1
= (Iy0 y0 − Iz0 z0 ) sin γ cos γ = − mR2 sin γ cos γ [b]
4
Introducing the value of Iyz from above and Ixz = 0 into Equation (11.99) yields the
resultant moments acting on the disk as
1
Mx = Iyz ω 2 = − mR2 Ω2 sin γ cos γ My = −Ixz ω 2 = 0 [c]
4

Y
Y y Y

F AY My F BY
B
Z A
G
F AX Mx
F BX
x X
X mg X
L/2 L/2

FIGURE 11.16
Free-body diagram of shaft.

Next, consider the free-body diagram of the shaft in Figure 11.16. The weight of the disk
acts on the shaft. The disk also exerts an equal and opposite moment −Mx on the shaft.
This moment needs to be expressed in terms of the inertial coordinates XY Z. Writing the
moment in vector form as M = −Mx i and noting that i = cos θI + sin θJ, components of
this moment in the X and Y directions become

MX = −Mx cos θ = −Mx cos Ωt MY = −Mx sin θ = −Mx sin Ωt [d]

These resultant moments are generated by the bearing forces at either end. Summing forces
in the X and Y directions gives

FAX = −FBX FAY + FBY = mg [e]

Summing moments about the center of mass of the shaft gives


L L
MX = (FBY − FAY ) MY = (FAX − FBX ) [f ]
2 2
588 Applied Dynamics

Introducing Equation [c] into Equation [d] and solving Equations [e] and [f] for the
reaction forces at the bearings, we obtain

1 R2 2
FBX = −FAX = − m Ω sin γ cos γ sin Ωt
4 L

1 1 R2 2
FAY = mg − m Ω sin γ cos γ cos Ωt
2 4 L
1 1 R2 2
FBY = mg + m Ω sin γ cos γ cos Ωt [g]
2 4 L
The misalignment between the shaft and the disk gives rise to bearing forces in both the
X and Y directions. These bearing forces are a function of the rotation speed Ω. Dynamic
wheel balancing involves adding counterweights to the disk (tire) so that the bearing forces
are only due to the weight. Knowing the mass m, radius R, and angular velocity Ω, we
measure the bearing forces and determines the value of sin γ cos γ by
2L
q
2 2
sin γ cos γ = (FBX − FAX ) + (FBY − FAY ) [h]
mR2 Ω2
from which we can calculate γ. Knowing from experience the amount of counterweight that
corresponds to a correction for γ, the balancer applies the counterweight, spins the tire
again, and observes the change in the bearing forces.
In actual tire balancing, the inaccuracies associated with the weight distribution are
more complex than the slanted disk model considered here.

11.10 Impulse and Momentum


As discussed in Chapter 5, impulse-momentum relationships describe the effects of forces
and moments on bodies over a period of time. These relationships are the same for two-
or three-dimensional motion. We will review these relationships here and discuss angular
impulse-momentum in more detail.
For linear momentum, we take Newton’s Second Law, F (t) = dp (t) /dt = m dvG /dt,
multiply this equation by dt, and integrate over time, which leads to the impulse-momentum
theorem as
Z t2 Z p(t2 )
F (t) dt = dp = p (t2 ) − p (t1 ) = mvG (t2 ) − mvG (t1 ) (11.101)
t1 p(t1 )

When the left side of the above equation is zero, that is, the integral of the applied force
over the time interval of interest is zero, the initial and final linear momenta of the system
are the same. This is known as the principle of conservation of linear momentum.
The angular impulse-momentum theorem is obtained by taking the general form of the
d
moment balance equation, MG = dt HG , and integrating it over time
Z t2 Z HG (t2 )
MG (t) dt = dHG = HG (t2 ) − HG (t1 ) (11.102)
t1 HG (t1 )

When the integral of the applied moment is zero over a range of time, the result is
conservation of angular momentum during that time period.
Dynamics of Three-Dimensional Rigid Body Motion 589

For three-dimensional motion, the angular impulse-momentum theorem can be expressed


in column vector format as
Z t2
[IG ] ({ω (t2 )} − {ω (t1 )}) = {MG (t)} dt (11.103)
t1

and, given initial conditions and the magnitude of the angular impulse, we can obtain the
final angular velocities from
Z t2
−1
{ω (t2 )} = {ω (t1 )} + [IG ] {MG (t)} dt (11.104)
t1
Rt
In the presence of an impulsive force or moment, we replace t12 F (t) dt with F̂ in
R t2
Equation (11.101) and t1 MG (t) dt with M̂G in Equation (11.102), keeping in mind that
the time interval involved is very short.

Example 11.13
Consider the rotating L-shaped rod in Figure 11.17 of mass = 2 kg and a = 45 cm. There
is a ball-and-socket joint at point O. The rod is initially rotating about the z axis with
angular velocity ω = 0.2 rad/s. The rod is hit at the tip P by a force F = 62 N in the y
direction, for a time period of ∆t = 0.001 seconds. Calculate the angular velocity of the rod
immediately after impact.

y
2a
z
O

P
x
F

FIGURE 11.17
Rod hit by an impulsive force.

The time period is small enough to consider the force as impulsive, so


F̂ ≈ F ∆t = 62 × 0.001 = 0.062 N · s [a]
and the resulting impulsive moment about point O is
M̂O = (ai + 2ak) × F̂ j = F̂ (−2ai + ak) = −0.0558i + 0.0279k N · m · s [b]
Applying Equation (11.104), we obtain the angular velocity after impact, denoted by
{ω 0 }, as
{ω 0 } = {ω} + [IO ]−1 {M̂O } [c]
where    
0 −0.0558
{ω} =  0  rad/s {M̂O } =  0  N·m·s [d]
0.20 0.0279
590 Applied Dynamics

From Equation [b] in Example 11.11, the inertia matrix about O is


 
20 0 −3
[IO ] = 0.045  0 21 0  kg · m2 [e]
−3 0 1

and the inverse of the inertia matrix is


 
2.0202 0 6.0606
1
[IO ]−1 =  0 1.0582 0  [f ]
kg · m2
6.0606 0 40.404

Introducing Equations [d] and [f] into Equation [c] gives the angular velocity immediately
after impact as
      
0 2.0202 0 6.0606 −0.0558 0.0564
{ω 0 } =  0  +  0 1.0582 0  0  =  0  rad/s [g]
0.20 6.0606 0 40.404 0.0279 0.9891

11.11 Energy and Work


This section discusses work and energy expressions for three-dimensional motion. Work
and energy expressions were first developed in Chapter 5. While the general form of these
expressions remains the same for any type of motion, some further treatment is needed for
three-dimensional analysis, especially for the kinetic energy.

11.11.1 Kinetic Energy


The kinetic energy of a body is defined as
Z
1
T = v · v dm (11.105)
2 body

where v is the absolute velocity of the differential element. Consider Figure 11.1 and express
the velocity of the differential element in terms of an arbitrary point on the body, say, B,
as

v = vB + ω × s (11.106)

Introducing this into Equation (11.105) and performing the dot product gives
Z Z
1 1
T = mvB · vB + ω × s) · (ω
(ω ω × s) dm + (vB · ω ) s dm (11.107)
2 2 body body

The first term on the right side of the above equationR involves a translation, while
the second term involves a rotation. The last term, (vB · ω ) body sdm, is a combination of
translational and rotational terms, and it involves the first moment of the mass distribution.
This term becomes zero if one of the following holds:
• vB = 0, implying that point B is fixed, and the rigid body is rotating about point B.
• ω = 0, implying that the body has no rotational motion.
Dynamics of Three-Dimensional Rigid Body Motion 591
R
• Point B coincides with the center of mass G, in which case s = ρ and body
sdm = 0.
It is preferable to express kinetic energy in terms of the center of mass. The first term
on the right side of Equation (11.107) gives the translational part of the kinetic energy,
denoted by Ttran , as
1
Ttran = mvG · vG (11.108)
2
The second term in Equation (11.107) is the rotational component of the kinetic energy,
denoted by Trot . The vector relationship

(a × b) · c = a · (b × c) (11.109)

can be used to evaluate this term. Letting a = ω , b = ρ , and c = ω × ρ results in


Z
1
Trot = ω · ρ × (ω
ω × ρ ) dm (11.110)
2 body
R
Recalling from Equation (11.6) the definition of angular momentum as HG = body ρ ×
ω × ρ )dm, the rotational kinetic energy can be expressed as

1
Trot = ω · HG (11.111)
2
Equation (11.111) can also be obtained using the column vector representation of Equa-
tion (11.110). Consider the relation

ω × ρ ) · (ω
(ω ω × ρ ) = (ρρ × ω ) · (ρρ × ω )

whose column vector representation is


T
[ρ̃] {ω})T ([ρ̃] {ω} = {ω}T [ρ̃] [ρ̃] {ω}

(11.112)

Recalling the column vector definition of inertia matrix, [IG ] = body [ρ̃]T [ρ̃]dm, the
R

column vector representation of the rotational kinetic energy is


Z
1 T 1 1
Trot = {ω}T [ρ̃] [ρ̃] {ω}dm = {ω}T [IG ] {ω} = {ω}T {HG } (11.113)
2 body 2 2

The total energy is T = Ttran + Trot . Table 11.2 summarizes expressions for kinetic
energy for different types of motion. Note that for the case of only translational motion, the
angular velocity is zero and all points on the body have the same velocity v.
Consider a body-fixed reference frame xyz. Writing the velocity of the center of mass
by vG = vx i + vy j + vz k and the angular velocity of the body as ω = ωx i + ωy j + ωz k, and
considering the inertia matrix [IG ], we can express the components of the kinetic energy as
1 T 1 1 T
m {vG } {vG } = m vx2 + vy2 + vz2

Ttran = Trot = {ω} [IG ] {ω}
2 2 2

1
Ixx ωx2 + Iyy ωy2 + Izz ωz2 − Ixy ωx ωy − Ixz ωx ωz − Iyz ωy ωz

= (11.114)
2
When the coordinate axes xyz are principal axes, the products of inertia vanish and the
rotational kinetic energy expression simplifies to
1 T 1
Ixx ωx2 + Iyy ωy2 + Izz ωz2

Trot = {ω} [IG ] {ω} = (11.115)
2 2
592 Applied Dynamics

TABLE 11.2
Kinetic energy expressions

Case Ttran Trot


1 1 T 1 1 T
General motion 2 mvG · vG = 2 m {vG } {vG } 2ω · HG = 2 {ω} [IG ] {ω}
1 1 T
Rot. about fixed C 0 2ω · HC = 2 {ω} [IC ] {ω}
1 1 T
Translation with v 2 mv ·v = 2 m {v} {v} 0

Let us write the rotational kinetic energy in terms of the Euler angles. Consider a 3-2-3
transformation from a1 a2 a3 to b1 b2 b3 . From Equation (9.38), the angular velocity expression
for a 3-2-3 transformation, with transformation angles ψ, θ, and φ, is
     
ω = −ψ̇sθcφ + θ̇sφ b1 + ψ̇sθsφ + θ̇cφ b2 + ψ̇cθ + φ̇ b3 (11.116)

where we have adopted the compact notation of s for sine and c for cosine. The rotational
kinetic energy becomes
1 T
Trot = {ω} [IG ] {ω}
2
  2 2 2 
1  
= I1 −ψ̇sθcφ + θ̇sφ + I2 ψ̇sθsφ + θ̇cφ + I3 ψ̇cθ + φ̇ (11.117)
2
For axisymmetric bodies, say, with I1 = I2 , the expression for rotational kinetic energy
simplifies. Setting I2 = I1 in the above equation gives
  2 
1 2 2 2
 
Trot = I1 θ̇ + ψ̇ s θ + I3 ψ̇cθ + φ̇ (11.118)
2
The sines and cosines of the third rotation angle, φ, are not present in the kinetic energy
for axisymmetric bodies, as well as in the angular velocity expression in terms of the shape
frame, so
 
ω = −ψ̇sθf1 + θ̇f2 + ψ̇cθ + φ̇ f3 (11.119)

11.11.2 Work and Conservation of Energy


The incremental work dW done by an applied force F over a displacement dr is

dW = F · dr (11.120)

where r describes the position of the point to which the force is applied. The work expression
has the form
Z r2
W = F · dr (11.121)
r1

For plane motion, the incremental work due to an applied moment has the form dW =
Dynamics of Three-Dimensional Rigid Body Motion 593

M dθ, where M is the applied moment and θ is the angle by which the body rotates.
Denoting the unit vector in the direction perpendicular to the plane of motion by k, the
moment becomes M = M k. The rotation angle is dθθ = dθk, so
dW = M · dθθ = M dθ (11.122)
For three-dimensional motion, angular velocity is not the derivative of a vector, but a
defined quantity. Hence, we cannot obtain a differential rotation by taking the derivative of
an expression. Rather, we define the differential rotation. The process involves taking the
expression for the angular velocity and changing the derivative terms in it to differentials.
For example, for the 3-2-3 transformation considered earlier, we replace ψ̇ with dψ, θ̇ with
dθ, and φ̇ with dφ, so the differential rotation term becomes
dθθ = (−dψsθcφ + dθsφ) b1 + (dψsθsφ + dθcφ) b2 + (dψcθ + dφ) b3 (11.123)
The incremental work due to an applied concentrated moment is
dW = M · dθθ (11.124)
As discussed earlier, it is customary to express the total effects of all the external ex-
citations as a resultant force F going through the center of mass and a resultant moment
MG about the center of mass. Hence, the incremental work done by all applied forces and
moments becomes
dW = F · drG + MG · dθθ (11.125)
Another way of obtaining the expression for work is to use the power, which is the rate
at which work is done, P = dW/dt. From Chapter 5, power is defined in terms of resultant
forces and moments as
P = F · vG + MG · ω (11.126)
Integrating the expression for power over time yields the work as
Z t2 Z t2
W1→2 = P dt = (F · vG + MG · ω ) dt (11.127)
t1 t1

Some of the forces and moments acting on a body may be conservative, that is, derivable
from a potential function. The work done on a body can be separated into work done by
conservative forces and by nonconservative forces as
dW = dWc + dWnc (11.128)
where the notation is obvious. Recall that for conservative forces, the differential work is
related to the derivative of the potential function (the potential energy) by dWc = −dV ,
where V is the potential energy. Total energy is defined as E = T + V , so the work-energy
theorem becomes
E (t1 ) + Wnc1→2 = E (t2 ) or T (t1 ) + V (t1 ) + Wnc1→2 = T (t2 ) + V (t2 ) (11.129)
where Wnc1→2 is the work done by nonconservative forces.
Chapter 5 discussed that spring forces and gravity are conservative forces. We can show
that an applied force that is constant can also be treated as conservative. So can a constant
moment for plane motion. However, for three-dimensional rotation, a constant moment
cannot be treated as conservative. This is because the three-dimensional motion equations
are always nonlinear, for all choices of the rotation variables.
Recall that, while for conservative natural systems the energy integral is E = T +
V = const, for conservative nonnatural systems the energy integral is the Jacobi integral
H = T2 − U = T2 − T0 + V = const.
594 Applied Dynamics

Example 11.14
Consider the rotating rod in Example 11.6. Calculate the kinetic and potential energies, as
well as the motion integral.
The angular velocity of the rod is

ω = ω1 b1 + ω2 b2 + ω3 b3 = −Ω cos θb1 + Ω sin θb2 + θ̇b3 [a]

The mass moments of inertia about the center of rotation B are


1
I1 ≈ 0 I2 = I3 = mL2 [b]
3
so the kinetic energy is
1 1  
I2 ω22 + ω32 = mL2 θ̇2 + Ω2 sin2 θ

T = Trot = [c]
2 6
Using point B as the datum, the potential energy has the form
mgL
V = − cos θ [d]
2
This is a conservative nonnatural system. The external forces and moments do not do
any work, as they are all reaction forces and moments. Hence, an integral of the motion is
the Jacobi integral
1   1
H = T2 − T0 + V = mL2 θ̇2 − Ω2 sin2 θ − mgL cos θ = const [e]
6 2
Because the system has one degree of freedom, we can obtain the equation of motion by
differentiating the Jacobi integral. Doing so, and dividing all terms by mL2 θ̇/3 yields
3 g
θ̈ − Ω2 sin θ cos θ + sin θ = 0 [f ]
2L
which is the same as Equation [h] of Example 11.6.

11.12 Analytical Equations for Rigid Bodies


The extended Hamilton’s principle and Lagrange’s equations discussed in Chapter 8, as well
as Kane’s equations, have the same form for two- or three-dimensional motion. This section
discusses the kinetic energy in terms of the Euler angles and an interpretation of Lagrange’s
equations. We also consider nonholonomic systems where use of Kane’s equations presents
an obvious advantage over traditional analytical mechanics.

11.12.1 Euler Angles as Generalized Coordinates


In Chapter 9 we saw that that a desirable set of coordinates to use when dealing with three-
dimensional motion is Euler’s angles. Using the Euler angles as generalized coordinates in
conjunction with Lagrange’s equations leads to interesting interpretations.
Consider the Euler angles of precession (ψ), nutation (θ), and spin (φ) associated with
a 3-2-3 Euler angle transformation from an A to a B frame. The angular velocity is
A
ω B = ψ̇a3 + θ̇a02 + φ̇b3
Dynamics of Three-Dimensional Rigid Body Motion 595
     
= −ψ̇sθcφ + θ̇sφ b1 + ψ̇sθsφ + θ̇cφ b2 + ψ̇cθ + φ̇ b3 (11.130)

The position of the center of mass is described using an inertial set of coordinates,
rG = A1 a1 + A2 a2 + A3 a3 (11.131)
so the velocity of the center of mass is vG = Ȧ1 a1 + Ȧ2 a2 + Ȧ3 a3
The kinetic energy then becomes
1  
T = Ttran + Trot = m Ȧ21 + Ȧ22 + Ȧ23 + Trot (11.132)
2
where Trot is defined in Equation (11.117).
The precession angle ψ is absent from the kinetic energy, so ψ can become a cyclic
coordinate (or ignorable coordinate) if it is not present in the potential energy and in the
virtual work. To see this better, write the generalized momentum associated with ψ, πψ , as
∂T ∂Trot
πψ = = (11.133)
∂ ψ̇ ∂ ψ̇
Recalling the definition of rotational kinetic energy in column vector format, we can write
the generalized momentum as
  T
∂ 1 T ∂ {ω}
πψ = {ω} [IG ] {ω} = [IG ] {ω}
∂ ψ̇ 2 ∂ ψ̇

∂ω1 ∂ω2 ∂ω3


= I1 ω1 + I2 ω2 + I3 ω3 (11.134)
∂ ψ̇ ∂ ψ̇ ∂ ψ̇
Differentiating the angular velocity vector with respect to the derivatives of the Euler
angles gives
∂ωω ω
∂ω ω
∂ω
= a3 = a02 = b3 (11.135)
∂ ψ̇ ∂ θ̇ ∂ φ̇
For a general Euler angle sequence (and not just for 3-2-3), with rotation angles ψ, θ, and
φ, define the unit vectors
∂ωω ω
∂ω ω
∂ω
eψ = eθ = eφ = (11.136)
∂ ψ̇ ∂ θ̇ ∂ φ̇
Comparing Equations (11.134)–(11.136), the generalized momenta are related to the
angular momentum by
πψ = eψ · HG πθ = eθ · HG πφ = eφ · HG (11.137)
Thus, generalized momenta associated with Euler angles are components of the angular
momentum along directions about which the Euler angle rotations have been performed.

11.12.2 Lagrange’s and Kane’s Equations


To write Lagrange’e equations, accumulate all external forces, conservative or not, in the
virtual work and write the virtual work as
n
X
δW = Qk δqk = Qψ δψ + Qθ δθ + Qφ δφ (11.138)
k=1
596 Applied Dynamics

where Qψ , Qθ , and Qφ are the generalized forces. The equations of motion associated with
the rotational coordinates become
d ∂T d ∂T d ∂T
πψ − = Qψ πθ − = Qθ πφ − = Qφ (11.139)
dt ∂ψ dt ∂θ dt ∂φ
It turns out that the Lagrange’s equations for the rotational motion of an unconstrained
rigid body are the angular momentum balances in the directions about which the Euler angle
transformations are made. The derivation is lengthy and is not given here. The interested
reader can consult the text by Baruh (pp. 460–462). We state below the major results from
the derivation. The generalized forces are

Qψ = eψ · MG Qθ = eθ · MG Qφ = eφ · MG (11.140)

where MG is the resultant moment about the center of mass. Also,


d ∂T d ∂T d ∂T
ḢG · eψ = πψ − ḢG · eθ = πθ − ḢG · eφ = πφ − (11.141)
dt ∂ψ dt ∂θ dt ∂φ
It follows that
d ∂T
πψ − = Qψ =⇒ ḢG · eψ = MG · eψ
dt ∂ψ

d ∂T
πθ − = Qθ =⇒ ḢG · eθ = MG · eθ
dt ∂θ

d ∂T
πφ − = Qφ =⇒ ḢG · eφ = MG · eφ (11.142)
dt ∂φ
The above equations are derivable using Kane’s equations as well. Selecting the gener-
alized speeds as u1 = ψ̇, u2 = θ̇, and u3 = φ̇, from Equation (11.136) the partial angular
velocities become
ω
∂ω ∂ωω ω
∂ω ω
∂ω
ω1 = = = eψ ω2 = = = eθ
∂u1 ∂ ψ̇ ∂u2 ∂ θ̇

ω
∂ω ω
∂ω
ω3 = = = eφ (11.143)
∂u3 ∂ φ̇
 
Applying Kane’s equations gives ḢG − MG · ω i = 0 (i = 1, 2, 3), or
   
ḢG − MG · ω 1 = ḢG − MG · eψ = 0

   
ḢG − MG · ω 2 = ḢG − MG · eθ = 0

   
ḢG − MG · ω 3 = ḢG − MG · eφ = 0 (11.144)

which are the same as Equation (11.142).


Similarly, we can also derive Euler’s equations (or the general rotational equations when
Dynamics of Three-Dimensional Rigid Body Motion 597

the coordinate axes are not principal axes) from Kane’s equations. Here, the generalized
speeds are selected as u1 = ω1 , u2 = ω2 , u3 = ω3 . Noting that the angular velocity is

ω = ω1 b1 + ω2 b2 + ω3 b3 (11.145)

the partial angular velocities become


ω
∂ω ω
∂ω ω
∂ω
ω1 = = b1 ω2 = = b2 ω3 = = b3 (11.146)
∂u1 ∂u2 ∂u3
Introduction of these partial velocities to the Kane’s equations gives the moment balance
equations, Equation (11.51), when the reference axes are not principal axes, and Equation
(11.52) when the coordinate axes are selected as the principal axes.

Example 11.15
Consider the rolling disk in Figure 9.16, which is rolling without sliding or slipping. Obtain
the equations of motion using a 3-1-3 Euler angle transformation and Kane’s equations.
The free-body diagram is shown in Figure 11.8. Rolling without slipping is a three-
degrees-of-freedom problem, and we select the generalized coordinates as the three Euler
angles of precession φ, nutation θ, and spin ψ. We also use the shape frame xyz.
There are two apparent choices for the generalized speeds: derivatives of the Euler angles
(φ̇, θ̇, ψ̇) or the angular velocity components in the xyz frame (ωx , ωy , ωz ). Selecting the
components of the velocity of the center of mass as generalized speeds is a valid but not a
very suitable choice and leads to complicated expressions.
Let us select the derivatives of the Euler angles as the generalized speeds so that

u1 = φ̇ u2 = θ̇ u3 = ψ̇ [a]

From Equation (9.62), the angular velocity is


 
ω = θ̇i + φ̇ sin θj + φ̇ cos θ + ψ̇ k = ωx i + ωy j + ωz k [b]

Roll without slip indicates that the velocity of the contact point C is zero, From this,
and noting that the vector from C to G is rG/C = Rj, we calculate the velocity of the center
of mass using  
vG = ω × rG/C = −R φ̇ cos θ + ψ̇ i + Rθ̇k [c]
The next step is to obtain the partial velocities, with the result

1 ∂vG 2 ∂vG 3 ∂vG


vG = = −R cos θi vG = = Rk vG = = −Ri
∂ φ̇ ∂ θ̇ ∂ ψ̇

ω
∂ω ω
∂ω ∂ωω
ω1 = = sin θj + cos θk ω2 = = i ω3 = = k [d]
∂ φ̇ ∂ θ̇ ∂ ψ̇
The only external force that does work is gravity, so that the resultant force and moment
are
F = −mgK = −mg cos θk − mg sin θj MG = 0 [e]
Kane’s (or Gibbs-Appell) equations are
 
j
(maG − F) · vG + ḢG − MG · ω j = 0 j = 1, 2, 3 [f ]

To implement Kane’s equations, it is necessary to calculate the acceleration of the center


598 Applied Dynamics

of mass and the rate of change of the angular momentum. The acceleration of the center of
mass is obtained by differentiating the velocity:

aG = v̇Grel + ω f × vG [g]

where ω f is the angular velocity of the shape frame, ω f = θ̇i + φ̇ sin θj + φ̇ cos θk. The
individual terms in Equation [g] become
 
v̇Grel = −R φ̈ cos θ − φ̇θ̇ sin θ + ψ̈ i + Rθ̈k [h]

     
ω f × vG = θ̇i + φ̇ sin θj + φ̇ cos θk × −R φ̇ cos θ + ψ̇ i + Rθ̇k
    
= Rφ̇θ̇ sin θi + −Rθ̇2 − Rφ̇ cos θ φ̇ cos θ + ψ̇ j + Rφ̇ sin θ φ̇ cos θ + ψ̇ k [i]

Using the above expressions, the translational parts of Kane’s equations become
 
1
(maG − F) · vG = (maG − F) · (−R cos θ)i = mR2 φ̈ cos θ − 2φ̇θ̇ sin θ + ψ̈ cos θ

  
2
(maG − F) · vG = (maG − F) · Rk = mR2 θ̈ + φ̇ sin θ φ̇ cos θ + ψ̇ + mgR cos θ
 
3
(maG − F) · vG = (maG − F) · (−R)i = mR2 φ̈ cos θ − 2φ̇θ̇ sin θ + ψ̈ [j]

The rotational terms are calculated next, beginning with the angular momentum. The
mass moments of inertia of the disk are
1 1
Ixx = Iyy = mR2 Izz = mR2 [k]
4 2
and the angular momentum is
1    
HG = Ixx ωx i + Iyy ωy i + Izz ωx k = mR2 θ̇i + φ̇ sin θj + 2 φ̇ cos θ + ψ̇ k [l]
4
Using the transport theorem, the rate of change of the angular momentum becomes

ḢG = ḢGrel + ω f × HG [m]

with the individual terms in Equation [m] having the form


1      
ḢGrel = mR2 θ̈i + φ̈ sin θ + φ̇θ̇ cos θ j + 2 φ̈ cos θ − φ̇θ̇ sin θ + ψ̈ k [n]
4
  1    
ω f × HG = θ̇i + φ̇ sin θj + φ̇ cos θk × mR2 θ̇i + φ̇ sin θj + 2 φ̇ cos θ + ψ̇ k
4
1     
= mR2 φ̇2 sin θ cos θ + 2φ̇ψ̇ sin θ i − φ̇θ̇ cos θ + 2θ̇ψ̇ j [o]
4
Noting that the resultant moment about the center of mass is zero, MG = 0, the
components of Kane’s equations involving the rotational terms become
 
ḢG − MG · ω 1 = ḢG · (sin θj + cos θk)
Dynamics of Three-Dimensional Rigid Body Motion 599
1  
= mR2 φ̈ sin2 θ + 2ψ̈ cos θ + 2φ̈ cos2 θ − 2θ̇ψ̇ sin θ − 2φ̇θ̇ sin θ cos θ
4
  1  
ḢG − MG · ω 2 = ḢG · i = mR2 θ̈ + 2φ̇ψ̇ sin θ + 2φ̇2 sin θ cos θ
4
  1  
ḢG − MG · ω 3 = ḢG · k = mR2 ψ̈ + φ̈ cos θ − φ̇θ̇ sin θ [p]
2
The last step is to add the translational and rotational components of Kane’s equations,
given in Equations [j] and [p], to obtain the equations of motion:
 
For u1 = φ̇: mR2 φ̈ cos2 θ − 2φ̇θ̇ sin θ cos θ + ψ̈ cos θ

1  
+ mR2 φ̈ sin2 θ + 2ψ̈ cos θ + 2φ̈ cos2 θ − 2θ̇ψ̇ sin θ − 2φ̇θ̇ sin θ cos θ
4
1     
= mR2 φ̈ sin θ − 2θ̇ψ̇ sin θ + 6ψ̈ + 6φ̈ cos θ − 10φ̇θ̇ sin θ cos θ = 0 [q]
4
  
For u2 = θ̇: mR2 θ̈ + φ̇ sin θ φ̇ cos θ + ψ̇ + mgR cos θ

1  
+ mR2 θ̈ + 2φ̇ψ̇ sin θ + φ̇2 sin θ cos θ
4
1  
= mR2 5θ̈ + 5φ̇2 sin θ cos θ + 6φ̇ψ̇ sin θ + mgR cos θ = 0 [r]
4
  1  
For u3 = ψ̇: mR2 φ̈ cos θ − 2φ̇θ̇ sin θ + ψ̈ + mR2 ψ̈ + φ̈ cos θ − φ̇θ̇ sin θ
2
1  
= mR2 6ψ̈ + 6φ̈ cos θ − 10φ̇θ̇ sin θ = 0 [s]
4
The terms inside brackets on the right side of Equation [s] also appear in Equation [q],
so that Equation [q] can be simplified to
1  
mR2 φ̈ sin θ − 2θ̇ψ̇ sin θ = 0 [t]
4
Hence, the equations of motion are Equations [r], [s], and [t].
Using Kane’s equations is cumbersome in that we have to obtain terms for the accelera-
tion of the center of mass and the rate of change of the angular momentum. This complexity
should be compared with the algebraic complexity of using Lagrange’s equations and con-
straints. Also, if we use the angular velocity components as the generalized speeds, the
resulting equations are simpler. We will demonstrate this in the example that follows.

Example 11.16
Consider the disk in the previous example that is rolling without sliding or slipping. Obtain
the equations of motion using a 3-1-3 Euler angle transformation and Kane’s equations. Use
the angular velocity components in the xyz frame ωx , ωy , ωz as the generalized speeds.
The free-body diagram is shown in Figure 11.8. There are three degrees of freedom and
the generalized coordinates are selected as the three Euler angles of precession φ, nutation
600 Applied Dynamics

θ, and spin ψ. As before, it is convenient to use the shape frame xyz. The generalized speeds
are
u1 = ωx u2 = ωy u3 = ωz [a]
From Equation (9.62) and Equation (9.63) the angular velocities of the disk and of the
shape frame are
ωy
ω = ωx i + ωy j + ωz k ω f = ωx i + ωy j + k [b]
tan θ
The velocity of the center of mass, given by Equation (9.70), is

vG = ω × rG/C = (ωx i + ωy j + ωz k) × Rj = −Rωz i + Rωx k [c]

The partial velocities and partial angular velocities are

1 ∂vG 2 ∂vG 3 ∂vG


vG = = Rk vG = = 0 vG = = −Ri
∂ωx ∂ωy ∂ωz

ω
∂ω ω
∂ω ω
∂ω
ω1 = = i ω2 = = j ω3 = = k [d]
∂ωx ∂ωy ∂ωz
The acceleration of the center of mass, as given in Equation (9.71), is
 ωy ωz 
aG = R (−ω̇z + ωx ωy ) i + R −ωx2 − j + R (ω̇x + ωy ωz ) k [e]
tan θ
From Equation [e] of Example 11.15, the resultant force and moment are F =
−mg cos θk − mg sin θj, MG = 0, so the components of Kane’s equations involving the
translational terms are
1
(maG − F) · vG = (maG − F) · Rk = mR2 (ω̇x + ωy ωz ) + mgR cos θ

2
(maG − F) · vG = 0

3
(maG − F) · vG = (maG − F) · (−R) i = mR2 (ω̇z − ωx ωy ) [f ]
The angular momentum about the center of mass is
1
HG = Ixx ωx i + Iyy ωy j + Izz ωz k = mR2 (ωx i + ωy j + 2ωz k) [g]
4
Using the transport theorem, the derivative of the angular momentum is

ḢG = ḢGrel + ω f × HG [h]

with the individual terms having the form


1
ḢGrel = mR2 (ω̇x i + ω̇y j + 2ω̇z k)
4
 ωy  1
ω f × HG = ωx i + ωy j + k × mR2 (ωx i + ωy j + 2ωz k)
tan θ 4
!
1 ωy2 1 ω ω
x y

= mR2 2ωy ωz − i + mR2 − 2ωx ωz j [i]
4 tan θ 4 tan θ
Dynamics of Three-Dimensional Rigid Body Motion 601

Noting that the resultant moment about the center of mass is zero, the components of
Kane’s equations involving the rotational terms are
!
 
1 1 2
ωy2
ḢG − MG · ω = ḢG · i = mR ω̇x + 2ωy ωz −
4 tan θ
  1  ωx ωy 
ḢG − MG · ω 2 = ḢG · j = mR2 ω̇y + − 2ωx ωz
4 tan θ
  1
ḢG − MG · ω 3 = ḢG · k = mR2 ω̇z [j]
2
The next step is to add the translational and rotational components of Kane’s equations,
given in Equations [f] and [j], to obtain the equations of motion, with the result
!
1 2
ωy2
mR 5ω̇x + 6ωy ωz − + mgR cos θ = 0
4 tan θ

1  ωx ωy  1
mR2 ω̇y + − 2ωx ωz = 0 mR2 (6ω̇z − 4ωx ωy ) = 0 [k]
4 tan θ 4
We can introduce the expressions for ωx , ωy , and ωz in terms of the Euler angles, so
the equations of motion become
1    
mR2 5θ̈ + 6φ̇ sin θ φ̇ cos θ + ψ̇ − φ̇2 sin θ cos θ + mgR cos θ = 0
4

1   1  
mR2 φ̈ sin θ − 2θ̇ψ̇ = 0 mR2 6ψ̈ + 6φ̈ cos θ − 10φ̇θ̇ sin θ = 0 [l]
4 4
which are, of course, the same as the equations of motion obtained in Example 11.15.
Comparing the algebraic effort with the previous example, using the angular velocity
components as generalized speeds leads to less cumbersome expressions.

Example 11.17
Consider the rod in Examples 11.6 and 11.14. Using an energy approach, find the moment
that needs to act on the pivot so that the rod rotates with a constant angular velocity.
We previously considered that the angular velocity of the pivot was given and a constant.
This resulted in a nonnatural system. Here, we will use the constraint relaxation method
from Chapter 8. The rotation of the rod will be treated as a variable φ̇ and a moment
M acts on the pivot to maintain the constant speed of the shaft. The equation of motion
associated with φ will be derived first, and then the constraint that φ̇ = constant will be
invoked. The kinetic energy is
1  1 1   1
T = Trot = I2 ω22 + ω32 + I φ̇2 = mL2 φ̇2 sin2 θ + θ̇2 + I φ̇2 [a]
2 2 6 2
where I is the mass moment of inertia of the rotating base to which the pivot is attached.
The virtual work is due to the moment M acting on the base to maintain the constant
rotation rate4 and has the form
δW = M δφ [b]
4 Such a moment can be generated by a servomotor.
602 Applied Dynamics

The rotation angle of the base, φ, does not contribute to the potential energy. Hence,
Lagrange’s equations associated with φ become5
 
d ∂T ∂T
− = Qφ [c]
dt ∂ φ̇ ∂φ

which, when expanded, gives


1 2
mL2 φ̈ sin2 θ + I φ̈ + mL2 φ̇θ̇ sin θ cos θ = M [d]
3 3

Next, invoke the constraint that φ̇ = Ω is a constant, so the expression for M becomes
2
M = mL2 Ωθ̇ sin θ cos θ [e]
3

Let us compare this answer with Equation [j] in Example 11.6, which is M2 = 2I2 Ωθ̇ cos θ.
From Figure 11.6 we recognize that

M = M2 sin θ [f ]

which confirms that the result here is the same as Equation [j] in Example 11.6.

11.13 Torque-Free Motion of Axisymmetric Bodies


An important application of Euler’s equations is for bodies that are not acted upon by any
external moments or when the resultant moment is zero.
Rotational motion of torque-free axisymmetric bodies is the topic of this section. The
results here are also valid for a body of arbitrary shape with two of its principal moments
of inertia the same (known as axially symmetric). The reader is referred to the texts by
Ginsberg or by Baruh for an analysis of torque-free motion of arbitrary bodies.

11.13.1 Integrals of the Motion


Consider an axisymmetric body, with I1 = I2 . Axisymmetry, as discussed in Sec. 11.7, not
only simplifies the equations of motion but also increases the stability properties. This is the
reason bodies that undergo substantial rotational motion are designed as axisymmetric; for
example, balls, rockets, and machinery parts. Using a body-fixed reference frame (not the
shape frame) and Euler’s equations in Equation (11.52), and setting the external moments
to zero gives

I1 ω̇1 − (I1 − I3 ) ω2 ω3 = 0 I1 ω̇2 − (I3 − I1 ) ω1 ω3 = 0 I3 ω̇3 = 0 (11.147)

The rightmost equation above, I3 ω̇3 = 0, indicates that ω3 is constant, so it is an integral


of the motion. Introducing the rotation constant
I3 − I1
Ω = ω3 (11.148)
I1
5 The equation of motion associated with θ is left as an exercise. See Problem 11.45.
Dynamics of Three-Dimensional Rigid Body Motion 603

the first two of Equations (11.147) can be expressed as

ω̇1 + Ωω2 = 0 ω̇2 − Ωω1 = 0 (11.149)

The above equations describe a gyroscopic system, with Ω describing the gyroscopic
properties, namely the rate at which the gyroscopic motion unfolds. Multiplying the left
equation by ω1 and the right equation by ω2 and adding the two equations leads to
1 d
ω12 + ω22 = 0

ω1 ω̇1 + ω2 ω̇2 = (11.150)
2 dt
which leads to the conclusion that ω12 + ω22 is constant. Hence, a second integral of the
motion is identified. The two first integrals can be combined to yield a third integral of the
motion

ω12 + ω22 + ω32 = |ω


ω |2 = constant (11.151)

indicating that the magnitude of the angular velocity vector is also constant.
Because the body is torque-free, both the angular momentum about the center of mass
and rotational kinetic energy are conserved, providing two other motion integrals:

HG = I1 ω1 b1 + I1 ω2 b2 + I3 ω3 b3 = constant

2Trot = ω · HG = I1 ω12 + I1 ω22 + I3 ω32 = constant (11.152)



 


 
 


 

FIGURE 11.18
Angular momentum vector and its projection onto the b1 b2 plane.

p The integrals of the motion help explain the motion of the body. Define by ω12 =
ω12 + ω22 the magnitude of the projection of the angular velocity vector on the b1 b2 plane,
as shown in Figure 11.18. Also define by H12 the projection of the angular momentum
vector on the b1 b2 plane, where
q
H12 = I1 ω12 + ω22 = I1 ω12 (11.153)

The above equations indicate that the projections of the angular momentum HG and
the angular velocity ω on the b1 b2 plane lie along the same line. Hence, as shown in Figure
604 Applied Dynamics










FIGURE 11.19
Common plane of the angular momentum and angular velocity vectors.

11.19, the vectors b3 , HG , and ω lie on the plane defined by the b3 and the b12 axes, where
the b12 axis is defined such that H12 = H12 b12 and ω 12 = ω12 b12 .
The motion of the body can be interpreted as the movement of the plane formed by the b3
and b12 axes. This plane rotates with angular velocity Ω. The rotation is about the angular
momentum vector, as both the magnitude and the direction of the angular momentum are
constant. While the magnitude of the angular velocity is constant, its direction changes,
and the angular velocity vector can be viewed as rotating around the angular momentum
vector and around the b3 axis. The angles α and β in Figure 11.19 are constants, and they
can be expressed as
H12 I1 ω12 ω12
tan α = = tan β = (11.154)
H3 I3 ω3 ω3
in which H3 = HG · b3 = I3 ω3 is the component of the angular momentum about the b3
axis. The relationship between α and β depends on the magnitudes of the mass moments
of inertia I1 and I3 . There are two possible cases:
• If I1 < I3 , then α < β. This case corresponds to the motion of a flat body, such as a
disk or a frisbee.

• If I1 > I3 , then α > β. This case corresponds to the motion of a slender body, such as
a rocket, a rod, or American football.

11.13.2 Body and Space Cones


The rotation of the angular velocity vector defines the body and space cones. These cones
roll on top of each other. The body cone is fixed to the body. It is generated by rotation
of the angular velocity vector as viewed from the b3 axis (the spin and symmetry axis).
The space cone is fixed in inertial space and is generated by the rotation of the angular
velocity vector about the angular momentum vector. Because the body we are dealing with
is axisymmetric, the body and space cones are right circular cones.
Figure 11.20 shows the body and space cones for a flat body (I1 < I3 ) and for a slender
body (I1 > I3 ). The space cone has an apex angle of |β − α|. The space cone usually is
easier to visualize than the body cone, as it is generated by the angular velocity vector
rotating about a fixed axis. For a flat body, the space cone rolls inside the body cone and
Dynamics of Three-Dimensional Rigid Body Motion 605

for a slender body the body and space cones roll outside of each other. Visualizing the body
and space cones is usually easier for slender bodies.

a) Flat body b) Slender body


b3 Space b3
. cone Fixed . Fixed
 
HG HG
. .
  
 
Body
 Space
  cone
Body cone
cone b12 b12

FIGURE 11.20
Body and space cones for a) a flat body, b) a slender body.

Consider, for example, a football being thrown. Under ideal circumstances and for a
perfectly thrown ball, the ball would have a spin only about its longitudinal axis, implying
an angular velocity vector that is constant in direction as well as magnitude, as shown
in Figure 11.21a. If the ball is tipped, or is thrown with a wobble, the angular velocity
has two components, a wobble and a spin as in Figure 11.21b. Consequently, the angular
velocity no longer has a constant direction, and the angular velocity vector rotates itself.
The magnitude of the wobble remains constant.

%' b! (' b!

v ! v !
G

#$#%&
b" " b"

FIGURE 11.21
Football in motion: a) well-thrown ball, b) ball with a wobble.

Another example of wobbling motion is the rotation of the Earth. The Earth is not a
perfect sphere—it is an oblate spheroid. Attaching a set of body axes to the Earth and
selecting the b3 axis as the polar axis, from measurements that have been taken, we get
I3 − I1
= 0.0033 (11.155)
I1
Taking the angular velocity of the Earth as 1 rev/day leads to Ω = 0.0033 rev/day. This
implies that the Earth should sweep one body cone in 1/0.0033 = 300 days.
The polar axis of the Earth indeed has some wobbling motion, a phenomenon called
variation in latitude or Chandler wobble. The period of the wobble has been measured
as 433 days and the radius swept by the poles in one body cone as about 4 meters. The
606 Applied Dynamics

difference from the predicted period of 300 days is attributed to two reasons; The Earth is
not rigid and it is not entirely torque-free. Because the Earth is not a perfect sphere, the
gravitational forces between it and the moon and the sun create a gravity gradient torque.
Also, the lack of rigidity of the Earth’s crust and mantle causes some energy transfer from
one component of the rotation to the other.6
An interesting interpretation of the body and space cones is obtained when considering
a 3-1-3 (or 3-2-3) transformation to locate the body and when the direction of the angular
momentum with the inertial is the a3 axis. Comparing Figure 11.20 with Figure 9.14, the
conclusion is that the nutation angle θ in Figure 9.14 is the same as α in Figure 11.20.

11.13.3 Direct and Retrograde Precession


Consider a 3-1-3 Euler angle sequence with the rotation angles of φ (precession), θ (nuta-
tion), and ψ (spin). The associated rotation matrices are
   
cφ sφ 0 1 0 0
[R1 ] =  −sφ cφ 0  [R2 ] =  0 cθ sθ 
0 0 1 0 −sθ cθ

 
cψ sψ 0
[R3 ] =  −sψ cψ 0 (11.156)
0 0 1

with the resulting transformation matrix as


 
cφcψ − sφcθsψ sφcψ + cφcθsψ sθsψ
[R] = [R3 ] [R2 ] [R1 ] =  −cφsψ − sφcθcψ −sφsψ + cφcθcψ sθcψ  (11.157)
sφsθ −cφsθ cθ

and the relationship between the coordinates of the rotated B frame is related to the
coordinates of the initial A frame by
   
b1 a1
 b2  = [R]  a2  (11.158)
b3 a3

Recall that the a3 direction is selected such that the angular momentum, which is
constant, is written as HG = HG a3 . The angular momentum vector can be expressed in the
B frame by expressing a3 in terms of the B frame. This requires multiplying the third row
of [R]T (or the third column of [R]) by [b1 b2 b3 ]T , which leads to

HG = HG a3 = HG (sin θ sin ψb1 + sin θ cos ψb2 + cos θb3 )

= I1 ω1 b1 + I1 ω2 b2 + I3 ω3 b3 = H1 b1 + H2 b2 + H3 b3 (11.159)

from which we obtain

H1 = I1 ω1 = HG sin θ sin ψ H2 = I1 ω2 = HG sin θ cos ψ

H3 = I3 ω3 = HG cos θ (11.160)
6 In recent years, severe earthquakes in the Pacific Ocean have slightly altered the mass moments of

inertia of the Earth, as well as the orientation of the axis about which the Earth revolves.
Dynamics of Three-Dimensional Rigid Body Motion 607

Notice from the last of these equations that because ω3 is constant, so too is the nutation
angle θ. Also θ = α in Equation (11.154), leading to

I3 ω3
cos θ = = constant (11.161)
HG
To obtain expressions for the precession and spin rates we can make use of the relation-
ship between the body-fixed angular velocities and the Euler angles. The angular velocity
for a 3-1-3 transformation has the form

ω = ω1 b1 + ω2 b2 + ω3 b3 = φ̇a3 + θ̇a01 + ψ̇b3 (11.162)

Setting the nutation rate θ̇ = 0 and expressing a3 in terms of the B frame by reading
T
from the third column of [R] (third row of [R] ) gives the angular velocities in terms of the
Euler angles as

ω1 = φ̇ sin θ sin ψ ω2 = φ̇ sin θ cos ψ ω3 = φ̇ cos θ + ψ̇ (11.163)

Comparing the above equation with Equation (11.159), and considering Equation (11.161)
leads to the conclusion that
HG I 3 ω3
φ̇ = = = constant ψ̇ = −φ̇ cos θ + ω3 = constant (11.164)
I1 I1 cos θ
The two terms in the above equation can be combined with the result
I3
φ̇ = ψ̇ (11.165)
(I1 − I3 ) cos θ

This equation states that for torque-free motion of an axisymmetric body, the precession
rate and spin rate are proportional to each other. For example, if a football is thrown with a
wobble, the magnitude, direction, and angle of the wobble remain the same while the ball is
traveling.7 Two cases are possible, depending on the magnitudes of the moments of inertia:

• When I1 > I3 , such as in a slender body, the precession and spin have the same direction.
This type of motion is called direct precession. The space cone rolls outside the body
cone, as shown in Figure 11.20b. The wobble is in the same direction as the spin.
• When I1 < I3 , such as in a flat body, the precession and spin have the opposite directions.
This type of motion is called retrograde precession. The space cone rolls inside the body
cone, as shown in Figure 11.20a. Recall that the motion of a rolling disk in Section 11.8
is in the form of retrograde precession, even though the motion there is not torque-free.

Direct precession is easier to observe than retrograde precession.

11.13.4 Energy Dissipation and Nutational Instability


We have so far considered cases where both angular momentum and kinetic energy are
conserved. This is not the case in actual applications. Energy can be lost due to internal
or external forces, such as friction and damping, as well as sloshing of fluids in aircraft and
spacecraft. Energy loss initiates nutational motion, as was discovered to be in the case of the
Explorer satellite. Motion that begins with the body rotating about the axis of minimum
7 In reality, the wobble changes due to aerodynamic effects and because the football is not completely

axisymmetric, as the laces create a small amount of asymmetry.


608 Applied Dynamics

moment of inertia eventually acquires nutation and ends up with the body rotating about
the axis of maximum moment of inertia.
The process is illustrated in Figure 11.22 for an axisymmetric body, where θ is the
nutation angle. Denoting the initial angular velocity by ωi , in Figure 11.22a the angular
momentum is HG = I3 ωi a3 . Nutational motion begins, as shown in Figure 11.22b. When
the nutation angle becomes θ = 90◦ (Figure 11.22c), the angular momentum is
HG = I3 ωi a3 = I1 ωf a3 (11.166)
so
I3
ωf = ωi (11.167)
I1
Since for a slender body I3 /I1 < 1, it follows that ωf < ωi . For example, for I3 /I1 = 1/3,
so that ωf = ωi /3. The initial and final kinetic energies are
1 1 1  ω 2 1
i
Ti = I3 ωi2 Tf = I1 ωf2 = 3I3 = Ti (11.168)
2 2 2 3 3

a) b) c) b3
b3

a3, b3 2 
 a3 a3
i f
12
b12

FIGURE 11.22
a) Initial motion about smallest inertia axis, b) nutational motion begins, c) rotation about
largest inertia axis.

Section 11.7 discussed the stability of the rotational motion and showed that for an
arbitrary body, rotations about the minimum and maximum axes of inertia are stable. For
an axisymmetric body, rotation about the axis of symmetry is stable, regardless of whether
the body is squat or slender. These results, obtained from linearized equations, assume no
energy loss. It was observed in 1958 that energy dissipated in the Explorer 3 satellite caused
nutational instabilities. The satellite decayed from orbit after only 93 days.
The following example demonstrates that rotation about the axis with the smallest
moment of inertia is unstable in the presence of energy loss.

Example 11.18—Nutational Instability due to Energy Loss


Consider an axisymmetric body, such as the squat and slender bodies shown in Figure 11.23.
The centroidal mass moments of inertia of the body are denoted by Ixx = Iyy and Izz , where
z is the symmetry axis. The angular velocities are ωx , ωy , ωz . Consider torque-free rotation
and assume that the body has no translational motion. The angular momentum is
HG = Ixx (ωx i + ωy j) + Izz ωz k [a]
and the kinetic energy is purely rotational
1
Ixx ωx2 + ωy2 + Izz ωz2
 
T = Trot = [b]
2
Dynamics of Three-Dimensional Rigid Body Motion 609

z x

FIGURE 11.23
A squat and a slender body.

The magnitude of the angular momentum is


2 2
ωx2 + ωy2 + Izz
2
ωz2

HG = Ixx [c]
2
and it is also constant. Consider the expression HG − 2T Ixx which can be written as
2
HG − 2T Ixx = Izz (Izz − Ixx ) ωz2 [d]

As discussed in the previous section, the nutation angle θ = α is constant for torque-free
motion. From Equation (11.161) we can write
2
Izz ωz2 = HG
2
cos2 θ [e]

so that Equation [d] can be rewritten as


 
2 Ixx
2T Ixx = HG − Izz (Izz − Ixx ) ωz2 = 2
HG − 2
HG 2
cos θ 1 − [f ]
Izz

Differentiating Equation [f] with respect to time, dividing the resulting expression by
2Ixx and rearranging gives
2
 
HG Izz
Ṫ = cos θ sin θ − 1 θ̇ [g]
Izz Ixx

Now, consider that there is some form of energy dissipation so that Ṫ < 0. Depending
on the shape of the body, two scenarios are possible:
• For a flat body, such as a frisbee, where Izz > Ixx , the nutation rate is negative, θ̇ < 0.
The nutation angle becomes smaller with time and the body eventually acquires simple
rotation about the z axis, albeit with a smaller angular velocity. A flat body spun with
a nutation angle eventually loses its nutational motion.
• For a slender body, as in a rod or rocket, where Ixx > Izz , the nutation rate is positive,
θ̇ > 0, so the nutation angle grows with time. The spin is eventually transferred to a
wobble about the transverse x axis. The slender body that should be rotating about its
symmetry axis no longer does and stability is lost.

Example 11.19
An American football has dimensions (both NFL and NCAA) of major axis 2a = 11–
11.25 in. (or longitudinal circumference 27.5 to 28.5 in.), minor axis 2c = 6.73–6.85 in. (or
circumference 20.75–21.25 in.), and weight 14–15 oz. The football should be pressurized to
610 Applied Dynamics

b1

G b3
a
c
F^

b2

FIGURE 11.24
Football (ellipsoid) that is tipped.

12.5 to 13.5 psi. The ball is initially spinning with angular velocity Ω about its symmetry
axis, when it is tipped by a player in a transverse direction. The tipping creates a moment
of magnitude I1 Ω. Obtain the precession and spin rates after the impulse is applied.
An approximate model of a football is that of a prolate ellipsoid, shown in Figure 11.24.8
The mass moments of inertia of a solid ellipsoid are
1 2
m a 2 + c2 mc2

I1 = I2 = I3 = [a]
5 5
with the mass of the ellipsoid as m = 4πρac2 /3. We can use this relationship and average
dimensions to obtain the mass and mass moment of inertia of an ellipsoidal shell of thickness
t as
m = m (a + t, c + t) I = I (a + t, c + t) − I (a, c) [b]
The procedure is tedious. Here, experimental results (Brancazio) are quoted for a football
of the dimensions given above, which are

I1 = 0.00321 kg · m2 I3 = 0.00194 kg · m2 [c]

Using these experimental results, the mass moment of inertia ratio becomes I1 /I3 = 1.655,
or I3 /I1 = 0.6044.
The angular momentum before the impulse is HG = I3 Ωb3 , and the angular momentum
after the impulse is
HG = I1 Ωb1 + I3 Ωb3 [d]
so the angular velocity after the impulse is

ω = Ωb1 + Ωb3 [e]

The motion of the ball after the tip is torque-free, with magnitude of the angular mo-
mentum as
p p p
HG = HG · HG = I3 Ω 1 + 1.6552 2 = I3 Ω 1 + 2.7392 [f ]
8 The difference in shape at the tips of the football does not affect the mass moments of inertia too much,

but it changes the aerodynamic properties.


Dynamics of Three-Dimensional Rigid Body Motion 611

Let us construct the body and space cones. The football is a slender body, so the motion
is in the form of direct precession. The cone angles are
H12 I1 Ω ω12 Ω
tan α = tan θ = = = 1.655 tan β = = =  [g]
H3 I3 Ω ω3 Ω
The body and space cones depend on the magnitude of the impulse. From Equations
(11.164)-(11.165) the precession and spin rates become
√ √
I3 Ω 1 + 3.2692 Ω 1 + 2.7392
 
HG I3
φ̇ = = = ψ̇ = ω3 1 − = 0.3956Ω [h]
I1 1.655I3 1.655 I1

The spin rate after the impulse ψ̇ is independent of the magnitude of the tipping impulse.
On the other hand, the nutation angle and precession rate depend on the strength of the
impulse. For small values of , the precession rate divided by the initial angular velocity,
which is also known as the wobble to spin ratio, becomes

φ̇ 1 + 2.7392 I3 I3
= ≈ (1 + 1.3692 ) ≈ [i]
Ω 1.655 I1 I1
indicating that for small wobbles (generated due to an impulsive transverse force or a
football that is thrown with a wobble), the wobble to spin ratio remains almost constant.
Because for the football I3 /I1 ≈ 3/5, this wobble to spin ratio is relatively easy to observe.

11.14 Bibliography
Baruh, H., Analytical Dynamics, McGraw-Hill, 1999.
Brancazio, P.J., “Rigid Body Dynamics of a Football,” American Journal of Physics, Vol.
55, 1987, pp. 415–420.
Ginsberg, J., Engineering Dynamics, 3rd Edition, Cambridge University Press, 2007.
Greenberg, M.D., Advanced Engineering Mathematics, 2nd Edition, Prentice-Hall, 1998.
Greenwood, D.T., Principles of Dynamics, 2nd Edition, Prentice-Hall, 1988.
Sidi, M.J., Spacecraft Dynamics and Control, Cambridge University Press, 1997.

11.15 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 11.2—Linear and Angular Momentum


11.1 (M) A solid rectangular block of mass 2 kg with sides a = 30 cm in the x direction,
b = 8 cm in the y direction, and d = 40 cm in the in the z direction rotates with angular
velocity ω about an axis OB as shown in Figure 11.25a. Point O is not fixed. Using the xyz
coordinates, calculate the angular momentum about the center of mass.
11.2 (E) The solid cylindrical rod of mass m, length L, and radius R in Figure 11.25b rotates
about the x axis with constant angular velocity ω, where θ is the angle between the fixed
612 Applied Dynamics

a) y b)
y 
O x
x
d
G L 
G X
m
B
b
 Radius R
A C
a
z

FIGURE 11.25
Figures for a) Problem 11.1, b) Problem 11.2.

g
O

L
y
R x
G
A L/2
m

FIGURE 11.26
Figure for Problems 11.3, 11.4, and 11.10.

X axis and x axis. Given that θ (t) = sin 2t, calculate the angular momentum about the
center of mass in terms of the xyz coordinates.
11.3 (M) The disk of mass m and radius R in Figure 11.26 is rotating with constant angular
velocity Ω about a light elbow. The elbow swings in the xy plane with angle θ. Calculate
the linear momentum and angular momentum about the center of mass G of the disk when
θ = 30◦ and θ̇ = Ω/2.

Section 11.3—Transformation Properties of Angular Momentum


11.4 (M) The disk of mass m and radius R in Figure 11.26 is rotating with constant angular
velocity Ω about a light elbow. The elbow swings in the xy plane with angle θ. Calculate
the angular momentum of the disk about the pivot O when θ = 30◦ and θ̇ = Ω/2.
11.5 (M) Consider the disk in Figure 11.27a spinning with ω. The pivot arm of length
L rotates with angular velocity Ω. Calculate the angular momentum about O. Show that
Equation (11.24) cannot be used to calculate the angular momentum about point O and
that we have to use Equation (11.22) to calculate HO . Explain why.
11.6 (M) A disk of mass m and radius R in Figure 11.27b rotates freely about a rod of
Dynamics of Three-Dimensional Rigid Body Motion 613

a) z b)

m g
O
R
O G y
M
L
  m
L G R
x

C
N

FIGURE 11.27
Figures for a) Problem 11.5, b) Problems 11.6 and 11.31.

mass M and length L. As the rod rotates about O with Ω, the disk rolls without slipping.
Calculate the angular momentum of the system about point O for when θ = 30◦ .

a1

a B b2
g

O
 L
G

m
b1

FIGURE 11.28
Figure for Problems 11.7, 11.13, 11.14, 11.35, and 11.40.

11.7 (M) The slender bar of mass m and length L, shown in Figure 11.28, is attached to
the arm of a shaft of length a that is rotating with a constant speed of Ω. Find the angular
momentum of the bar about point B.
11.8 (M) Calculate the angular momentum of the rolling cone in Figure 9.20 about the pivot
point, using the shape frame coordinate axes.
11.9 (M) Calculate the angular momentum of the rolling disk in Figure 9.43 about point O.

Sections 11.4 and 11.5—General Describing Equations, Body-Fixed Coordinates


11.10 (M) The disk of mass m and radius R in Figure 11.26 is rotating with constant angular
velocity Ω about a light elbow. The elbow swings in the xy plane with angle θ. Calculate
the gyroscopic moment in terms of θ and θ̇.
614 Applied Dynamics
a
a) b)
y 
z

a
O y v
b m x
L O L/2


L
x

FIGURE 11.29
Figure for Problem 11.11. a) Front view. b) top view.

11.11 (M) The propeller blade in Figure 11.29 is modeled as a thin rectangular prism of
length L and sides a and b, pivoted at its center. The blade is rotating with angular velocity
ω while the aircraft is traveling with speed v in the xy plane and taking a turn with radius
of curvature ρ. There is a damping moment Md = −cωi acting on the blade due to the
aerodynamics. Calculate the moment that is exerted on the pivot as a result of the turning
of the plane and aerodynamic moment.

Z, z
z'

m
D
L/5
G
L/5
0

L

x
O 

x'

FIGURE 11.30
Figure for Problem 11.12.

11.12 (M) The acrobat in Figure 11.30 of mass m can be modeled as a slender rod connected
to a ball-and-socket joint O. The acrobat is rotating with constant angular velocity ω = ω0 K
and is maintaining the angular velocity and constant lean angle θ = 15◦ by holding on to
the pole. Calculate the reaction forces at joint O. The force between the acrobat’s hands
and pole is in the horizontal (x) direction.
11.13 (M) The slender bar of mass m and length L, shown in Figure 11.28, is attached to
the arm of a shaft of length a that is rotating with a constant speed of Ω. Find the equation
of motion of the bar using Equations (11.43)–(11.46).
11.14 (M) The slender bar of mass m and length L, shown in Figure 11.28, is attached to
Dynamics of Three-Dimensional Rigid Body Motion 615

the arm of a shaft of length a that is rotating with a constant speed of Ω. Find the equation
of motion of the bar using Equation (11.40).

Section 11.6—Angular Momentum Balance for Axisymmetric Bodies


11.15 (M) Obtain the equation of motion of the disk rolling without slipping in Figure 9.18
by using the Euler angles described in Section 9.5.2.
11.16 (M) Consider the unicycle in Figure 9.42 and roll without slip. Obtain the equations
of motion assuming that the rider remains upright at all times (θ = 0). Select the Euler
angle transformation of your choice.
11.17 (D) Consider the rolling cone in Figure 9.20. Pivot O is fixed. It is given that friction
is sufficient to prevent slipping. The cone will tip if its rolling speed exceeds a critical value.
Find this critical speed in terms of v and L, for when R/L = 0.2. Hint: Use the shape frame
and model the normal force acting on the cone as a single force acting at a distance d from
O and relate the angular velocity to d.

Section 11.7—Stability Analysis of Rotational Motion

b"
$% b" &%
b"
N" N"
M M!
N! M M! N!
 G b!
G b! G b!
M#

b# M#
N# N#

b# b#

FIGURE 11.31
Figure for Problems 11.18, 11.19, 11.20, and 11.44. a) Gyrostat, b) free-body diagrams.

11.18 (D) A gyrostat is a spacecraft that has a rotor attached inside it. The rotor is placed in
the gyrostat for providing additional stability and for maneuvering (rotating) the spacecraft.
Consider the gyrostat in Figure 11.31a that is inside a body with principal moments of
inertia I1 , I2 = I, and I3 along the b1 b2 b3 axes. External moments N1 , N2 , and N3 act
on the spacecraft. The disk has mass moments of inertia J1 and J2 = J3 . The free-body
diagrams are shown in Figure 11.31b. Obtain the equations of motion of the spacecraft,
using the reference frame attached to the body as the shape frame. Neglect any translational
motion. A servomotor keeps the angular velocity of the gyrostat constant.
11.19 (M) Consider the gyrostat in Problem 11.18. Assume initially that the rotor is spinning
with Ω and the spacecraft does not have any angular motion. A small perturbation is
applied and the spacecraft acquires angular velocities of 1 , 2 , and 3 about b1 , b2 , and
b3 , respectively. Linearize the equations of motion and conduct a stability analysis on the
motion. Can the rotor stabilize the spacecraft if b1 is the intermediate axis of inertia (that
is, it is unstable when the rotor does not rotate)?
11.20 (M) Consider the gyrostat in Figure 11.31a which now is inside an axially symmetric
616 Applied Dynamics

body with principal moments of inertia I1 , I2 = I1 , and I3 along the b1 b2 b3 axes. Obtain
the equations of motion of the spacecraft, using the reference frame attached to the body as
the shape frame. Neglect any translational motion. A servomotor keeps the angular velocity
of the gyrostat constant.
11.21 (C) Consider a rectangular box of dimensions a, b, and c, in the body-fixed x, y, and
z directions, respectively. Conduct a parametric study of the level of instability (amount
of wobbling) as a function of the dimensions of the container. Let b = 1 and let the initial
rotation be about the y axis with Ω = 1. Vary the other dimensions as 0 ≤ a ≤ 5, 0 ≤ c ≤ 4
and create a three-dimensional plot of the positive root (when it exists) of the characteristic
equation.

Section 11.8—Steady Precession of a Rolling Disk


11.22 (C) How slanted can a disk be while it is rolling? Calculate and plot the speed v and
friction coefficient µs that is needed to keep a disk rolling as a function of the nutation
angle θ.

Section 11.9—Rotation about a Fixed Axis

a) g b)
y g
x m/10
R = L/2 A L/4 M
G 2L C m B
A D

m, 6L 2L z
B L/4 L/2 L/4
xyz fixed to shaft

FIGURE 11.32
Figures for a) Problem 11.23, b) Problem 11.24.

11.23 (M) The rod in Figure 11.32a of mass m, length 6L, and radius R = L is rotating
with constant angular velocity Ω. The imbalance in the shaft is modeled by a concentrated
mass m/10 that is at a distance d = 2L from the axis of the shaft and distance 2L from the
bearing at B. Calculate the bearing reactions at A and B due to this imbalance.
11.24 (M) The triangular plate of mass m in Figure 11.32b is attached to a massless rod.
The rod spins with constant angular velocity Ω. Find the reactions at the supports A and B
and the moment M that needs to be exerted on the shaft to maintain the constant angular
velocity.

Section 11.10—Impulse and Momentum


11.25 (M) A rigid body with the inertia matrix shown below is rotating with angular velocity
ω = 3i − 2k rad/s when it is subjected to an impulsive moment of M̂G = 23i − 10j N·m·s.
Calculate the angular velocity of the body immediately after impact.
 
12 0 3
[IG ] =  0 6 −4  kg · m2
3 −4 7
Dynamics of Three-Dimensional Rigid Body Motion 617

z, Z g

G y, Y
Q
 90 – 
P
 r

x, X F

FIGURE 11.33
Figure for Problem 11.26.

11.26 (M) A soccer player kicks the ball of radius 4.4 in. and weight 1 lb at point P ,
where θ = 60◦ and φ = 15◦ as shown in Figure 11.33. The applied impulsive force is
F̂ = −0.6i + 0.8k lb·sec. Calculate the translational and angular velocity of the ball after
the kick. Neglecting aerodynamics and the effects of rotation on the ball, how far does the
ball travel before it hits the ground?

a) g b)

O X y
y m
12" B
0.5 m
x
G x
Y 24" M
1m x
0.2 m
z
F B z
z 4"
Y
Z

FIGURE 11.34
Figures for a) Problem 11.27, b) Problem 11.28.

11.27 (M) The rectangular plate of weight 12 lb in Figure 11.34a is at rest with one of its
edges connected to a spherical joint. The plate is hit by a force F = 500 lb in the negative
Y direction for a period of 0.01 seconds. Calculate the angular velocity of the plate and the
velocity of point B after this impulsive force is applied.
11.28 (M) The slab of mass M = 10 kg in Figure 11.34b is tumbling freely in space when a
rock of mass m = 0.5 kg traveling with speed of 12 m/s in the −y direction hits it. Upon
impact, the rock gets lodged in the slab. Given that the angular velocity of the slab before
618 Applied Dynamics

impact is ω = 0.1i + 0.2j − 0.05k rad/s, calculate the angular velocity of the slab after
impact.

y
m/80
v
m v0 x
 G O

FIGURE 11.35
Figure for Problem 11.29.

11.29 (M) The spacecraft in Figure 11.35 is of mass m shaped as a cylindrical rod of length
L and radius R = L/6. It is moving with velocity of v0 in the x direction and also spinning
about the x axis with angular velocity ω = v0 /50L. A meteorite of mass m/80 traveling
with velocity v = −v0 /3j + v0 /5k impacts the spacecraft and miraculously gets lodged onto
point O, without damaging the spacecraft. Calculate the angular velocity and velocity of
point O immediately after impact.

Section 11.11—Energy and Work


11.30 (M) Calculate the kinetic energy lost in Problem 11.29 due to the impact.
11.31 (M) A disk of mass m and radius R in Figure 11.27b rotates freely about a rod of
mass M and length L. As the rod rotates about O with ω, the disk rolls without slipping.
Calculate the kinetic energy of the system.

#$ L %$

y y Y
z .
x
L!" 
F vA R x
 A
A m x
M D
L  X
R

FIGURE 11.36
Figure for Problems 11.32 and 11.37. a) General view, b) top view.

11.32 (M) The two wheels of mass m each in Figure 11.36a and radius R are mounted on
an axle of mass M and length L. The top view is shown in Figure 11.36b. Assume that the
wheels roll without slipping or sliding. Calculate the kinetic energy of the system.
11.33 (M) The two wheels in Figure 11.37a, each of mass m and radius R, are mounted on
an axle of mass M and length L. The side view is shown in Figure 11.37b. The wheels roll
Dynamics of Three-Dimensional Rigid Body Motion 619

B
M!"&'(L!) z
#$ L %$ B
M!"&'(L!)
y 
z
x
L!"
F vA R
kT
kT m x
vA
A M D
L A
R R

FIGURE 11.37
Figure for Problems 11.33 and 11.47. a) General view, b) side view.

without slipping or sliding. Attached to the axle is a rod of mass M/2 and length 3L/4,
which pivots about A about the y axis. The rod is stabilized by means of a torsional spring
of constant kT . Calculate the kinetic energy and potential energy of the system.

Y y
g
a B x
M
a
s
s
l M, L
e
s
s

O m
 R
G

FIGURE 11.38
Figure for Problem 11.34.

11.34 (M) The disk of mass m and radius R in Figure 11.38 rotates with spin rate ω at
the end of a pivoting rod of mass M and length L. The base rotates with constant angular
velocity Ω. Calculate a) the angular momentum of the disk about point B, b) the kinetic
energy of the system. The shaft to which the rod is attached is massless.

Section 11.12—Analytical Equations for Rigid Bodies


11.35 (M) The slender bar of mass m and length L, shown in Figure 11.28, is attached to
the arm of a shaft of length a that is rotating with a constant speed of Ω. Find the equation
of motion of the bar using Lagrangian mechanics.
11.36 (M) Elbow OAB in Figure 11.39 has radius R/3 and rotates about pin joint O with
620 Applied Dynamics

g
O

5L
s
L R x
B
A R/3

4L
y
z

FIGURE 11.39
Figure for Problems 11.36 and 11.46.

angle θ (t) = π6 sin 2t. A collar of mass m, radius R and length L slides over AB (variable
s denotes the slide) without friction and also rotates around AB with angle φ. The collar
does not rotate with respect to AB. Assuming the link is massless, obtain the equation of
motion of the collar.
11.37 (M) The two wheels of mass m each in Figure 11.36a and radius R are mounted on
an axle of mass M and length L. The top view is shown in Figure 11.36b. The wheels roll
without slipping or sliding. A force F = Fx i + Fy j acts at the center of the axle. Use Kane’s
equations to obtain the equations of motion.
11.38 (M) Consider the unicycle and rider in Figure 9.42. The mass of the unicycle is m
and the mass of the rider is 8m. Using the shape frame associated with a 3-2-3 Euler angle
transformation and assuming roll without slip, obtain the equations of motion using Kane’s
equations, as well as the kinematic differential equations. Assume that the rider does not
move with respect to the shape frame of the wheel. Select the generalized speeds as the
angular velocities ωi , (i = 1, 2, 3) as expressed in the shape frame.
11.39 (D) Consider the unicycle and rider in Figure 9.42. The mass of the unicycle is m and
the mass of the rider is 8m. Here, the rider moves with respect to the unicycle shape frame.
Obtain the equations of motion using Kane’s equations.
11.40 (M) Consider the rotating pendulum in Figure 11.28 and calculate the equilibrium
positions for when a) a = 0, b) a 6= 0.
11.41 (M) Obtain the equation of motion of the spinning top in Figure 11.9 by means of
the Lagrange’s equations.
11.42 (M) Identify the integrals of the motion of the spinning top in Figure 11.9 for when
the applied moment M = 0.
11.43 (M) Obtain the equation of motion of the spinning top in Figure 11.9 by means of
the Kane’s equations using a) the derivatives of the Euler angles, and b) the body angular
velocities as the generalized speeds.
11.44 (M) Obtain the equations of motion of the gyrostat in Figure 11.31 using Kane’s
equations. The angular velocity of the rotor Ω is not constant. Neglect translational motion.
11.45 (M) Consider the rotating rod in Example 11.6. Obtain the equation of motion using
the constraint relaxation method of Section 8.11.2. Treat Ω as a variable φ̇, derive the
equations of motion and then impose the constraint that φ̇ = const.
11.46 (D) Consider the collar in Problem 11.36. Now, the elbow angle θ is a motion variable
Dynamics of Three-Dimensional Rigid Body Motion 621

and the elbow is of mass M . Obtain the equations of motion for x, θ and φ using Lagrange’s
equations.
11.47 (D) The two wheels of mass m each in Figure 11.37a and radius R are mounted on
an axle of mass M and length L. The side view is shown in Figure 11.37b. The wheels roll
without slipping or sliding. Attached to the axle is a rod of mass M/2 and length 3L/4
which pivots about A about the y axis. The rod is stabilized by means of a torsional spring
of constant kT . A force F = Fx i + Fy j acts at the center of the axle. Obtain the equations
of motion using Kane’s equations.

Section 11.13—Torque-Free Motion of Axisymmetric Bodies


11.48 (E) Consider the tipped football in Example 11.19 and calculate the strength of the
impulse, as measured by the parameter  so that the nutation angle becomes θ = 25◦ . Find
the corresponding wobble to spin ratio.
11.49 (M) A slender axisymmetric body with I1 = I2 and I3 < I1 is given an initial
spin ω0 about the longitudinal axis (b3 ). Due to energy loss, the slender body acquires a
nutation angle of θ, at which point the energy loss ends and θ becomes constant. Writing
the angular momentum of the body as HG = I1 ω12 b12 + I3 ω3 b3 , show that ω3 = ω0 cos θ,
ω12 = I3 ω0 sin θ/I1 .
11.50 (M) A slender axisymmetric body with I1 = I2 = 3I3 is given an initial spin ω0 about
the longitudinal axis (a3 ). Due to energy loss, the slender body acquires a nutation angle
of θ = 15◦ , which then remains constant. Calculate the precession and spin rates and the
percentage of energy that is lost.
11.51 (M) A slender axisymmetric body with I1 = I2 = 2.4I3 is given an initial spin ω0
about the longitudinal axis (a3 ). Due to external effects, 20% of the energy gets lost after
which the body exhibits torque-free motion. Calculate the resulting nutation angle, as well
as precession and spin rates.

y' y

x'
 x
^
M
z

FIGURE 11.40
Figure for Problem 11.52.

11.52 (M) The fighter jet in Figure 11.40 is symmetric about the xy plane. The principal
axes of the aircraft are denoted by the x0 y 0 z axes and we assume that Iy0 y0 = Izz (enabling
us to treat the principal axes as that of an axisymmetric body). Also, Ix0 x0 = 0.4Izz . The
fighter jet is in steady flight when it is given an impulsive roll moment M̂ about the x axis
(roll axis of aircraft). Calculate the angular velocity of the aircraft and sketch the body and
space cones. The motion that ensues is called roll coupling and it led to loss of control of
high-performance fighter jets before its cause was understood.
622 Applied Dynamics


12
Vehicle Dynamics—Basic Loads and Longitudinal
Motions

12.1 Introduction
This and the following three chapters apply concepts and developments of the previous
chapters to the motion of ground vehicles. We study forces that act on vehicles, as well as
the ensuing motions. A vehicle can accelerate, decelerate, bounce up and down, or acquire
an angular velocity, such as yaw, pitch, or roll. While the general motion of a vehicle is
three-dimensional, we can, for an initial study, separate these motions and analyze them
independently.
This chapter discusses longitudinal motions of vehicles and presents a simple model
of roll. Coordinate systems used for describing vehicle motions are introduced. Tire and
aerodynamic forces are discussed in Chapter 13. Lateral motions (sideslip, yaw) are discussed
in Chapter 14 and pitch, bounce, and roll motions in Chapter 15.
A four-wheeled vehicle is statically indeterminate. The wheel loads cannot be determined
by the laws of statics or dynamics alone. We develop an approximate technique to calculate
wheel loads as well as center of mass location.
It is particularly important in vehicle dynamic analysis to know the units of the terms
with which we are dealing, as well as the positive directions associated with the motion
variables and the coordinate systems being used.

12.2 Vehicle Coordinate Systems and Nomenclature


A commonly used coordinate system in vehicle dynamics is the one adopted by the SAE
(Society of Automotive Engineers in the U.S.) and it is shown in Figure 12.1. It consists
of an inertial coordinate system defined by the XY Z coordinates and a set of body-fixed
coordinates xyz. The reference frame defined by xyz coordinates move with the vehicle.
The XY plane denotes the horizontal, and the Z axis is vertical, with the positive direction
as downward. For the body-fixed coordinates, the x direction is the longitudinal direction of
the vehicle, y direction is the lateral, and z is perpendicular to the xy plane of the vehicle,
positive downwards.
In a well-designed and well-built vehicle, the center of mass G is equidistant from the
left and right sides of the vehicle, and the plane formed by x and z axes that goes through
the center of mass G becomes the plane of symmetry of the vehicle.
For a stationary vehicle on level ground the xyz and XY Z coordinates are aligned. When
the vehicle is moving, the body-fixed coordinates are obtained by rotating the inertial XY Z
coordinates by a 3-2-1 rotation sequence by three angles:

623
624 Applied Dynamics

  






FIGURE 12.1
SAE vehicle coordinate system.

a) Top view b) Side view c) Rear view


X'' 
 X 
 X'
 Y''
X' 
 y
Y' Y  
Z'' z
Z' Z''

FIGURE 12.2
Vehicle body coordinates after a 3-2-1 rotation sequence: a) top view of rotation by heading
angle ψ about Z, b) side view of rotation by pitch angle θ about Y 0 , c) rear view of roll
rotation by φ about X 00 .

• A rotation about the Z axis, shown in Figure 12.2a, by the heading angle ψ, resulting
in the coordinate system X 0 Y 0 Z 0 with Z 0 = Z.
• A rotation about the resulting Y 0 axis, shown in Figure 12.2b, by the pitch angle θ,
resulting in the coordinate system X 00 Y 00 Z 00 with Y 00 = Y 0 .
• A rotation about the X 00 axis, shown in Figure 12.2c, by the roll angle φ, resulting in
the coordinate system xyz with x = X 00 .
The angular velocity of the vehicle is ω = ωx i + ωy j + ωz k. ωx is called roll, ωy is pitch,
and ωz is yaw. The translation of the vehicle in the longitudinal direction is called travel,
in the lateral (y) direction is called sideslip, and translational motion in the z direction is
called bounce.
Three-dimensional rotations and the orientation of bodies is discussed in Chapters 2
and 9. The order of the rotations needs to be specified, as the same rotations made in a
different sequence result in different orientations. In vehicle dynamics, the rotation angles
are usually are small, and we can assume that the order of the rotations is insignificant.
The study of vehicle motions in this text, for the most part, analyzes the rotational mo-
tions pitch, roll, and yaw independently of each other. For example, for the lateral motion
Vehicle Dynamics—Basic Loads and Longitudinal Motions 625

(XY plane) the only orientation angle considered is the heading angle ψ. Figure 12.3 illus-
trates the coordinates of interest. The angular velocity of the vehicle for two-dimensional
motion is ω = ωz k = ψ̇k.

X  x'
x
X
 x

Y

G Y y


FIGURE 12.3
Top view of vehicle coordinates for plane motion.

The coordinate systems used for tire and aerodynamic forces have the same orientation
as the body-fixed xyz coordinates. It is convenient to specify a reference point for tire and
aerodynamic forces. The resultant of the tire and aerodynamic forces is represented by a
force that acts through these reference points and a moment.

12.3 Loads on Vehicles


A vehicle is acted upon by forces that are generated externally as well as internally. Forces
that act on the vehicle can be summarized as follows:
• Weight (W = mg). The weight of a vehicle acts through the center of mass. In the
inertial frame, weight acts in the Z direction. On an inclined or banked road, the weight
has components in the lateral (y) or longitudinal (x) directions.

• Tire forces. Tire forces include wheel loads, tractive and lateral forces, as well as rolling
resistance forces. The magnitude of tire forces depends on the nature of the road surface,
friction, tire flexibility, and hysteresis.
• Tire moments. Tire moments are generated because the tire force distributions are
not uniform in the area where the tires make contact with the round, known as the
contact patch.
• Aerodynamic forces and moments. These forces are due to the interaction between
the vehicle and surrounding air and they are functions of the relative air speed.
• Excitations generated by components of the vehicle. These forces are generated
due to rotation of the engine, transmission components, as well as tires. Imbalances
626 Applied Dynamics

associated with these rotations lead to periodic forces that adversely affect the vehicle
by causing vibrations, faster wear of tires, loss of stability, and discomfort to occupants.
To propel a land vehicle, the driver steps on the accelerator. This action opens the
throttle and allows for a mixture of air and fuel to enter the engine. The engine turns more
rapidly and generates the power needed to drive the vehicle. This power is transmitted to
the wheels through a transmission system.

12.4 Acceleration

v
b x
c G
g B
FRR 
W
h  F Tf Z z
A z Wf

F Tr
Wr
L

FIGURE 12.4
Vehicle moving uphill.

Figure 12.4 shows the free-body diagram of the side view of a vehicle that is moving uphill
along a straight line. For the basic acceleration model, we ignore tire flexibility, suspension
systems, and the inertia of the tires. On level ground, the z direction is the vertical. The
grade angle of an inclined road is denoted by θ. The following forces act on the vehicle:
• Weight W and axle loads (normal forces) Wf and Wr . The subscripts f and r denote
front and rear, respectively.
• Tractive forces FTf and FTr . The tractive forces are in the x direction for acceleration
(FAf , FAr ) and in the −x direction for braking (FBf , FBr ). Tractive forces are generated
between the tires and the road as the consequences of the driving or braking torques at
the wheels.
The dynamics of an accelerating wheel is discussed in Chapter 5. The torque generated
at the axle is clockwise and, in the absence of friction, it gives the contact point between
the wheel and the road a velocity in the negative x direction. Friction between the tire
and the road fights this torque and generates a tractive force in the positive x direction.
The reverse occurs for braking.
• Rolling resistance forces, denoted by FRR , and aerodynamic forces. It is custom-
ary to show the cumulative rolling resistance force as a horizontal force along the road
surface acting on the front tire. Rolling resistance and aerodynamic forces will be dis-
cussed in detail in the next chapter. In this initial state of the analysis, aerodynamic
forces will not be included in the formulation.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 627

12.4.1 Acceleration Analysis


Consider first a rear wheel drive vehicle that is accelerating and set FTf = 0 and FTr = FAr .
Using the free-body diagram in Figure 12.4, the sum of forces in the x and z directions gives
X
+% Fx = maGx =⇒ FAr − mg sin θ − FRR = max (12.1)
X
+& Fz = 0 =⇒ −Wr − Wf + mg cos θ = 0 (12.2)
Assuming that there is no pitch angular velocity or acceleration, summing moments
about the center of mass gives
X
MG = 0 =⇒ Wf b − Wr c + (FAr − FRR ) h = 0 (12.3)
where h is the height of the center of mass from the ground. Equation (12.1) gives the
tractive force necessary to provide acceleration and to overcome the horizontal components
of gravity and rolling resistance as
FAr = max + mg sin θ + FRR (12.4)
Equations (12.2) and (12.3) can be solved together to find the axle loads, with the result
b h b h
Wr = mg cos θ + (FAr − FRR ) = mg cos θ + (max + mg sin θ) (12.5)
L L L L
c h c h
Wf = mg cos θ − (FAr − FRR ) = mg cos θ − (max + mg sin θ) (12.6)
L L L L
The first term on the right in each of the expressions above denotes the static axle loads
b c
Wrs = mg cos θ Wfs = mg cos θ (12.7)
L L
and the second and third terms denote the weight shift due to acceleration, max h/L, and
shift due to the incline, mg sin θh/L. The following observations can be made from Equations
(12.5)–(12.6):
• When a vehicle accelerates, the tractive force is positive, and it has the effect of shifting
weight to the rear wheels.
• The rolling resistance force reduces the acceleration and hence shifts weight to the front
wheels.
• When a vehicle decelerates, weight shifts to the front wheels. This a major reason why
the front brakes are designed to be larger than the rear brakes and why they usually
wear out faster than the rear brakes.
• When a vehicle moves uphill, weight shifts to the rear wheels, as the tractive force is
increased by mg sin θ.
• When a vehicle moves downhill, weight shifts to the front wheels. The front brakes work
even harder when decelerating a vehicle that is going downhill.
Note that for low inclinations θ, we can use a small angles approximation of sin θ ≈ θ
and approximate the effect of the incline on the acceleration as
a = −g sin θ ≈ −gθ (12.8)

Indeed, road inclinations rarely exceed 9 . Freeways are generally designed so that the
maximum inclination is 5◦ . Portions of freeways with higher inclinations usually have an
extra lane that is designated for traffic that cannot move uphill at posted highway speeds.
628 Applied Dynamics

12.4.2 Maximum Acceleration


An interesting special case of acceleration analysis is the maximum acceleration that can
be attained, assuming friction is large enough to sustain the tractive forces needed for such
acceleration. Let us ignore rolling resistance and consider a rear wheel drive vehicle so that
FAf = 0. Equation (12.6) indicates that as the vehicle accelerates, the front axle load Wf
becomes smaller. The limiting case is when the front wheels lose contact with the ground
and Wf = 0. Solving for the tractive force from Equation (12.6) gives
c
FAr = mg cos θ (12.9)
h
Setting Wf = 0 in the vertical force balance gives Wr = mg cos θ. Substitution of this
expression into the above equation leads to a relationship for the tractive force in terms of
the axle load as
c
FAr = Wr (12.10)
h
The maximum value of the tractive force is limited by the static friction coefficient, µs ,
so the maximum value of the tractive force is FAr = µs Wr , which leads to the conclusion
that the static friction coefficient must have a value of
c
µs ≥ (12.11)
h
in order to reach the maximum possible acceleration. Assuming there exists this much
friction and ignoring rolling resistance, we can calculate the maximum acceleration by sub-
stituting the value of the tractive force into Equation (12.1), with the result
c c 
maxmax = mg cos θ − mg sin θ =⇒ axmax = g cos θ − sin θ (12.12)
h h
Let us analyze the above equation for some realistic numbers. Assuming a level surface
(θ = 0), c/L ≈ 0.6, and h/L ≈ 1/4 for a passenger vehicle gives the result c/h ≈ 2.5,
so amax = µs g = 2.5g, which is a high number. The friction coefficient for passenger tires
is around µs ≈ 0.85. Hence, maximum possible acceleration cannot be reached with most
vehicles. An exception is drag cars.
Consider next finding the maximum acceleration for a given amount of friction µs . The
maximum value of the tractive force is FArmax = µs Wr and, substituting this relationship
into Equation (12.3), we obtain the front axle load in terms of the rear axle load as
c − µs h
Wf = Wr (12.13)
b
Introduction of the above relationship into the force balance in the z direction, Equation
(12.2), and solving for the rear axle load Wr lead to the result
b b
Wr = mg cos θ FAr = µs Wr = µs mg cos θ (12.14)
L − µs h L − µs h
Introducing Equation (12.14) to the force balance in the x direction and dividing by the
mass gives the maximum acceleration as
b
axmax = µs g cos θ − g sin θ (12.15)
L − µs h
c
Note that setting µs = h in the above equation, and recalling from Equation (12.11)
Vehicle Dynamics—Basic Loads and Longitudinal Motions 629

that this is the amount of friction needed for maximum acceleration in the previous case,
the expression for maximum acceleration becomes
b c
axmax = µs g cos θ c − g sin θ = g cos θ − g sin θ (12.16)
L − hh h

which is the same result as Equation (12.12).


Next, consider the maximum incline a rear wheel drive vehicle can climb. At maximum
incline, the acceleration of the vehicle is zero. Setting axmax = 0 in Equation (12.15) and
solving for the incline angle give the result
!
  b
µs b µ s
θmax = tan−1 = tan−1 L
h
(12.17)
L − µs h 1 − µs L

The results here are for a rear wheel drive vehicle. Calculation of the maximum acceler-
ation for a front wheel drive vehicle is left as an exercise. We can show that the maximum
acceleration of a front wheel drive vehicle is less than the maximum acceleration of a rear
wheel drive vehicle. This is a major reason why race cars are designed with rear wheel drive.
Front wheel drive is preferred for vehicles whose power to weight ratio is not large and when
the vehicle is expected to navigate low-friction conditions (icy roads, rain, off-road).

Example 12.1
Calculate the maximum incline that a rear wheel drive vehicle with Lb = 0.55, L
h
= 0.2, and
µs = 0.8 can climb and the maximum acceleration it can attain on a level road.
Substituting these values into Equation (12.17) results in
!
µs Lb
 
−1 −1 0.8 × 0.55
θmax = tan h
= tan = tan−1 (0.5238) = 27.65◦ [a]
1 − µs L 1 − 0.8 × 0.2

On a level road (θ = 0), the maximum acceleration is obtained from Equation (12.15)
as
axmax b/L 0.8 × 0.55
= µs = = 0.5238 [b]
g 1 − µs h/L 1 − 0.8 × 0.2
Let us compare this result with the maximum acceleration that can be attained if infinite
friction is available. For a level road, Equation (12.12) becomes

axmax c c/L 0.45


= = = = 2.25 [c]
g h h/L 0.2

Example 12.2—Drag Racer Design


Drag racing cars are designed to reach high speeds in extremely short periods of time. Zero
to 300 mph in 4.5 seconds corresponds to acceleration levels higher than what a fighter jet
experiences when catapulted from an aircraft carrier. Figure 12.5 shows the configuration
of a drag racer. The drag racer is designed so that most of the weight of the vehicle is
supported by the rear wheels.
Consider Equation (12.15), which gives the maximum acceleration of the drag racer for a
given amount of friction. Let us consider level ground and write the maximum acceleration
as
b/L
amax = µs g h
[a]
1 − µs L
630 Applied Dynamics

G
h
FT
A B
c b
Wr W
Wf

FIGURE 12.5
Drag racer.

One way of increasing the maximum acceleration is to increase b, the distance from the
front axle to the center of mass. Concentrating the weight of the vehicle in the rear helps.
Another way is to increase friction, which serves two purposes: it increases the numerator
and decreases the denominator.
Yet another way is to give the drag racer a higher center of mass, so that the term
h
1 − µs L will be as small as possible. However, raising the center of mass height has sev-
eral disadvantages, including reduced stability for lateral and roll motions. Most dragsters
achieve higher levels of acceleration by bringing the center of mass as close to the rear wheel
as possible and by using rear tires that have extremely high coefficients of friction.

12.5 Power
The amount of power that can be generated by the engine, the power that can be transmitted
to the driving wheels, and how the transmitted power translates into a tractive force are of
utmost importance in the study of vehicles.
As discussed in Chapter 5, power is the rate at which work is done, or
P = dW/dt (12.18)
where P is the power and W is the work done by a force or a torque applied to a body.
Consider the particle in Figure 12.6 moving under the influence of a force F, the incremental
work done by moving the particle from a point described by the vector r to a point r + dr
is defined by
dW = F · dr (12.19)
Dividing the above equation by dt and noting that dr/dt = v, where v is the velocity, leads
to
dW
P = = F·v (12.20)
dt
For rotational motion on a plane and for an object with centroidal mass moment of
inertia IG and angular velocity ω that is acted upon by a torque T , the expression for
power becomes
P = IG ω (12.21)
Vehicle Dynamics—Basic Loads and Longitudinal Motions 631

F
Tangent line
dr

FIGURE 12.6
Force F acting on a body.

The unit of power is force times velocity, or energy over time. The base unit of power
in the U.S. system is ft·lb/sec. In the SI system, the base unit of power is Watts (W)
where 1 W = 1 J/sec = 1 N·m/sec, where J stands for Joule, with 1 Joule = 1 N·m. It is
customary to express power in terms of horsepower (hp). There are five different definitions
of horsepower: mechanical, metric, electrical, boiler, and hydraulic. The most commonly
used definition is mechanical power, where 1 hp = 550 ft·lb/sec = 3300 ft·lb/min.
For rapidly rotating systems encountered in vehicles (engine, tires), it is customary to
express the angular velocity in terms of revolutions per minute (rpm). Noting that
rad 1 rad 60 sec 1 rev
1 = = 9.549297 rpm (12.22)
sec sec min 2π rad
we obtain horsepower in terms of rpm as
ft · lb
1 hp = 550 × 9.549297 rpm sec = 5252.11 ft · lb · rpm (12.23)
sec
The conversion from horsepower to Watts can be shown to be

1 hp = 745.699872 W ≈ 745.7 W 1 W = 0.00134102 hp (12.24)

The power vs. engine speed and torque vs. engine speed curves characterize the proper-
ties of an engine. Generic curves are shown in Figure 12.7 for a passenger vehicle. Engine
designers place the peak power and peak torque values depending on the intended use of
the vehicle.
For a particle model, Newton’s Second Law and the expression of power become ma = F
and P = F v. Combining these two expressions we obtain
F Fg P g
a = = = (12.25)
m W v W
The torque generated by the engine is transferred to the wheels through a transmission
system and shafts. We can write Twheel = GTengine , where G is the gear ratio. During the
process, there usually is some loss of power. Denoting the efficiency of the transfer by η, so
Pwheel = ηPengine , the torque provided to the wheels is

Twheel = Tengine Gη (12.26)


632 Applied Dynamics

 




       

FIGURE 12.7
Power vs. rpm and torque vs. rpm.

A lower gear has a higher gear ratio, which provides more torque to the wheels at lower
speeds. For example, for a 2011 Honda Accord sedan, the gear ratios are 1st: 3.267, 2nd:
1.778, 3rd: 1.154, 4th: 0.870, 5th: 0.647.

12.5.1 Constant Power Approximation


The power vs. rpm and torque vs. rpm plots of an engine powered by fossil fuels, such as
gasoline or diesel, are usually not in the form of a flat straight line, so that engine power
changes with engine speed. Engine designers place the highest power at a desired engine
speed, depending on the intended use of the vehicle. Electrical motors have flatter engine
torque curves than gasoline engines.
It is sometimes convenient to assume that the engine power is constant for a short
duration of time. In such cases, we get interesting results about velocities, accelerations,
and distances traversed. Let us model the vehicle as a point mass. The constant power
approximation basically says that P = F v = constant, where P is the power, F is the
tractive force, and v is the vehicle speed. Noting from Newton’s Second Law that F =
ma = m dv/dt and introducing this expression into the power expression gives
dv
P = mv (12.27)
dt
The above equation can be integrated in a number of ways. Consider first integration
over time. Multiplying both sides of the above equation by dt and integrating give
Z tf Z vf
1
mv dv =⇒ P (tf − ti ) = m vf2 − vi2

P dt = (12.28)
ti vi 2
with subscripts i and f denoting initial and final, respectively. Multiplying both sides of
Equation (12.27) by dx, where x is the position variable, and noting that v = dx/dt and
integrating yield
Z xf Z xf Z vf
dv 1
mv 2 dv =⇒ P (xf − xi ) = m vf3 − vi3 (12.29)

P dx = mv dx =
xi xi dt vi 3
The above two relationships can be used to analyze displacement and velocity performance.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 633

Example 12.3
In a top fuel drag race, dragsters can reach speeds of 300 mph1 in 1/4 mile.2 Calculate the
power used by the drag car, the average value of the acceleration during the race, and the
elapsed time. The vehicle weighs 2400 lb.
Equation (12.29) gives for xf = 5280/4 = 1320 ft, vi = 0, vf = 300 × 88/60 = 440
ft/sec, and mass m = 2400/32.17 slugs. Using Equation (12.29), we can solve for the power
needed as
1 mvf3 1 2400 4403
Php = = = 2917.8 hp [a]
550 3xf 550 32.17 3 × 1320
where Php is power expressed in terms of horsepower. The elapsed time is found by using
Equation (12.28), which gives (multiply power in hp by 550 to convert to lb·ft/sec)

mvf2 2400 4402


tf = = = 4.50005 sec [b]
2P 32.17 2 × 2917.8 × 550
The average acceleration can be obtained from aave = vf /tf . It is customary to express
the acceleration in terms of the gravitational constant, so
440
aave = = 3.04g [c]
4.5 × 32.17
which is a large number. In the beginning second of a drag race, the acceleration is closer
to 6g, which results in a tremendous force exerted on the driver, as well as on all of the
components of the drag car. When the drag race is over, the driver deploys a parachute to
slow down, and this time experiences deceleration levels of around 6g.
Next, let us carry out the same analysis for a passenger car. Considering a 4000 lb car
that reaches 0 to 60 mph in 10 seconds, the average power use can be obtained by
1
∆ Energy mv 2 4000 882
P = = 2 = = 48, 144 ft · lb/sec = 87.53 hp [d]
∆ time ∆ time 2 × 32.17 10
which is a much smaller number.
For a particle, the maximum acceleration is the product of the coefficient of friction times
g (ma = µs mg). The coefficient of friction between a tire and road surface is µs ≈ 0.85, and
for a specialty tire slightly over µs = 1. For a vehicle, the maximum acceleration is given
by Equation (12.12).
Race car drivers conduct a tire spinning action before the race begins called burnout,
which heats the tires and making them stickier, thus increasing the value of the coefficient
of friction.3 Even then, the tire friction coefficient µs does not come anywhere close to six.
How can a dragster attain accelerations of 2.5g or more? The answer lies in the way
the rear tires are designed. The rear tire of a dragster is big and wide. It is 36 inches in
diameter and the tire pressure is about 7 psi, which results in a large contact area between
the tire and the road surface, almost 10 inches longer than in a passenger vehicle.
The tremendous torque created by the engine and the centrifugal forces from the rapid
spin of the tires have a ballooning effect on the low-pressure tires, instantaneously increasing
their diameter to about 43 inches. This results in a further increase of the contact patch,
permitting generation of higher friction forces. Deformation of the rear tires of a dragster
during acceleration is shown in Figure 12.8.
1 Thisspeed is higher than the takeoff speed of some aircraft.
2 Inrecent years, the length of the track has been reduced from 1/4 mile to 1000 ft.
3 Burnout also cleans the tires and rids them of any debris that may be stuck on them.
634 Applied Dynamics

FIGURE 12.8
Deformation of rear tire of a dragster. Printed with permission of Mickey Thompson Tires,
mickeythompsontires.com.

12.6 More Advanced Model Including Wheel Inertia


The acceleration model above can be made more realistic by including wheel inertia terms.
We assume the tires to be rigid so that the contact between a tire and the road is modeled
as point contact.

x
g
G b
z Motion
direction
c
h
B
Road
A 

FIGURE 12.9
Half-car model with wheel inertias.

Consider a front wheel drive model. Figure 12.9 shows the half-car model. The mass and
mass moment of inertia of the body are denoted by M and IG , the combined mass of the
wheels and axle by m, and mass moment of inertia of the wheels on each axle by Iw .4 The
radii of the wheels are R.
The free-body diagrams of the vehicle body and the front and rear axles (two wheels
per axle) are shown in Figure 12.10. The propulsive power is transmitted as a torque from
the engine to a drive shaft and from the driveshaft to the axle and from the axle to the
4 In Chapter 15, when studying pitch and bounce motions and suspensions, the wheels and axle are

referred to as the unsprung mass and the vehicle body is called the sprung mass.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 635

v, a
x
b
c
T
h–R B
M, IG
Bx
A
Mg
Ax Bz
 z
Z
Az
T
Rear m, Iw Front
Az Bz v x
x
A B
Ax Bx
FT
Fr Wr Wf
mg  mg 
z z

FIGURE 12.10
Free-body diagrams of vehicle and wheels.

propulsive front wheel. Assume that the wheels roll without slip and the body of the vehicle
does not pitch up or down. Hence, the body of the vehicle and the centers of the wheels all
have the same acceleration a = RΩ̇, which is in the x direction.
Consider the front wheel first. Denoting the angular velocity of the tires by Ω and
summing moments clockwise about the center of the wheel lead to
X
 M = Iw Ω̇ = T − FT R (12.30)

and the force balances yield


X
+% Fx = ma = FT − Bx − mg sin θ

X
+& Fz = 0 = Bz + mg cos θ − Wf (12.31)

The tractive force FT is a friction force that is generated between the tire and road,
from the resistance of the tire to the torque acting on the tire. Its maximum value is
FTmax = µs Wf .5 From Equation (12.30), the tractive force in terms of the axle torque T
is obtained as FT = (T − Iw Ω̇)/R. Substituting this into the force balance in the motion
direction (x) in Equation (12.31), and noting that a = RΩ̇, the moment balance becomes

Iw + mR2 Ω̇ = T − Bx R − mgR sin θ



(12.32)
5 When friction levels are low, as when driving over a slick road surface or on ice, the coefficient of friction

becomes small and the wheel may spin in place (roll with slip).
636 Applied Dynamics

Let us now consider the rear wheel. The free-body diagram indicates that the friction
force Fr acts opposite the direction of motion. This is because in the absence of friction,
the point of contact between the tire and ground will move in the positive x direction. The
force balances for the rear axle become
X
+% Fx = ma = Ax − Fr − mg sin θ

X
+& Fz = 0 = Az + mg cos θ − Wr (12.33)

and the moment balance becomes


X
 M = Iw Ω̇ = Fr R (12.34)

Solving from the above for the friction force Fr = I Ω̇/R and substituting this value into
the sum of forces in the motion direction in Equation (12.33) gives

Iw + mR2 Ω̇ = Ax R − mgR sin θ



(12.35)

Adding Equation (12.32) and Equation (12.35) results in the describing equation for the
axles

2 Iw + mR2 Ω̇ = T − (Bx − Ax ) R − 2mgR sin θ



(12.36)

In the free-body diagrams in Figure 12.10, the reaction forces are drawn in their correct
directions for acceleration. The drivetrain applies the engine torque to the front wheels. The
front wheels exert a force on the vehicle through the axle. The front axle “pulls” the vehicle
body forward. In turn, the vehicle body exerts a force on the rear axle and “pulls” the rear
axle forward, which makes the rear wheels turn. Looking at the free-body diagram of the
vehicle, the axle force Bx is in the x direction and the axle force Ax is in the −x direction.
The force balances for the vehicle body are
X
+% Fx = M a = Bx − Ax − M g sin θ

X
+& Fz = 0 = −Az − Bz + M g cos θ (12.37)

and, noting that the angular velocity of the vehicle body is zero, the moment balance
equation about the center of mass is

 MG = 0 = Az b − Bz c + (Ax − Bx ) (h − R) − T (12.38)

The sum of forces in the vertical direction in Equation (12.31) and Equation (12.33)
gives the vertical forces between the vehicle and the tires as

Bz = Wf − mg cos θ Az = Wr − mg cos θ (12.39)

Substitution of these expressions into Equation (12.37) gives the expected result

Wr + Wf = (M + 2m) g cos θ (12.40)

Let us now rewrite Equation (12.36) in terms of the axle reaction forces in the horizontal
direction, (Bx − Ax ) R, as

(Bx − Ax ) R = T − 2 Iw + mR2 Ω̇ − 2mgR sin θ



(12.41)
Vehicle Dynamics—Basic Loads and Longitudinal Motions 637

and substitute it into the force balance in the motion direction equation for the vehicle body
in Equation (12.37), with the result
M R2 + 2 Iw + mR2 Ω̇ = T − (M + 2m) gR sin θ

(12.42)
which becomes the equation of motion for the vehicle. Denoting by M 0 = M + 2m, I 0 =
M R2 +2 Iw + mR2 and recalling that the acceleration is related to the angular acceleration
by a = RΩ̇, the equation of motion can be rewritten as
R M 0 R2
a = T − g sin θ (12.43)
I0 I0
We can use the above equations to calculate the wheel loads Wf , Wr and the axle loads
Ax , Az , Bx , Bz . When the mass of the axles and wheels and the inertia of the wheels are
ignored, setting m ≈ 0, Iw ≈ 0, we end up with M 0 = M, I 0 = M R2 , and the expression
for acceleration reduces to
T
a = − g sin θ (12.44)
MR

12.7 Braking
Mathematically, braking is the opposite of acceleration, with the tractive forces acting
opposite the direction of the forward velocity, so all the results obtained in Section 12.4
apply for braking when the sign of the tractive force is reversed. However, because the forces
applied during braking are much larger, and brake forces are applied to all four wheels, with
differing brake pressures at the rear and front wheels, analysis of braking warrants separate
treatment. Braking, like acceleration, is a friction mechanism. The braking force is limited
by the friction between the tire and the road.

12.7.1 Brake Effects


As an illustration, consider the 4000 lb car in Example 12.3, accelerating from 0 to 60 mph
in 10 seconds. Assuming constant acceleration and ignoring rolling resistance, the tractive
force becomes
∆v 4000 88
FA = ma ≈ m = = 1094 lb (12.45)
∆t 32.17 10
For the same vehicle to come to a stop in three seconds, the braking force FB must be
10/3 = 3.33 times higher, so that FB = 3640 lb. The braking force needed for a 50,000 lb
truck to stop in three seconds is over 45,000 lb. These tremendous brake forces are generated
by two small brake pads on each wheel. The contact between the brake pads and the wheel
is in the form of sliding friction, which generates significant heat.
Performance demands of braking require that brake mechanisms be extremely accurate
regarding alignment, wear, and sensitivity. Brakes must be very reliable and resistant to
failure, deformation, and high heat. They must be able to dissipate the heat generated on
the brake pads in a short period of time.
Rolling resistance, aerodynamic and driveline drag,6 and going uphill all help with brak-
ing. As will be discussed in the next chapter, rolling resistance provides deceleration of about
6 Driveline drag is the loss of power that occurs when the engine power is transmitted to the wheels

through the transmission, drive shafts, differential, and axle shafts.


638 Applied Dynamics

0.01g to 0.02g, and aerodynamic drag at normal highway speeds provides about 0.03g. A
grade of x% (tangent of the grade ≈ sine of the grade ≈ value of the grade in radians)
provides 0.01xg in deceleration. Driveline drag is more prevalent at low speeds. However, at
low speeds aerodynamic drag is not strong. As a ballpark estimate, braking of about 0.1g
is available to the driver at any time. Conversely, a vehicle has to continuously overcome a
resistive force that provides a deceleration of 0.1g.

12.7.2 Force Analysis

G
c b

h
W
FBr FBf
L
Wr Wf

FIGURE 12.11
Braking on level road.

Figure 12.11 depicts a vehicle on level ground to which braking forces are applied. The
braking forces on the front and rear axles are denoted by FBf and FBr , respectively. The
total braking force is FB = FBf + FBr . The sum of forces in the tractive direction (x) gives

maGx = −FB = − FBf + FBr (12.46)

Introducing the dimensionless deceleration factor Dx = −aGx /g, the deceleration can
be expressed as Dx g and it can be related to the brake force as FB = W Dx .
It follows from the analysis for acceleration in the previous section that the wheel loads
for braking become
c h h
Wf = W + FB = Wfs + W Dx
L L L

b h h
Wr = W − FB = Wrs − W Dx (12.47)
L L L
The braking force is a friction force between the tires and the road surface. It is generated
through a torque that acts on the tire in a direction opposite the impending angular accel-
eration of the tire. This braking torque is generated with yet another friction mechanism,
the brake pads.
Figure 12.12 shows the free-body diagram of a wheel to which a braking torque TB is
applied.7 Neglecting the inertia of the tire (Iw Ω̇ << TB so the sum of moments about
7 Note that the friction force opposes impending slip due to the torque.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 639

W
R

v
TB
Rl

FB
N

FIGURE 12.12
Free-body diagram of braking wheel. Undeformed portion has radius R and deformed por-
tion has radius Rl .

the center of the tire gives Rl FB = TB ), we can calculate the braking force by summing
moments about the center of the tire, which yields the relationship
TB
FB = (12.48)
Rl
where Rl is the distance between the center of the tire and the ground and is referred to as
the loaded radius.
The brake force is strongly affected by the loaded radius. An overinflated tire has a
larger radius, implying that it will generate a smaller brake force. An underinflated tire can
provide higher brake forces, but it also leads to higher rolling resistance forces, as well as
reduced stability of the sidewall.

Piston

Caliper

Brake pads
Wheel
attaches here
Rotor
Hub

FIGURE 12.13
Disc brake mechanics.

There are two commonly used brake mechanisms: disc brakes or drum brakes. In a disc
brake, two pads press against a thin rotor that rotates with the tire, as shown in Figure
12.13. The brake pads are pressed against the rotor by calipers, which are actuated by a
piston. The piston is actuated by a hydraulic mechanism in smaller vehicles and by air in
larger vehicles, such as trucks.
640 Applied Dynamics

Brake cylinder
Pistons
Drum
Brake pads
Cable to
emergency
brake lever
Emergency
brake mechanism

Adjuster mechanism Brake shoes

FIGURE 12.14
Drum brake mechanics.

A drum brake, shown in Figure 12.14, consists of a thin cylinder called a drum that
rotates with the wheel. Two brake shoes, in the shape of the inside circumference of the
cylinder, press against the cylinder, creating the friction forces. The motion of the pads is
initiated through a hydraulic or air pressure system. An advantage of a drum brake is that
it can accommodate emergency (parking) brakes more efficiently than can disc brakes.
Disc brakes are more expensive but they are more precise than drum brakes. The brake
torque that they generate does not fade as rapidly as in a drum brake. Friction forces gener-
ated by the brake shoes decrease when the shoes get hot, lowering the braking effectiveness,
a phenomenon known as fading. In drum brakes, the heat generated by friction between
the brake shoes and drum stays inside the drum, having less opportunity to dissipate than
in a disc brake. Figure 12.15 gives a generic plot of fading of brake torque.










FIGURE 12.15
Fading of brake torque.

As described earlier, weight shifts to the front during braking, so the front brakes work
harder. The majority of passenger vehicles have disc brakes in front and drum brakes in
Vehicle Dynamics—Basic Loads and Longitudinal Motions 641

the rear. Trucks and buses used to have only drum brakes. However, since 1980 disc brakes
have been introduced in trucks and buses. Disc brakes also perform better in wet weather,
because the centrifugal force associated with the rotation of the wheel flings water off the
brake rotor, whereas in drum brakes the water stays longer inside the drum.
It is important to note that even small effects, such as differences in tire pressure (impacts
tire radius and friction amount), tire size (impacts tire radius), and tire wear (impacts tire
radius, as well as friction coefficient), as well as disproportionate loading of the vehicle
(impacts the wheel loads), lead to differences in the brake torques. Major problems may
emerge when braking, such as wheel lock-up and instability.
It is important to ensure that the vehicle is loaded evenly, the tires are properly inflated,
and the wear on tires is even. A set of new tires in the front and a set of worn out tires in
the rear, or regular tires in the front and snow tires in the rear, can cause changes in the
braking force.8
The problem of brake lock-up can occur when the driver slams on the brakes. If the force
exerted by the brake pads on the rotor (or by the brake shoes on the drum) becomes very
large, the wheel stops rotating before the vehicle comes to a halt. This results in slipping
of the wheel, so that the friction force between the wheel and the ground is dictated by the
kinetic coefficient of friction, which is smaller than the static coefficient of friction.
To make matters worse, slipping tires lose their capability to provide directional stability.
If the rear wheels lock up, even a small disturbance to the rear wheels initiates a large
amount of yawing. If the front wheels lock up, steering ability is lost but the rear wheels
continue to provide some stability. It is widely accepted that rear wheel lock-up is more
dangerous than front wheel lock-up. Brake designers bias their design so that front brakes
will lock before rear wheels.

12.7.3 Brake Proportioning


One of the key issues in the design of brakes is determining the torques that are applied to
the wheels so that the desired brake forces are obtained, no lock-up occurs, and the brakes
are used efficiently. The brake forces should not exceed the limit values of µs Wr for the rear
and µs Wf for the front tires. Defining the effective friction coefficients µr and µf , so that

FBr FB f
µr = µf = (12.49)
Wr Wf

it is preferable to have these coefficients close to each other.


We can measure the effectiveness of the braking system by comparing the values of the
effective friction coefficients µr and µf with the friction coefficient between the tires and
road, µs , as well as the deceleration of the vehicle, measured in terms of the deceleration
factor Dx = −aGx /g. Braking efficiency is defined as

Dx
ηb = (12.50)
µb
where ηb is the braking efficiency and µb is the larger of µr and µf .
The braking torques are generated by a hydraulic mechanism which, when the brake
pedal is depressed, comes under pressure and pushes the brake pads against the disc (or the
drum). Front wheel loads become higher during braking, so we can apply larger brake forces
to the front wheels than to the rear wheels. We therefore need the brake torque applied to
8 Some dealers who sell snow tires will not sell fewer than four tires at a time, or will ask the customer

to sign a disclaimer when only two tires are purchased.


642 Applied Dynamics

the front and rear wheels to be different. The different brake forces are generated through
brake proportioning, where a valve sends hydraulic oil in different pressures to the front
and rear brake lines.
Because brake performance is of utmost importance, strict rules have been developed
with which automakers must comply. In the United States, the Highway Safety Act of
1965 established the National Highway Traffic Safety Administration (NHTSA), which has
developed several vehicle safety standards called Federal Motor Vehicle Safety Standard
(FMVSS).9 Of these, FMVSS 105 and 13510 deal with passenger vehicle and light truck
braking safety, and 121 and 122 deal with air brakes (trucks) and motorcycles, respectively.
These braking standards specify stopping distances and brake pedal forces for a variety of
driving conditions, such as dry and wet roads, loaded vehicles, partial brake failure, brake
temperature, fade, recovery, and new vs. worn-out brakes.
We can relate stopping distances to acceleration levels using a constant deceleration
approximation of v 2 = 2ds, where v is the initial speed, d = Dx g is the deceleration, and
s is the stopping distance. For example, FMVSS requires that a vehicle on a dry road stop
from an initial speed of 100 km/h (62.1 mph) in 70 m (230 feet). As an illustration, taking
a speed of 60 mph = 88 ft/sec, the average deceleration that the vehicle must sustain is
d = v 2 /2s = 882 /460 = 16.83 ft/sec2 , which is about 0.52g.
Consider Equation (12.47) and denote the brake forces on each axle by Ff and Fr . The
maximum brake force on each axle then becomes
   
h h
Ffmax = µWf = µ Wfs + W Dx Frmax = µWr = µ Wrs − W Dx (12.51)
L L

We cannot solve for the maximum braking forces from the above two equations directly,
because the total deceleration that the vehicle attains is due to the total braking forces
generated in the front and rear. The total acceleration in terms of the maximum front and
rear forces become

For Ffmax : mgDx = Ffmax + Fr For Frmax : mgDx = Frmax + Ff (12.52)

Introduction of the relationships above into Equations (12.51) yields


   
µ h µ h
Ffmax = h
W fs + Fr F rmax = h
W rs − F f (12.53)
1 − µL L 1 + µL L

which gives two equations for the maximum brake forces, where the maximum force for the
front axle is in terms of the rear brake force and vice versa. Both equations are linear.
In addition to the requirements in Equation (12.53), vehicles must comply with minimum
deceleration requirements, which can be expressed as

Ff + Fr ≥ W Dx (12.54)

It is interesting to plot Equations (12.53)–(12.54), which is shown in Figure 12.16. The


two maximum force curves define the safety envelope, inside which there is no wheel lock-
up. The triangle formed by the intersection of the two lines in Equation (12.53) and the
minimum desired deceleration line in Equation (12.54) defines the safe zone. Inside this
triangle, which is shown shaded in Figure 12.16, the brake forces do not lead to lock-up,
while the total deceleration is above the required minimum.
Note that in Figure 12.16, the horizontal coordinate is Ff and vertical coordinate is Fr ,
9 See the web site: http://www.nhtsa.dot.gov/cars/rules/import/fmvss/index.html
10 FMVSS 135 has replaced FMVSS 105 for model year 2000 and after.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 643

corresponding to the x and y coordinates of an xy plot. This implies that for front brake
lock-up, we express the first relation in Equation (12.53) in the form y = mf x + bf . For
the rear brake, the second relation in Equation (12.53) can be written can be written as
x = m0r y + b0r . We can rewrite this equation as

y = mr x + br (12.55)

where mr = 1/m0r , br = −b0r /m0r .


The proportioning line should go through the safe region. Because it is less dangerous to
have the front wheels lock before the rear wheels, it is preferable to draw the proportioning
line so that it crosses the front lock-up line before it crosses the rear lock-up line, as shown
in Figure 12.16.11

W fs
WDx
1 – h/L Front and
rear lockup
Front lockup
h/L
No Slope =
1 – h/L
lockup
area Safe region
Front Brake Force (Ff)

–h/L
Slope =
1 + h/L

Minimum
deceleration
Rear lockup

Proportioning line
line

W rs
1 + h/L

45°
Rear Brake Force (Fr) WDx

FIGURE 12.16
Brake proportioning diagram for a given friction level µ.

The challenge with brake proportioning is that the friction coefficients between the tire
and road change depending on the road conditions. In addition, the vehicle may be loaded in
different ways, which changes the weight distribution. The proportioning valve design must
take into consideration all these consequences. A generic brake proportioning is shown for
two braking requirements in Figure 12.17. Here, the proportioning curve is not a straight
line, which ensures that the line goes through both shaded regions.
In recent years, the concept of electronic brake distribution (EBD) has emerged as a
way to optimize brake force distribution based on conditions such as changes in wheel
loads, friction levels, and sliding of tires.
11 Note: We omit the subscript s from the coefficient of friction µ to simplify the expressions in the figure.
644 Applied Dynamics

1st Requirement

Front Brake Force (Ff )


Proportioning curve

2nd Requirement

45° 45°
Rear Brake Force (Fr)

FIGURE 12.17
Brake proportioning for two braking requirements.

12.7.4 Anti-Lock Brake Systems


The anti-lock brake system (ABS) is a technology where a controlling computer fluctuates
the pressure applied to the brakes when the wheels are about to lock or if the vehicle is
about to lose traction. In essence, the ABS pumps the brakes on behalf of the driver.
An ABS consists of four components: speed sensors, valves, a pump, and a controller:
• The speed sensor (one on each wheel controlled by ABS or at the differential) senses if
a certain wheel is turning much slower than the others, a sign that lock-up impending.
• The valves (one on each wheel) control the flow of brake fluid from the master cylinder
to the piston. They reduce or increase the pressure in the brake line.
• The pump regulates the pressure in the ABS system.
• The controller is a computer that receives information from the sensors. It determines
if the ABS system should kick in and sends the necessary commands to the valves and
pump.
The success of ABS systems has led to more sophisticated safety systems, such as elec-
tronic stability control (ECS)12 and traction control systems (TCS). Both of these systems
make use of additional sensors, such as steering wheel angle sensors, accelerometers, and
yaw velocity sensors, to check if the vehicle is sliding out of control (or is about to slide) or
if a tire is beginning to lose traction. These systems restore stability by properly adjusting
the brake torque and/or throttle.

Example 12.4
This example involves an accident on a level highway, where investigators noticed skid
marks left by the rear tires, while no skid marks were left by the front tires. Why would
12 Electronic stability control systems are discussed in Chapter 14.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 645

the rear tires skid and front tires not skid? Obviously, there was locking in the rear brakes.
The driver testified that he was driving at normal highway speed and gradually applied the
brakes upon seeing an animal. The vehicle went out of control. What went wrong?
Skidding (or locking brakes) occurs when the torque applied by the brake pads is higher
than the brake forces that can be generated by the tires. It turns out in this case that the
driver had different size tires in the front and back, with the rear tires having a radius of
12.7 in. while the front tire radius was 13.8 in. Under such circumstances, the rear tires
would lock first, and they did.
Assuming that the wheel loads are Wf = 0.65W and Wr = 0.35W , and all tires have
the same friction coefficient µs , the maximum braking forces are
FBf = µs Wf = 0.65µs W FBr = µs Wr = 0.35µs W [a]
The maximum brake torque that can be applied to each wheel then becomes
Tfmax = Rf FBf = 13.8 × 0.65µs W = 8.970µs W

Trmax = Rr FBr = 12.7 × 0.35µs W = 4.445µs W [b]


so that the maximum torques that can be applied differ by
Tfmax 8.970
= = 2.017 [c]
Trmax 4.445
It is likely that the brake proportioning valves did not have that high a ratio. Hence,
the rear brakes locked before the front wheels. As discussed earlier, locking of rear wheels is
more dangerous than locking of front wheels; it takes longer for the vehicle to recover from
the sliding when the rear tires begin to slide.

Example 12.5
The vehicle of weight 3500 lb in Figure 12.18 is decelerating with the brake forces shown.
Calculate the braking efficiency of the rear and front axles.

3500 lb
G a x
28"
z
A B
540 lb 900 lb
6' 4'

FIGURE 12.18
Vehicle with brakes applied.

The total braking force is


FB = 540 + 900 = 1440 lb [a]
so that the nondimensional deceleration factor Dx is
FB 1440
Dx = = = 0.4114 [b]
W 3500
646 Applied Dynamics

From Equation (12.47), the wheel loads are

c h 6 28/12
Wf = W + FB = 3500 + 1440 = 2436 lb
L L 10 10

b h 4 28/12
Wr = W − FB = 3500 − 1440 = 1064 lb [c]
L L 10 10
The effective braking coefficients on the front and rear axles are
FBf 900 FBr 540
µf = = = 0.3695 µr = = = 0.5075 [d]
Wf 2436 Wr 1064

It is preferable that the effective friction coefficients for the front and rear brakes be
similar to each other. It is this goal for which the proportioning valves should be designed.
The braking efficiency based on the rear axle is
Dx 0.4114
η = = = 0.8107 = 81.07% [e]
µr 0.5075

Example 12.6—Proportioning Valve Design


Develop a proportioning valve design for the brake system of a vehicle to provide deceleration
levels of 0.7g while the brakes are fully operational, where the friction coefficient is µs = 0.81.
The vehicle wheelbase is 120 in. and the center of mass height is h = 30 in. The static
wheel loads are Wfs = 2000 lb, Wrs = 1800 lb. We will draw the brake proportioning
diagram covering this condition and select a proportioning line that will satisfy the braking
requirements for the vehicle.
We begin by calculating the intercepts of the three lines. The deceleration curve is

Fr = W Dx − Ff [a]

Let us write this equation as y = mb x + bb , where the slope mb = −1 and the y intercept is
bb = W Dx = 3800 × 0.7 = 2660 lb. The subscript b denotes braking. This line crosses the
x and y axes at Ff = Fr = 2660 lb.
Noting that the center of mass height to wheelbase ratio is h/L = 30/120 = 0.25, the
front brake lock-up equation can be written as y = mf x + bf , where

µs h/L 0.81 × 0.25


mf = = = 0.2539
1 − µs h/L 1 − 0.81 × 0.25

µs Wfs 0.81 × 2000


bf = = = 2031.3 lb [b]
1 − µs h/L 1 − 0.81 × 0.25
The rear brake lock-up equation can be written as x = m0r y + b0r , where

µs h/L 0.81 × 0.25


m0r = − = − = −0.1684
1 + µs h/L 1 + 0.81 × 0.25

µs Wrs 0.81 × 1800


b0r = = = 1212.5 lb [c]
1 + µs h/L 1 + 0.81 × 0.25
We can write the rear brake lock-up equation as y = mr x + br where, as shown earlier,
1 b0r
mr = = −5.9383 br = − = 7200 lb [d]
m0r m0r
Vehicle Dynamics—Basic Loads and Longitudinal Motions 647

Next, we calculate the points at which the three lines intersect. For example, for the
intersection of the front brake lock-up and rear brake lock-up equations we equate
y = mf x + bf = mr x + br [b]
with the result
br − bf
x = = 834.70 lb y = mf x + bf = 2243.3 lb [e]
mf − mr

y (Ff ) lb Rear lockup curve

2660

2250 Front lockup curve


2031

1500

750 Dx = 0.7

Proportioning Dx = 0.6
line
x (Fr) lb
0
750 1500 2031 2250 2660

FIGURE 12.19
Brake proportioning design.

We find the other two intersections in a similar way, with the results
Front brake lock-up and deceleration curve: (501.35, 2158.7) lb
Rear brake lock-up and deceleration curve: (919.35, 1740.7) lb [f ]
The next step is to plot these equations, which is done in Figure 12.19. The proportioning
line, as discussed earlier, is drawn to the left of the intersection of the two lock-up curves.
The safe region is small, which points to the difficulties associated with finding realistic
values for desired decelerations and selecting the center of mass to wheelbase ratio.
If we reduce the desired deceleration level to Dx = 0.6, the deceleration curve crosses
the Ff and Fr axes at 2280 lb. This line is also given in Figure 12.19. As can be seen, the
safe zone has increased substantially.

12.8 Rollover and Lateral Instability


This section presents a simplified model to calculate the weight shift on a vehicle due to
the application of lateral forces, and a way to calculate the rollover propensity of a vehicle.
648 Applied Dynamics

The lateral forces may result from cornering forces, a side force applied to the vehicle in the
form of a wind load, or the component of gravity on a banked road. We are not considering
side forces generated as a result of impact with another vehicle.

Path

x v

Outside (o)

Inside (i)
y G L

FIGURE 12.20
Top view of vehicle taking a right turn.

Figure 12.20 shows the top view of a vehicle taking a right turn with speed v. The
wheels closer the turn center are referred to as the inside wheels and the wheels further
away from the turn center are called the outside wheels. The subscripts i and o denote the
inside and outside. The distance between the wheels on the same axle is referred to as track
and denoted by t. The free-body diagram of an axle is shown in Figure 12.21. The center
of mass is at a height h from the ground. The weight carried by the axle is denoted by W
and the lateral forces are depicted for the case when the vehicle is taking a right turn.

y
G 
z
h

Fo Fi
t
Wo Wi

FIGURE 12.21
Rear view of axle, while taking a right turn.

Summing forces and moments gives


+
X mv 2
→ Fy = may =⇒ Fi + Fo = (12.56)
R
X
+↓ Fz = 0 =⇒ Wi + Wo = W (12.57)
Vehicle Dynamics—Basic Loads and Longitudinal Motions 649
X t t
 MG = 0 =⇒ Wo − Wi − (Fi + Fo ) h = 0 (12.58)
2 2
Solving the above equations for the wheel loads gives

W mv 2 h W mv 2 h
Wi = − Wo = + (12.59)
2 R t 2 R t
indicating that weight shifts to the outside wheels during a turn. This weight shift affects
the lateral forces as well as the brake forces that can be applied. A higher center of mass
increases the amount of weight shift and a longer track reduces the amount of weight shift.
Race cars are designed so that they are wide and their center of mass is close to the ground.
The analysis above is valid for any type of lateral load acting on the vehicle, including
2
a vehicle hit by an object in the y direction. Indeed, replacing vR with ay in the above
equation gives the general relationship for lateral load transfer
W h W h
Wi = − may Wo = + may (12.60)
2 t 2 t
Two types of scenarios need to be considered when a lateral force is applied. One is
rollover, which is studied in this section. The other scenario involves sliding of the tires
sideways and loss of lateral stability. In this scenario, the tires are not able to withstand the
lateral forces acting on the body and the vehicle begins to slide. This case can be studied
more accurately after modeling tire forces, and it will be discussed in the following chapters.
The maximum lateral force that can be generated by a tire is µs times the wheel load,
where µs is the static coefficient of friction between the tire and the road. Let us calculate
the limiting case, when the inside wheel loses contact with the ground. Setting Wi = 0 in
Equation (12.60) and solving for the maximum lateral acceleration results in
aymax t
= (12.61)
g 2h

This ratio is known as rollover propensity or static stability factor (SSF). For most cars
and SUVs, the SSF is in the range of 0.90 to 1.50. As an example, a sporty car will have
a track length of 60 inches and a center of mass height of 20 inches, resulting in an SSF of
1.5, while an SUV with the same track width will have a center of mass height of 28 inches,
resulting in an SSF 1.07. Table 12.113 gives ranges for SSF for various types of vehicles.

Vehicle Type Ctr. of Mass Height (in.) Track (in.) SSF


Sports Car 18–20 50–60 1.2–1.7
Compact Car 20–23 50–60 1.1–1.5
Family Car 20–24 60–65 1.2–1.6
SUV, Pickup Truck 30–35 65–70 0.9–1.1
Medium Truck 45–55 65–75 0.6–0.8
Heavy Truck 60–85 70–72 0.4–0.6

TABLE 12.1
Static stability factor ranges for vehicle types

NHTSA currently uses a five-star rating based on SSF and statistical models that predict
13 From the web site http://www.rqriley.com/suspensn.htm.
650 Applied Dynamics

rollover.14 For example, an SSF larger than 1.40 usually receives a five-star rating. Minivans
usually receive four and SUVs three stars. NHTSA classifies rollovers in two categories:
tripped and untripped. In a tripped rollover, the vehicle encounters a vertical force, such as
a bump, curb, ditch, or other road imperfection that initiates the rollover (or “trips” the
vehicle); 95% of rollovers are in this category. An untripped rollover occurs when a vehicle
takes a turn with high speed, as discussed in the formulation in this section.
It should be reiterated that the above analysis does not consider any roll of the vehicle
body and the effects of any suspension stiffness or tire flexibility. Chapter 15 will discuss
more accurate models that take into consideration the roll center. It should be noted that,
while the above derivation of rollover propensity is based on a simple model, research
on rollover accidents has shown that the SSF is a reliable predictor of rollover accident
probability.
Experimental results suggest that rollover occurs rapidly. This indicates that the static
analysis conducted above is not sufficiently accurate and that the angular acceleration of
rolling should be considered when analyzing rollover instability.

12.8.1 Simple Combined Lateral and Roll Analysis


Let us proceed with a simple analysis of lateral stability. Suppose a vehicle enters a turn
with high speed. Is the vehicle more likely to slide or is it more likely to roll over?
The maximum acceleration that can be attained before rollover is given by Equation
(12.61). Considering now the lateral forces, the maximum lateral load that can be applied
by the tires is governed by the friction coefficient

FLmax = Fi + Fo = µs W (12.62)

where µs is the coefficient of static friction. It follows that the maximum lateral force that
can be generated by the tires is
W
maymax = ay = µs W (12.63)
g max
which gives the expression for the maximum lateral acceleration as
aymax
= µs (12.64)
g
We have developed two expressions for the maximum lateral acceleration, Equation
(12.61) and Equation (12.64). For most vehicles, the static coefficient of friction is around
µs ≈ 0.85, which is smaller than the values of the static stability factor t/2h considered in
Table 12.1, except for trucks. This implies that a passenger vehicle is more likely to slide
than to roll, a fact known to most drivers. However, the slightest road imperfection, or
hitting a curb or some other impediment, increases the likelihood of tripped rollover. For a
truck, the probability of rollover is equal to or higher than the probability of sliding.

12.8.2 Critical Sliding Velocity


Most rollover accidents are tripped rollovers. They occur because the vehicle, while sliding,
hits an obstacle or slides into pavement or curb that has a higher friction coefficient. The
obstacle “trips” the vehicle. The critical sliding velocity (CSV) is the lowest slide speed
with which a sliding vehicle hitting an obstacle will roll over.
14 The NHTSA web site for crash and rollover ratings is http://www.safercar.gov.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 651

a) b)

Path of G
G
G GC
vy y
s
z s
h
h obstacle 
 
t D D y

FIGURE 12.22
Rear view of sliding and tripping: a) before tripping, b) rolling about D after tripping.

Consider a vehicle that is sliding with speed vy , as shown in Figure 12.22a. Tripped
rollover takes place in two stages. When the vehicle hits an obstacle, an impact occurs and
the sliding motion is converted to roll about the obstacle D. We can assume that there is
angular momentum conservation about point D during this stage. The angular momentum
about D before impact is
HDbefore = mvy h (12.65)
and the angular momentum about D immediately after impact is
HDafter = ID ω (12.66)
in which ID is the roll moment of inertia about D, and ω is the roll angular velocity.
In the second stage of tripped rollover, the vehicle rotates about point D, as shown
in Figure 12.22b. How much the vehicle rotates can be analyzed by using the principle of
conservation of energy. The total energy of the vehicle the instant roll begins is
1
E = T + V = ID ω 2 + mgh (12.67)
2
For the vehicle to roll over, it must rotate to the position where the center of mass is
directly above point D (denoted by GC in Figure 12.22b) and still have an angular velocity.
The critical value is found by setting ω = 0 so that the total energy is due to the potential
energy, or
E = mgs (12.68)
Equating Equations (12.67) and (12.68) we can calculate the minimum sliding velocity,
or CSV, for rollover. First, by equating Equations (12.65) and (12.66), the angular velocity
immediately after tripping is expressed in terms of the sliding speed as
mvy h
mvy h = ID ω =⇒ ω = (12.69)
ID
Introducing this value to Equation (12.67) and equating Equations (12.67) and (12.68)
results in
1 1 m2 vy2 h2
mgh + ID ω 2 = mgh + = mgs (12.70)
2 2 ID
Solving this equation for the speed vy yields the critical sliding velocity as
r
2gID (s − h)
vy = (12.71)
mh2
652 Applied Dynamics

Example 12.7—Critical Sliding Velocity and Rollover


Accident statistics show that for passenger vehicles, tripped rollovers due to sliding occur
at sliding speeds of 10 to 15 mph. Let us use this information in an example. Consider a
vehicle with weight of 3600 lb, track width of 62.6 in., and SSF of 1.45, as well as sliding
speed for a tripped rollover speed of 14 mph. What is the roll moment of inertia about the
center of mass for this vehicle?
Solving Equation (12.71) for the roll moment of inertia ID gives

mh2 vy2
ID = [a]
2g (s − h)
The value of the center of mass height h from the SSF is found as
t t 62.6
SSF = =⇒ h = = = 21.59 in. [b]
2h 2 × SSF 2 × 1.45
and s  2
t p
s = h2 + = 21.592 + 31.32 = 38.02 in. [c]
2
Substituting the value of s into Equation [h] gives the roll moment of inertia as

mh2 vy2 3600 × (21.59/12)2 × (14 × 88/60)2


ID = = = 1731 slug · ft2 [d]
2g (s − h) 2 × 32.172 × (38.02 − 21.59)/12
The parallel axis theorem gives the centroidal roll moment of inertia as
 2
3600 38.02
2
Iφ = IG = ID − ms = 1731 − = 1731 − 1123 = 608 slug · ft2 [e]
32.17 12

Let us check if this answer makes sense. Assume that the rear of the vehicle is modeled
as a rectangle of width W and height H. Also assume that the width W is 7 inches longer
than the track, and the height H is 55 inches. We can then estimate the mass moment of
inertia about the center of mass as

1 1 3600 702 + 552
2 2
= 513 slug · ft2

Iφ = IG = m W +H = [f ]
12 12 32.17 144
This value for the roll moment of inertia is relatively close to the calculated value. The
actual value is expected to be higher, as Equation [f] assumes that the center of mass is at
the geometric center of the rectangle. This is not the case in a car, where the center of mass
is higher than the geometric center due to the tires.
The roll moment of inertia of a car is much lower than the yaw and pitch moments of
inertia. Remember that the analysis above uses some simple assumptions (for example, we
assume that the tires do not deform while the vehicle is rolling) and we should be aware of
the limitations of the analysis.

Example 12.8—Destabilizing a Fleeing Vehicle


When police chase a fleeing vehicle, one method they use to destabilize the fleeing vehicle
is to give it a bump from the rear with a sideways component. While the bump is most
effective when the fleeing vehicle is taking a turn, a bump can also destabilize a vehicle
moving along a straight line, causing it to spin out and become uncontrollable. This action
is known as the PIT (Precision Immobilization Technique) maneuver.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 653

x x'

vB
y
B

G
IC after bump
vA
vA before
after bump
bump
A
Bump IC before bump
force F

FIGURE 12.23
Top view of vehicle bumped by police vehicle and resulting instant center.

Figure 12.23 shows the geometry and location of the instant center of the vehicle before
and after the bump. We use the bicycle model of a car and assume that the front wheel
does not slip and the steer angle remains unchanged during the bump.
The instant center of the bumped vehicle is smaller than before the bump. This implies
that the angular velocity of the vehicle becomes larger after it is bumped. This leads to two
likely effects. First, the rear wheels will most likely slide and the vehicle will be destabilized.
Second, the impact from the bump and from the increased angular velocity will disorient
the driver and make it harder for the driver to stabilize the vehicle.
To observe these effects, consider the dynamics of the bump applied to a vehicle moving
along a straight line (steering angle is zero). The free-body diagram is shown in Figure
12.24. We will consider a bump force only along the −y direction. The lateral loads are the
tire forces and their maximum values are related to the wheel loads by

Frmax = µs Wr Ffmax = µs Wf [a]

where we note that the wheel loads are determined by the center of mass location as
c L−c b
Wf = W Wr = W = W [b]
L L L

In order to ascertain the correct direction of the tire forces, we need to first solve the
impact problem in the absence of friction. The sum of forces integrated over time along the
y direction and the sum of moments about the center of mass give
X F̂
+↓ F ∆t = −F̂ = mvGy =⇒ vG y = − [c]
m
F̂ c
 MG ∆t = F̂ c = IG ω = Iψ ω =⇒ ω = [d]

where F̂ = F ∆t is the impulsive force, ∆t is the impulse duration, and Iψ is the yaw
654 Applied Dynamics

c b B
A
v G =0 x

L
y

Fbump Fr Ff

FIGURE 12.24
Free-body diagram of bumped vehicle.

moment of inertia about the center of mass. The components of the velocities of points A
and B in the y direction are calculated from relative motion equations as

c2
 
1
vAy = vGy − cω = −F̂ + [e]
m Iψ
 
1 bc
vBy = vGy + bω = −F̂ − [f ]
m Iψ
As expected, the velocity of A is in the negative y direction, but how about the velocity
of B? To this end, let us consider a simple model of a vehicle of wheelbase L and width
t = L/2. The yaw moment of inertia becomes
 2 !
1 1 L 5
m L2 + t2 = m L2 + mL2

Iψ = = [g]
12 12 2 48

For simplicity (and without loss of generality), let b = c = L/2 which gives
 
1 bc 7 F̂
vBy = −F̂ − = [h]
m Iψ 5m

so that the velocity of B is in the positive y direction. The friction force Ff acts in the
negative y direction. We can show that the velocity of point A becomes

c2
 
1 17 F̂
vAy = −F̂ + = − [i]
m Iψ 5 m

The rear friction force Fr acts in the positive y direction. We conclude that the directions
of the friction forces in Figure 12.24 are correct.
Next, consider the force and moment balances integrated over the impulse and in the
presence of the lateral forces, which are
X
+↓ F ∆t = −F̂ + F̂r − F̂f = mvGy [j]
 
 MG ∆t = F̂ − F̂r c − F̂f b = Iψ ω [k]

Depending on the magnitude of the bump force, there are three scenarios that are
possible when a lateral bump is applied:
Vehicle Dynamics—Basic Loads and Longitudinal Motions 655

1. The force is small enough to be countered by the lateral tire forces and the vehicle does
not slide. Here, vGy = 0 and ω = 0. The magnitudes of the lateral forces are unknown
and need to be solved for.
2. The bump force is such that only the rear wheels slide. In this case, vGy 6= 0 and
ω = vGy /b. Also, Fr = Frmax = µs Wr , while the magnitude of the lateral force on the
front has not reached its maximum value.
3. Both the front and rear wheels slide. In this case, vGy 6= 0 and ω 6= 0, and the lateral
forces generated by the tires have reached their maximum values of µk Wr and µk Wf .
Equations [j] and [k] can be solved for the three possibilities to ascertain which wheels
will slide. Let us find the smallest value of F̂ for which sliding will occur. Both wheels will
slide when the lateral forces reach their maximum values of
c b
Frmax = µs Wr = µs W Ffmax = µs Wf = µs W [l]
L L
and vG < 0, ω > 0. Introducing these values to Equation [j] and [k] leads to the two
conditions that F̂ needs to satisfy as
µs W
F̂ = F ∆t > F̂rmax − F̂fmax = (b − c) ∆t
L

cFrmax + bFfmax µs W b2 + c2
F̂ = F ∆t > = ∆t [m]
c L c
The second of the two relationships in Equation [m] is the greater value and the critical
value of the bump that would make both the front and rear wheels slide. For a vehicle
of weight 3400 lb, wheelbase of L = 9 ft, friction coefficient µs = 0.8, and c = 5 ft, the
magnitude of the bump force becomes
µs W b2 + c2 0.8 × 3400 2
4 + 52 ∆t = 2478.2 ∆t lb · sec

F̂ > ∆t = [n]
L c 9×5
This is not a large force, especially when considering that the magnitudes of impulsive
forces that can be generated through impact are usually quite large. A small bump is
usually sufficient to destabilize the vehicle. To illustrate this, consider a bump force of 3000
lb applied over 0.3 seconds. The resulting impulsive force is 3000 × 0.3 = 900 lb·sec.
Assuming that the vehicle has a centroidal yaw moment of inertia of Iψ = 2150 slug·ft2 ,
let us calculate the angular velocity after an impulsive force of 3000 lb is applied over 0.3
seconds. Substituting the values of Frmax and Ffmax into Equation [k] and solving for the
angular velocity we obtain
 
∆t ∆t µW 2
b + c2

ω = (cF − (cFrmax + bFfmax )) = cF −
IG Iψ L
 
0.3 0.8 × 3400 2
4 + 52

= 5 × 3000 − = 0.364 rad/s [o]
2150 9
which is quite high, around 21◦ per second.
Note that this example neglected the horizontal motion of the vehicle, the component of
the bump in the travel direction, tire flexibility, and energy absorption by the rear bumper
and the vehicle body during impact. The component of the bump in the horizontal direction
creates a sudden change in velocity in the travel direction, which results in some of the
friction force being used up to counter sliding in the tractive direction. Also, an additional
angular velocity of the vehicle is created. The combined effects of the lateral and horizontal
components of the bump become even more devastating.
656 Applied Dynamics

12.9 Weight Shift and Statical Indeterminacy


The previous sections considered three simplified models of vehicle motion, the first two
known as half-car models:
• The half-car model for the longitudinal motion, as in Figure 12.4, viewed motion from
the side and used the front and rear axle loads.
• The half-car model for roll analysis, as in Figure 12.21, viewed motion from the rear,
considered axles separately, and dealt with inside and outside wheels.
• The bicycle model and associated lateral loads were used in the previous example on
destabilizing a fleeing vehicle.
When considered separately, all models above are statically determinate. However, when
each wheel load is considered individually, we deal with a four-wheeled vehicle that is
indeterminate. As with a table with four legs, the wheel loads cannot be determined by
the laws of statics and/or dynamics alone. Considering only vertical forces, there are four
unknowns: the four wheel loads and three equations, which are sum of forces in the vertical
(z) direction and sum of moments about the x and y axes.
This section introduces an approximate way of calculating wheel loads. Many times, car
designers will put a scale under each tire and measure the wheel loads experimentally.

12.9.1 Statically Indeterminate Systems

EI
A C B

L/2 L/2
FA FC FB

FIGURE 12.25
Statically indeterminate beam.

Consider a beam, say under a uniform load, that is supported at three points with pin
and roller joints, as shown in Figure 12.25. From statics, there are three reactions to solve
for but only two equations (sum of forces in the vertical and moment balance about a point,
or two moment equations about two different points) are available. Summing forces in the
vertical direction gives

FA + FB + FC = wL (12.72)

where w is the load per unit length and L is the length of the beam. The resultant of the
distributed load acts through the center of the beam, point C. Hence, summing moments
about point C gives
X L
MC = (FB − FA ) = 0 (12.73)
2
Vehicle Dynamics—Basic Loads and Longitudinal Motions 657

w
2
1 EI + EI
A C B A C B
FC

FIGURE 12.26
Loading on beam split into two parts.

from which we conclude that FA = FB .


A third equation is needed. This equation is obtained from the flexibility of the beam
and it is used together with the principle of superposition. The approach treats FC as a force
that acts on the beam and separates the system into two parts, as shown in Figure 12.26.
One part is a simply supported beam acted upon by the uniform load w. From mechanics of
materials, the deflection of point C, denoted by δ1 , is δ1 = 5wL4 /384EI (positive deflection
taken as downwards). The second part is another simply supported beam acted upon by a
force FC . The deflection at C, denoted by δ2 , is δ2 = −FC L3 /48EI. The sum of the two
deflections adds to zero, thus

5wL4 FC L3
δ1 + δ2 = − = 0 (12.74)
384EI 48EI
The solution of this equation for FC gives FC = 58 wL. We can then solve for the other
3
two unknown forces with the result FA = FB = 16 wL.

12.9.2 An Approximate Method for Calculating Wheel Loads


The approximate method outlined here is based on the assumption that the vehicle has a
plane of symmetry and that the center of mass lies on the plane of symmetry. Hence, in
the absence of lateral loads and roll moments, on each axle the right and left (or inside and
outside) static wheel loads are the same.
Figure 12.27 shows the individual wheel loads, which are labeled as Wri , Wro , Wfi , Wfo ,
where r, f , i, and o denote rear, front, inside, and outside, respectively. The axle loads are
Wr = Wri + Wro for the rear axle and Wf = Wfi + Wfo for the front, with Wr + Wf = W .
Based on the plane symmetry assumption, at static equilibrium, that is, when the only
external force is the weight W , the inside and outside wheel loads are the same, so
W rs Wfs
W ri = W ro = Wfi = Wfo = (12.75)
2 2
where Wrs = W Lb and Wfs = W Lc are the static axle loads.
The procedure to calculate the wheel loads is as follows.
1. Calculate all forces and moments that contribute to the wheel loads. External forces in
the x and z directions and moments about the y axis affect the axle loads. The forces
in the lateral (y) direction, moments about the x axis, and vertical loads that do not
act on the axis of symmetry affect the inside and outside wheel loads. Moments about
the z axis affect only the lateral wheel loads.
In the presence of a vertical external force Fz that does not go through the axis of
symmetry of the vehicle, a recommended procedure is to express the effect of the force as
658 Applied Dynamics

x
v W fo
G
t h y Path of G
Wro b
c

W fi
z
Wri

FIGURE 12.27
Wheel loads on a vehicle.

a force Fz that goes through the axis (plane) of symmetry and a roll moment Mx = qFz ,
as illustrated in Figure 12.28.

 

 







 
 

 



FIGURE 12.28
Dealing with a vertical force that does not go through the axis of symmetry.

2. First calculate the axle load shift. To do this, we consider the half-car model in the xz
plane. A representative free-body diagram is given in Figure 12.29, where the external
loads include a tractive force FT , a vertical load Fz , and a pitch moment My . Such a
moment may arise from the aerodynamics.
The sum of forces in the horizontal direction gives FT = maGx . The sum of forces in
Vehicle Dynamics—Basic Loads and Longitudinal Motions 659

My

My x
s Fz W
y z
G

A B
FT
c b
Wr Wf

FIGURE 12.29
Free-body diagram of side view.

the vertical direction and the sum of moments about the center of mass15 give
X
+↓ Fz = 0 =⇒ −Wr − Wf + W + Fz = 0 (12.76)

X
MG = 0 =⇒ Wf b − Wr c + FT h + My + Fz (c − s) = 0 (12.77)

Solving these two equations for the wheel loads we obtain


 
h My s
Wf = Wfs − FT + − Fz = Wfs + ∆Wf (12.78)
L L L

h My L−s
Wr = Wrs + FT + + Fz = Wrs + ∆Wr (12.79)
L L L
where ∆Wf and ∆Wr are the load shifts on the front and rear axles, respectively. Note
that these two load shift terms are not equal in magnitude and that ∆Wf + ∆Wr = Fz ,
reflecting the contribution of the external load in the vertical direction, Fz .

3. Calculate the weight shift from left to right (inside to outside) wheels. Conduct this
analysis for each axle individually. To calculate this load shift, we need to know the
distribution of the lateral forces and the roll moment on each axle.16 We express the
lateral force Fy and roll moment Mx as

Fy = Fyr + Fyf Mx = Mxr + Mxf (12.80)

where the subscripts r and f denote the rear and front axles, respectively.
The forces on each axle are apportioned in terms of the proximity of these forces to the
front and rear axles. For example, if the lateral force acts through a point that is at a
distance p from the rear axle, we apportion it as
L−p p
Fyr = Fy Fyf = Fy (12.81)
L L
660 Applied Dynamics

B F yf
b
G y
L
Fy
c
p
A F yr

FIGURE 12.30
Lateral force distribution to the axles.

The procedure is illustrated in Figure 12.30. A similar procedure is followed for the roll
moment Mx , which gets distributed in proportion to the static axle loads (p = c). Thus,

W rs b Wfs c
Mx r = Mx = Mx Mx f = Mx = Mx (12.82)
W L W L

M xr
Wr
x
aGyr
G y

z
h

F Lr
t
Wro Wri

FIGURE 12.31
Free-body diagram of rear axle for vehicle taking a right turn.

The free-body diagram of the rear axle of a vehicle taking a right turn is shown in
Figure 12.31. Here, there is no external force in the y direction, but there is a lateral
15 Alternatively, we can sum moments about another point, such as A or B. If you do so, make sure that

the inertia force is properly marked on the free-body diagram.


16 Note that the contribution of a vertical load F that does not act on the symmetry axis is accounted
z
for in the expression for Mx .
Vehicle Dynamics—Basic Loads and Longitudinal Motions 661

acceleration in the form


v2 b
aGyr = (12.83)
RL
which is generated by the lateral tire force FLr . There also is an external moment Mxr .
As the roll motion of the vehicle is ignored, there is no angular acceleration. Summing
forces in the vertical direction and taking moments about the center of mass in the x
direction gives
X
+↓ Fz = 0 =⇒ −Wri − Wro + Wr = 0 (12.84)

X t
 MG = 0 =⇒ (Wro − Wri ) − FLr h + Mxr = 0 (12.85)
2

Solving for the wheel loads we obtain


Wr h Mxr Wr h Mx r
W ri = − FLr + W ro = + FL r − (12.86)
2 t t 2 t t

Note that the load shifts in the inside and outside tires are equal and opposite, ∆Wri =
M
−∆Wro = −Fyr ht + txr .
The combined weight shift is due to the horizontal and lateral forces. The result for the
rear tires becomes
Wr W rs ∆Wr
W ri = + ∆Wri = + + ∆Wri
2 2 2

Wr Wr s ∆Wr
W ro = + ∆Wro = + + ∆Wro (12.87)
2 2 2
and for the front tires
Wf Wfs ∆Wf
Wfi = + ∆Wfi = + + ∆Wfi
2 2 2

Wf Wfs ∆Wf
Wfo = + ∆Wfo = + + ∆Wfo (12.88)
2 2 2
There are three terms in each wheel load expression. The first term is the static wheel
load (or half the static axle road). The second term is the weight shift due to the longitudinal
forces, vertical forces, and pitch moment. The third term is the weight shift caused by the
lateral forces, vertical forces, and roll moment. A pitch moment My shifts weight from front
to rear. A roll moment Mx shifts weight from the outside to the inside wheels.
Because braking forces are higher than acceleration forces, the concern is with weight
shift during braking. Articles about roll of vehicles usually show pictures of the outside front
tire while the vehicle is taking a turn and braking at the same time, when the weight shift
is the greatest for that wheel.
662 Applied Dynamics

12.9.3 Calculation of Center of Mass Location


Many times, location of the center of mass of a vehicle is not known and needs to be
calculated experimentally. This section discusses an experimental technique to calculate the
center of mass location, based on wheel load measurements.
Selecting an arbitrary reference point D on the vehicle, the center of mass location can
be expressed in vector form as
rG = rG/D = xG/D i + yG/D j + hG/D k (12.89)
The position vectors of the wheels with respect to the reference point are denoted by rri ,
rro , rfi , and rfo , as shown in Figure 12.32.
The experimental procedure consists of two steps. In the first step, values of xG/D and
yG/D are calculated. The vehicle is placed on level ground, scales are placed under each
wheel, and the wheel loads are measured.
x
W

rro rfo
Wro W fo
G
D rG/D
rri

rfi
y z
Wri W fi

FIGURE 12.32
Wheel loads on each tire and center of mass.

At static equilibrium, the sum of forces and the sum of moments about any point should
be zero. Summing forces in the vertical direction gives the total weight as
W = Wri + Wro + Wfi + Wfo (12.90)
Summing moments about the center of mass gives
X
MG = (rri − rG ) × Wri k + (rro − rG ) × Wro k

+ (rfi − rG ) × Wfi k + (rfo − rG ) × Wfo k = 0 (12.91)


which represents two equations (there is no component in the z direction) in terms of the
two unknowns xG/D and yG/D . Solving these equations yields the location of the center of
mass on the xy plane.
It remains to calculate the height of the center of mass from the ground, h = hG/D . A
convenient way to calculate the center of mass height is to use the half-car model, as viewed
from the side, and to deal with axle loads. The rear axle is hoisted up to position A0 while
keeping the front wheels locked in place, as shown in Figure 12.33. The front axle load is
measured. The center of mass height is h = Rl + h1 , where Rl is the loaded radius of the
tire, which was introduced in Section 12.7. Summing moments about point A0 gives
Wf L cos θ = W c1 (12.92)
Vehicle Dynamics—Basic Loads and Longitudinal Motions 663

T
c

d A'
G'
x
h1
 
z c1
G h1
B  W h A
Rl O

Fx
Lcos
Wf b c Wr
L

FIGURE 12.33
Free-body diagram of hoisted vehicle. Source: Race Car Vehicle Dynamics, by Milliken and
Milliken, redrawn with permission.

Solving the above for c1 , we obtain c1 = Wf L cos θ/W . From the geometry, d = h1 tan θ,
and c1 is related to h1 by
c1 = (c + h1 tan θ) cos θ (12.93)
Combining the above two equations gives an expression for h1 as
Wf L − W c
h1 = (12.94)
W tan θ
In an actual test, we would hoist the vehicle to a specified angle and measure the front
axle load Wf . Use of Equation (12.94) provides the value for the center of mass height. The
procedure is sensitive to measurement errors. It is recommended that the hoist test be done
for several values of the angle θ and the values for h obtained for the different hoist angles
be averaged.
Hoisting a vehicle to calculate the center of mass height presents several challenges. The
front wheels need to be secured without creating any new tire forces. There should be no
weight shift in the vehicle due to the motion of parts (such as transmission) and of fluids.17
In addition, the loaded radius of the tire changes with the added axle load, which should
be included in the calculations. The hoisting wires need to be perfectly vertical.

Example 12.9
A 1650 kg vehicle (wheelbase 2.8 m, track 1.2 m, CG height 70 cm, c = 1.5 m from rear, is
traveling at a speed of 70 kph on a curved level road (turning right) with radius of curvature
of 65 ft, while decelerating at the rate of 0.2g. Calculate the wheel loads.
The free-body diagrams of the vehicle are shown in Figure 12.34. We proceed with
calculating the static axle loads. Using the value g = 9.807 m/s2 , the weight of the vehicle
is 1650g = 16, 181.6 N
b 1.3
W rs = W = 16, 181.6 × = 7512.9 N
L 2.8
17 This test is usually conducted with an empty fuel tank and with the engine oil drained.
664 Applied Dynamics

a) b)
W Wr

G G
x y
1.2 m
0.7 m
z z
A B
FB F Lr
1.5 m 1.3 m
Wr Wf Wro Wri

FIGURE 12.34
Free-body diagrams of vehicle: a) side view with axle loads, b) rear axle.

c 1.5
Wfs = W= 16, 181.6 × = 8668.7 N [a]
L 2.8
The longitudinal acceleration is ax = −0.2g. Summing forces in the travel direction, the
tractive force becomes
ax
FT = −FB = max = W = 16, 181.6 × (−0.2) = −3236.3 N [b]
g
Noting that h/L = 0.7/2.8 = 0.25, the weight shift on the rear axle is
h
∆Wr = −∆Wf = FT = −3236.3 × 0.25 = −809.1 N [c]
L
so that the front axle carries a load of Wf = Wfs + ∆Wf = 8668.7 + 809.1 = 9477.8 N and
the rear axle load becomes Wr = 7512.9 − 809.1 = 6703.8 N.
The lateral acceleration is due only to the lateral force generated by the tires:
mv 2 (75 × 25/90)2
FL = maGy = = 1650 × = 9597.6 N [d]
R 65
The contribution of the lateral force on each axle is calculated as
W rs 7512.9
FL r = FL = × 9597.6 = 4456.0 N
W 16, 181.6

Wfs 8668.7
FLf = FL = × 9597.6 = 5141.6 N [e]
W 16, 181.6
The weight shift in each axle from the inside to the outside wheels is
h 2.4 h 2.4
FLr = 1122.4 × = 673.4 N FLf = 1247.2 × = 748.3 N [f ]
t 4 t 4
The next step is to calculate the individual wheel loads. From Equations (12.87)–(12.88)
Wr h 6703.8
W ri = − FLr = − 2599.3 = 752.6 N
2 t 2

Wr h 6703.8
W ro = + FL r = + 2599.3 = 5951.2 N
2 t 2
Vehicle Dynamics—Basic Loads and Longitudinal Motions 665

Wf h 9477.8
Wfi = − FLf = − 2999.2 = 1739.7 N
2 t 2
Wf h 9477.8
Wfo = + FLf = + 2999.2 = 7738.1 N [g]
2 t 2
The inside rear tire is close to losing contact with the road because braking shifts weight
to the front axle and cornering shifts weight to the outside tires.

12.10 Bibliography
Automobile Ride, Handling, and Suspension Design, http://www.rqriley.com/suspensn.htm
Genta, G., Motor Vehicle Dynamics: Modeling and Simulation, World Scientific, 1997.
Gillespie, T.D., Fundamentals of Vehicle Dynamics, SAE Publications (R114), 1992.
Huang, M., Vehicle Crash Mechanics, CRC Press, 2002.
Jazar, R.N., Vehicle Dynamics: Theory and Application, 2nd Edition, Springer, 2013.
Milliken, W.L., and Milliken, D.F., Race Car Vehicle Dynamics, SAE Publications (R146),
1995.
Tire Tech, http://www.tirerack.com/tires/tiretech/tiretech.jsp

12.11 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 12.4—Acceleration
12.1 (E) Show that the maximum acceleration of a front wheel drive vehicle on a level road
c
is amax = gµs L+hµ s
using the formulation and parameters in Section 12.4. Compare with
the results of a rear wheel drive vehicle. Ignore rolling resistance.
12.2 (M) A 4000 lb vehicle (wheelbase 9.5 ft, CG height 2.5 ft, Wf = 2200 lb under static
conditions on a level surface) is going uphill on a 3% grade. The coefficient of friction
between the tires and the road is µ = 0.85. Calculate the maximum acceleration of this
vehicle if a) the vehicle is front wheel drive, b) the vehicle is rear wheel drive.
12.3 (M) The front wheel drive vehicle of mass 2200 kg in Figure 12.35 is dragging an object
of mass 250 kg by a rope of length 2 m. The coefficient of friction between the object and
the road is µk = 0.5. Assuming that the rope stays taut, calculate the axle loads and the
tractive force needed to move the vehicle with an acceleration of 0.1g. Compare FT , Wr ,
and Wf with their counterparts when the rope is cut and the vehicle no longer drags the
object.
12.5 (M) Calculate the wheel loads of the tractor-trailer in Figure 12.36 when the tractor
is not moving. Note: the hitch that connects the tractor to the trailer can be treated as a
pin joint.
12.6 (M) Calculate the wheel loads of the tractor-trailer in Figure 12.36 when the tractor
is accelerating with 0.2g. Neglect rolling resistance and aerodynamic forces.
12.7 (M) Calculate the wheel loads of the tractor-trailer in Figure 12.36 when the tractor is
666 Applied Dynamics
x

z
G

D 1.1 m
250 kg 2.0 m
A B
 0.5 m 1.0 m 2.0 m 1.5 m

FIGURE 12.35
Vehicle dragging an object.

x
7.5 m 7.5 m
z 0.3 m
G2
8,000 kg 15,000 kg
G1
H hitch 2.4 m

1.5 m 1.3 m
A B C

1.6 m 2.5 m

FIGURE 12.36
Tractor trailer.

decelerating with 0.2g on a downhill grade of 5%. Neglect rolling resistance and aerodynamic
forces.

Section 12.5—Power
12.8 (M) A vehicle of mass 1500 kg is designed to go from rest to 90 kph in a distance of
210 m. How long will it take the vehicle to travel those 210 m and how much power will
be consumed if a) the engine provides constant power P0 , b) the engine power decreases
according to the relationship P (t) = P0 (1 − t/10) (t in seconds). Use for P0 the value you
obtained for power in part a.
12.9 (E) A dragster weighs 2500 lb and can generate a maximum power of 2700 hp. Using
the constant power approximation, calculate the speed the dragster attains in a 1/4 mile
and how much time it takes to travel that 1/4 mile. Compare these values with the speed
at 1/2 mile and the time it takes to travel that 1/2 mile.

Section 12.6—More Advanced Model Including Wheel Inertia


12.10 (M) Consider the more advanced model in Section 12.6. Calculate the wheel loads
Wf , Wr , as well as the axle loads Ax , Az , Bx , Bz for a front wheel drive vehicle.
12.11 (M) Consider the wheel loads in Section 12.6. Calculate the wheel loads Wf , Wr , as
well as the axle loads Ax , Az , Bx , Bz for a rear wheel drive vehicle.
12.12 (M) Derive an expression for the maximum torque that can be applied to the front
Vehicle Dynamics—Basic Loads and Longitudinal Motions 667

wheel drive vehicle shown in Figure 12.10 in terms of the static friction coefficient µs , so
that there will be roll without slip of the wheels.

Section 12.7—Braking
12.13 (E) The front wheel drive automobile of mass 1600 kg is traveling on a level road.
The wheelbase is 2.80 m and the center of mass is 1.30 m from the front wheel and 70
cm above the ground. The coefficient of friction between the front wheels and road is 0.75,
while the coefficient of friction between the rear wheels and the ground is 0.65. The vehicle
is being driven at 80 kph when the driver hits the brakes and all the brakes lock. The vehicle
eventually comes to a rest. Calculate the length of the tire marks on the road due to the
locked brakes.
12.14 (E) A 4000 lb vehicle (wheelbase 9 ft, CG height 2.5 ft, Wf = 2200 lb under static
conditions) is decelerating at the rate of 0.3g. It is observed that the total braking force on
the front axle is 30% more than on the rear axle. Find the minimum coefficient of friction
that would be needed to provide this braking action.
12.15 (M) A vehicle of mass 1800 kg, wheelbase 2.9 m, center of mass 1.6 m from rear,
and 65 cm high is braking. The rear brake force is 2500 N and front brake force is 4100 N.
Calculate the braking efficiency.
12.16 (M) You are at the scene of an auto accident where a driver slammed on the brakes
and left one set of skid marks on the ground that are 9 m long. The vehicle involved in the
accident does not have anti-lock brakes, it has a wheelbase of 2.5 m, the rear wheels carry
40% of the weight when the vehicle is stationary, and the center of mass height is 0.7 m.
The coefficient of friction between the road and the vehicle that left the tread-marks is in
the range of µk ≈ 0.5. The impact speed (relative speed between the colliding vehicles) at
the instant the accident occurred is 20 km/h. Find the speed of the vehicle before the driver
slammed on the brakes if a) the front brakes were not operational, b) the rear brakes were
not operational.
12.17 (M) Two distinctive features of bicycles are that their center of mass is high and
closer to the rear wheel. Consider a bicycle with center of mass height h = 0.9L and rear
static wheel load Wrs = 0.6W . Given a coefficient of friction between the road and tires of
µs = 0.8, calculate the maximum deceleration of the bicycle and the change in axle loads
if a) only the rear brakes are used, b) only the front brakes are used, and c) both sets of
brakes are used.
12.18 (D) Consider the previous problem and calculate the maximum deceleration when the
bicycle is going a) uphill in a 6% grade, b) downhill in a 6% grade.

Section 12.8—Rollover and Lateral Instability


12.19 (E) The vehicle in Figure 12.37 is parked on a banked road. Calculate the wheel loads
W1 and W2 when h = t/2. Calculate the minimum amount of friction needed, as a function
of the bank angle φ, as well as the minimum required friction forces F1 and F2 to prevent
sliding. Hint: When the vehicle begins to slide, F1 = µW1 , F2 = µW2 .
12.20 (M) For what value of the bank angle φ would the vehicle in the above problem roll
over, instead of slide? Hint: Find the friction coefficient to prevent sliding and define what
constitutes rollover of the vehicle.
12.21 (M) Consider the vehicle in the previous two problems. The vehicle is parked on a
banked road, but now the vehicle makes an angle of γ with the curb, as shown in Figure
12.38. Calculate the value of the bank angle when the vehicle begins to roll over. Hint: It
is easier to use vector notation and to calculate the moment of the gravity force about the
axis about which the vehicle would roll.
668 Applied Dynamics

x
G
z

F2
W

h
t/2
 W2
t
F1

W1

FIGURE 12.37
Vehicle parked on a banked road.

x
g
y
B
 G

t L
h
A

FIGURE 12.38
Vehicle not parallel to banked road.

12.22 (C) Consider Example 12.7 and plot the critical sliding velocity as a function of the
static stability factor. Use the range 0.5 ≤ SSF ≤ 1.7. Comment on the results.
12.23 (M) A vehicle has the following dimensions: t = 155 cm, h = 50 cm, Iφ = 900 kg·m2 .
Calculate the critical sliding velocity.
12.24 (M) A vehicle of mass m, wheelbase L, center of mass at a distance c = 2L/3 from
the rear axle, and yaw moment of inertia Iψ = mL2 /6 is to be destabilized by a bump force
along the y axis (see Figure 12.23). Calculate the largest amount of bump force F that can
be applied so that the front wheels B will not slide.
12.25 (M) A vehicle of track t = 66 in. is traveling on a curved road with radius of curvature
ρ = 100 ft. The road is also banked by 10◦ . The center of mass height is h = 22 in. Friction
is sufficient to prevent slip. Calculate the wheel loads when the vehicle speed is 45 mph.
You can model the vehicle as an axle.
Vehicle Dynamics—Basic Loads and Longitudinal Motions 669

Section 12.9—Weight Shift and Statical Indeterminacy


12.26 (E) By placing scales under all four tires, the wheel loads of a vehicle of wheelbase
2.7 m and track 1.05 m are observed to be the following: left front 600 kg, right front 580
kg, left rear 415 kg, right rear 405 kg. Calculate the location of the center of mass.
12.27 (M) In a test to calculate the center of mass height, a vehicle of weight 3000 lb,
wheelbase 9 ft and b = 4 ft is hoisted to an angle of 30◦ . The front axle load is measured
to be 1970 lb. Given that both wheels have a loaded radius of 1 ft, calculate the center of
mass height.
12.28 (M) A vehicle has the following dimensions: wheelbase 2.70 m, track 1.3 m. You are
informed that the weight of the vehicle, 2100 kg, acts through a point that is 1.55 m in
front of the rear axle and 0.58 m from the left side of the car. Calculate the wheel loads.
12.29 (M) A 3500 lb vehicle (wheelbase 9.5 ft, track 4 ft, CG height 2.4 ft, c = 5 ft from
rear), is traveling at a speed of 45 mph on a curved level road with radius of curvature of
200 ft, while decelerating at the rate of 0.2g. Calculate the wheel loads.
12.30 (M) A vehicle of mass 2000 kg, wheelbase 2.5 m, center of mass 1.1 m from the
front wheels, and height of 45 cm, is taking a right turn with speed 30 km/h and radius
of curvature of ρ = 40 m. If at the same time the brakes are applied to give the vehicle a
deceleration of 0.15g, calculate the wheel loads according to the approximate relationships
derived in this chapter.
12.31 (M) A 2800 lb vehicle (wheelbase 2.8 m, track 120 cm, CG height 58 cm, c = 1.5 m
from rear), is traveling at a speed of 80 kph on a curved level road with radius of curvature
of 120 m, while accelerating at the rate of 0.22g. Calculate the wheel loads.
12.32 (M) For the problem above, what should be the acceleration of the vehicle so that
one of the wheels will lose contact with the road?
12.33 (M) A hoist test is conducted to determine the center of mass height of a vehicle
with the following parameters: W = 3000 lb, L = 9 ft, Wf = 2000 lb, h − h1 = 1.5 ft.
The rear axle of the vehicle is hoisted by angles of 5◦ , 10◦ , 15◦ , 20◦ , and the following front
wheel loads are measured (measurements include some small random error) 2073.9, 2144.8,
2220.7, and 2303.3 lb, respectively. Estimate the center of mass height.
13
Vehicle Dynamics—Tire and Aerodynamic Forces

13.1 Introduction
This chapter discusses the two most important forces that act on vehicles: tire forces and
aerodynamic forces. Tires provide contact between the vehicle and the road surface, and
while traveling a vehicle encounters aerodynamic forces. We study the effects of tire forces,
aerodynamic forces, and rolling resistance forces on acceleration, braking, and wheel slip.
The analysis of tires is a complex topic and the contents of this chapter at best serve as
an introduction. The interested reader is referred to the texts by Gillespie, Jazar, Milliken
and Milliken, and Pacejka for more details.

13.2 Tires
Pneumatic tires are the most common way of creating contact between a land vehicle and
the road. Tires combine the elastic properties of rubber (natural or synthetic) with that of
compressed air to accomplish a variety of tasks, such as
• Maintain contact with the ground, and generate vertical forces (wheel loads) to support
the weight of the vehicle;
• Generate tractive forces that accelerate and decelerate the vehicle;

• Generate lateral forces that provide steering control and directional stability and enable
the vehicle to take a turn or to resist a side force;
• Provide cushioning and absorb forces due to road imperfections, such as bumps and
potholes; and
• Provide aligning torque and camber that improve stability.

A tire accomplishes all these tasks through its contact with the road. The area of contact
is called contact patch or tire print. The size of the contact patch is a design issue. Too
large a contact patch leads to high frictional forces and loss of traction, while a small tire
print leads to a bumpy ride, low frictional forces, and reduced lateral stability.
The pneumatic tire was first developed over 100 years ago. Mathematical models of tires
are complex and beyond the scope of this text. A pneumatic tire is constructed of an inner
torus, made of high-strength tensile cords, with a rubber covering over the threads. The
torus usually is manufactured using radial ply, as opposed to bias ply that was used in the
1950s. Radial ply tires have better cornering capability due to less roll under the treads,
better directional stability, and less distortion of the contact patch.

671
672 Applied Dynamics

13.2.1 Tire Terminology

Tireprint width
(thread width)
0 R 15 9
05/7 5S
P2

Sidewall
Sidewall height

Rim
Sidewall
Rim
Tire width

FIGURE 13.1
Side view and cross section of a tire.

Tires come in a variety of diameters, widths and aspect ratios. Figure 13.1 shows a side
view and cross-section of a tire. In essence, a tire can be viewed as consisting of a tread
surface and sidewall. Terminology using both U.S. and SI coordinates has been adopted to
describe tire size and load rating. This information is embossed in raised lettering on the
sidewall. Rating of a tire looks like P205/70 R15 95S. The different symbols indicate the
following:
P: Designation for passenger car. For a light truck, the designation is LT.
205: Tire width (not tread width) in mm.
70: Aspect ratio in %. The aspect ratio is the ratio of the sidewall height to tire width.
R: Indicates that the tire is radial. Bias ply tires are marked by a B.
15: Rim diameter, in inches. Rim radius plus sidewall height gives the radius of the unloaded
tire.
95: Load rating, indicating the expected weight that is to be carried by the tire. This ratio
ranges from 0 to 279, with passenger car tires having values in the range 75–100. The load
rating is related to the Gross Axle Weight Rating (GAWR) for the front and rear axles.
The GAWR numbers are usually printed on a label located on the driver’s side door pillar.
S: Speed rating in letters (A–Z), with each letter corresponding to a maximum recommended
speed for the tire. For passenger cars, the commonly found ratings (in mph) are P = 94, Q
= 100, R = 106, S = 112, T = 118, U = 124, H = 130.
Tires need to display the following additional information (on the sidewall) known as
the UTQG (Uniform Tire Quality Grading):
Tread Wear Rating: Indicates the projected tread life under normal operating conditions.
The tread wear rating is with respect to other tires from the same manufacturer. In general,
the tread wear rating multiplied by 100 gives the expected life of a tire in miles.
Traction Rating: Describes straight-line stopping ability on wet concrete and asphalt.
Uses A, B, and C scale, with A being the best. This index does not measure handling and
cornering capability.
Temperature Rating: Describes a tire’s ability to dissipate and resist heat, on a scale of
A, B, and C. A tire rated A (best) must withstand a half-hour run at 115 mph without
failure.
Vehicle Dynamics—Tire and Aerodynamic Forces 673

Tires also have the following ratings (usually this information is in smaller letters and
is located closer to the rim):
Radial Construction: Most passenger tires are radial. This has to be stated on the tire.
Tubeless: Most passenger cars have tubeless tires. The tire must state that it is tubeless.
If there is a tube inside the tire, the type of tube must be indicated.
Department of Transportation Code: In the U.S., all tires must have a designation in
the form DOT XXXX XXX. DOT means the tire meets the Department of Transporta-
tion safety standards. The next characters give the tire serial numbers which identify the
manufacturer, plant, type of tire construction, and date the tire was manufactured.
Tires are optimized for the expected use of the vehicle on which they are mounted. A
popular type of tire used on passenger cars is all-season tires, which are designed to give
reliable performance on dry, wet, and snowy surfaces. The designations M/S, M+S, or M&S
state that the tire meets the Rubber Manufacturer’s Association (RMA) definition for mud
and snow. Some manufacturers use the number 4 to denote all-season tires.

13.2.2 Tire Components


A pneumatic tire has the shape of a toroid and it is filled with compressed air. Components
of a tire can be categorized as follows:
• Outer Elements, which consist of the sidewall and the tread. The tread comes into
contact with the road and thus generates the tire forces. In general, a tread pattern
consists of lugs (or blocks), which maintain contact with the ground, and grooves, also
known as voids. While grooves reduce the contact area between the tire and the road,
grooves also let water, snow, mud, or other pieces of matter escape from the tire through
the effect of centrifugal forces generated by the rotation of the wheel.
Race cars have “shaved” tires, which have no grooves. This increases the contact area
between the tire and the road, which increases friction. On the other hand, such tires
cannot deal with rain, road imperfections, and loose objects on the road. Car races are
usually stopped when it rains. Racetracks must be kept very clean.
Wide and straight grooves are the most commonly encountered tread pattern. They
develop more lateral friction and also lead to a relatively quieter ride. Lateral grooves
are found in snow tires and tend to increase traction but also are noisier. Most tires have
a combination of straight and lateral grooves, and some also have V-shaped grooves.
Components of the treads of a tire are shown in Figure 13.2. Tires with large lugs, are
preferred for vehicles designed to travel in rough terrain and in muddy conditions. Tires
with large lugs provide better traction and better drainage of water, but they are noisy
and they wear out faster. The void ratio is the amount of open space in the tread. Sipes
are small grooves that are cut in the shape of slits, and they add to the flexibility of the
lugs and thus help traction. Dimples are small indentations that help with the cooling
of tires.
The sidewall provides an enclosure for the tire, keeps air from escaping, helps in the
formation of the vertical forces, and provides lateral stability. In lower aspect ratio tires,
reduction of sidewall height improves the structural integrity by making sidewall collapse
less likely. In addition, shorter sidewalls can accommodate larger rim diameter without
increasing tire diameter. A larger rim is desirable for anti-lock braking systems.
• Inner Elements, which consists of the inner liner, which protects the carcass and
prevents air from escaping, and the bead bundle or bead. The bead bundle is made of
high strength steel cable that is coated with rubber, providing support for the contact
between the tire and the rim and transmitting forces from the tire to the rim.
674 Applied Dynamics

Grooves Ribs
Dimples Shoulder
Sipes

Lugs Voids

FIGURE 13.2
Components of tire tread. Permission by Chris Longhurst, www.carbibles.com.

• Middle Elements, which consists of the carcass, which is made of low modulus of
elasticity rubber with high modulus of elasticity cords in a radial pattern. The cords can
be of metal, such as steel, or of synthetic material. The carcass is the main component
that supports loads. The cords are attached to the bead. There also some belts that are
outside the circumference of the carcass. The belts enhance contact with the ground, add
strength and improve tread wear. Tires meant for passenger cars have one carcass, while
tires that take larger forces, such as airplane tires, have multiple layers of carcasses.

13.3 Tire Forces


Tire forces are generated due to friction coupling between the tire and road. There are two
primary sources that generate tire friction forces and two secondary sources:
• Hysteresis is a material property that results in loss of energy because the force versus
deflection curve is different for loading and unloading (or for expansion and restitution).
Hysteresis is discussed in more detail in Section 5.5.7. A typical hysteresis curve is
shown in Figure 13.3 and has the form of a loop. The area inside the loop is the lost (or
dissipated or unrecoverable) energy.
In a tire, hysteresis manifests itself as energy loss as the tire deforms repeatedly while
the vehicle is moving. This lost energy is converted into heat, which is one reason why
tires heat up while driving.1 The rolling resistance force that acts on tires is primarily
generated by hysteresis, as described in Section 13.7.
High hysteresis tires, such as those in race cars, are designed for higher friction so that
the race car can accelerate or decelerate more rapidly. As expected, such tires wear out
faster. Tires on passenger cars are designed with less hysteresis so that they will have
lower rolling resistance, less wear and lower temperatures.
1 Another reason is the finite area of the contact patch, and sliding friction in a part of the contact patch.
Vehicle Dynamics—Tire and Aerodynamic Forces 675

Force
Loading

Unloading

Deflection

FIGURE 13.3
Hysteresis in force vs. deflection curve.

• Adhesion arises from the intermolecular bonds between the rubber and aggregate in
the road surface, which results in some amount of sticking between the tire and road.
The adhesion properties of tires are designed depending on the application, such as race
car vs. passenger vehicle.

FIGURE 13.4
Tire-road friction due to hysteresis and adhesion.

On a dry road, adhesion contributes more to friction (shear stresses) than does hysteresis
(normal stresses). However, a wet surface affects adhesion much more than hysteresis,
so on a wet road tire friction is mainly due to hysteresis and the other two friction
mechanisms discussed below. Hysteresis and adhesion are depicted in Figure 13.4.
• Deformation and Wear. Deformation friction takes place when the rubber in the
contact patch deforms from its original shape and assumes the shape of the irregularities
on the road surface. This increases the friction between the tire and road. Wear friction
is due to the large stresses that the rubber in the tire experiences. These high stresses,
when they are beyond the elastic limit of the rubber, tear the tips of the rubber treads
676 Applied Dynamics

a small amount. The torn rubber has a higher area of contact with the road and hence
higher friction is generated. Deformation and wear friction are less affected by wet roads
than is adhesion.
The coordinate system for the contact patch used here is based on the SAE (Society of
Automotive Engineers in the U.S.) tire coordinate system, which has the same directions as

 


FIGURE 13.5
Contact patch and SAE coordinate system for tires.

the vehicle coordinate system (x–longitudinal, y–lateral, and z–vertical positive down), as
shown in Figure 13.5. A reference point, or origin of the coordinate system, is needed and
it is placed at the middle of the contact patch. The wheel plane is obtained by rotating the
wheel by the inclination angle, also known as camber, about the x axis. Basic forces acting
on a tire are shown in Figure 13.6. Forces that are generated by a tire are


 
      
  


  
  


  

 




FIGURE 13.6
Wheel plane and basic forces acting on tires.

• In the longitudinal, or travel, (x) direction: Tractive force FT , due to acceleration or


braking; rolling resistance force FRR , due to energy loss at contact patch.
Vehicle Dynamics—Tire and Aerodynamic Forces 677

• In the lateral (y) direction: Lateral force FL (cornering force), due to taking a turn or
reaction to a lateral force, such as wind gust or force of gravity on a banked road, as
well as camber thrust (lateral force due to having camber).
• In the vertical (z) direction: Normal force FN , also known as wheel load, which results
from the weight carried by the tires and forces generated due to road imperfections,
such as bumps.

13.3.1 Resultant Tire Forces and Moments


Forces generated by tires are due to contact between the tires and the road surface; hence,
tire forces are reaction and friction forces. Tire forces act over the entire contact patch. Even
when the contact patch is small, the force distribution and resulting stress distribution are
not uniform along the contact patch. The resultants of tire forces usually do not act through
the center of the contact patch.
Rather than deal with the contact patch geometry and resultant forces that do not act
through the center of the contact patch, it is convenient to describe the effects of tire forces
as a resultant force going through the center of the contact patch and a resultant moment.
This is the same type of analysis conducted in Chapter 4, where the combined effects of
forces and moments acting on a body are viewed. Consider Figure 13.7a, depicting a generic
force f (x) (0 ≤ x ≤ c) acting along a patch of length c. The resultant force F and resultant
moment MO are calculated as
Z c Z c
F = f (x) dx MO = xf (x) dx (13.1)
0 0

The net effect of the acting force, shown in Figure 13.7b, is represented as a resultant
force F going through point D, which is a distance d from O (the origin of the point through
which the resultant force acts), where d can be calculated from

a) b) c)
f(x) f(x) f(x)

= F
= F
M = F(d – c/2)
x x x
O O c/2 Dc O c/2 c
c
d

FIGURE 13.7
a) Generic force, b) resultant force, and c) resultant moment.

Z c
MO 1
d = = xf (x) dx (13.2)
F F 0

The net effect of the acting force can also be represented by the force F acting through
the center, x = c/2, and a moment of magnitude M = F d − 2c , as depicted in Figure
13.7c. We will use this approach to model resultant tire forces and moments in subsequent
sections. The advantages of using the resultant tire forces and moments are

• Ease of dealing with resultant point forces, instead of continuous functions;


678 Applied Dynamics

• Simplification resulting from carrying the resultants to the middle of the contact patch;

• Clearer visualization of the net effects of the forces and moments that are acting.

13.4 Lateral Forces and Tire Slip


Our study of tire forces begins by analyzing lateral forces and how these forces prevent a
vehicle from sliding sideways. Tire slip is one of the most interesting characteristics of tires
and how tires generate lateral reaction forces.
We must exercise caution when using the word slip, as this word is used in different
contexts in vehicle dynamics. Chapters 5 and 11 discussed roll without slip, as well as roll
with slip and sliding. In vehicle dynamics, the term slip is encountered in two additional
situations: a) when studying lateral (or sliding) forces on a tire we deal with slip angles, and
b) when studying the longitudinal motions and spinning of tires, a quantity called slip ratio
is defined. In addition, the term sideslip is defined in Chapter 14 to describe the lateral
motion of the center of mass of a vehicle.

W x W x
a) b) c)
y
z
Fy v x y
Fy Fy FL

FL y

FN FL FN
Rear view Side view Top view

FIGURE 13.8
a) Rear view of tire acted upon by a lateral force Fy , b) side view, c) top view.

Consider the rear view of a tire in Figure 13.8a, which is acted upon by a force Fy in the
lateral direction. Examples of this force include a side force due to wind, or the component
of gravity when the tire is on a banked road, or a lateral force transmitted to the tire by
the body of the vehicle.
The most common lateral force is the inertia force associated with the normal accelera-
tion that results when taking a turn. In keeping with our sign convention, consider a turn
to the right (rotation about the positive z axis), where the lateral acceleration is in the y
direction, or a = an j = v 2 /ρj, with ρ denoting the radius of curvature of the turn. In vehicle
dynamics, the radius of curvature is also referred to as the turn radius. Consequently, the
inertia force on the tire is in the opposite direction. In vector form, Fy = −man j.
To counter the inertia force acting at the center of the tire, a lateral tire force FL is
developed through the contact between the tire and the road. This lateral force acts in the
opposite direction to the inertia force, so that it is in the positive y direction. In vector
form, FL = FL j. As a result of the vertical force (weight of the car) and the lateral tire
force FL , the tire will distort to the right and will assume a shape like the one shown in
Vehicle Dynamics—Tire and Aerodynamic Forces 679

Figure 13.8a. For steady-state motion, the tire does not slide along the lateral direction (y),
and FL = Fy . The side and top views of the tire are shown in Figure 13.8b and Figure
13.8c, respectively.
Next, consider the longitudinal motion of the tire. There are several ways of visualizing
this. One is to consider that you are walking along a straight line and you are subjected
to a continuous side force. Even if the side force does not make you slide, you will follow a
trajectory that is different from your heading, as shown in Figure 13.9.

v
Path

x
x Heading

y
FL
Lateral force

FIGURE 13.9
Top view of change in path due to a lateral load. Source: Race Car Vehicle Dynamics, by
Milliken and Milliken, redrawn with permission.

Similarly, due to the application of a side force, the wheel will follow a path that is
different from the heading of the tire, as shown in Figure 13.10. The velocity of the tire
is along the wheel path. The angle that the wheel path makes with the wheel heading is
known as the slip angle and is denoted by α. Here, the tire is not really slipping or sliding,
but its velocity is deviating (slipping) from its heading. A similar terminology will be used
in Chapter 14, when considering the sideslip angle.

v x (Wheel heading)
Travel 
direction x

Fy FL

FIGURE 13.10
Top view of tire path due to slip.

The slip angle is a velocity ratio. It usually is a small quantity, enabling us to use the
small angles approximation. Denoting the longitudinal velocity of the tire by V and the
680 Applied Dynamics

lateral velocity by v, we write the velocity of the tire as v = V i + vj, and the slip angle can
be expressed as

v v
α = tan−1 ≈ (13.3)
V V

FL
1 2 3

Elastic or Transitional Frictional


linear

Slip angle for


peak lateral force
Slope (C)

1 2 3 4 5 6 7 8
Slip angle (deg.)

FIGURE 13.11
Lateral force FL vs. slip angle α. Source: Race Car Vehicle Dynamics, by Milliken and
Milliken, redrawn with permission.

It is of interest to measure the lateral tire force FL as a function of the slip angle. This
relationship is usually measured experimentally. When the results are plotted, a curve that
looks like Figure 13.11 is obtained. We can split the curve into three regions:

• In the first region, referred to as the elastic or linear region, there is an almost linear
relationship between the lateral force and slip angle. The constant of proportionality is
called the cornering stiffness and is denoted by Cα , so we can write

FL = Cα α (13.4)

The cornering stiffness is usually expressed in the units of lb/◦ , lb/rad, N/◦ , or N/rad.
The cornering stiffness is affected by a large number of factors. These factors are sum-
marized in Table 13.1. These relationships are not linear and they are interdependent.
In the linear range, the lateral force is generated entirely by the elasticity and slip
characteristics of the tire. For most passenger tires, the linear range ends around α ≈
3–4◦ . A ballpark estimate of a passenger car tire cornering stiffness is about 15% to 20%
of the wheel load per degree.
• The second range in the lateral force vs. slip angle curve is called the transitional stage.
Here, the lateral force does not increase as much as it did in the linear region, as the
lateral force generation capability of the tire decreases. Recall that the lateral force is
generated by slip as well as friction between the tire and the road. The lateral force
reaches its maximum value at the end of this region. For passenger tires, the transition
region ends around α ≈ 6◦ .
Vehicle Dynamics—Tire and Aerodynamic Forces 681

TABLE 13.1
Summary of factors that affect cornering stiffness

Increase in or Having Effect on Cornering Stiffness


Radial (vs. bias ply) Increase
Inflation pressure Increase
Aspect ratio Decrease
Tire size Increase
Rim size Increase
Tire temperature Increase or decrease
Vertical tire load Increase
Vehicle speed Decrease

• The third region of the curve is called the frictional stage, as the lateral force is gener-
ated primarily by friction. Here, the magnitude of the lateral force, after reaching the
maximum value of FL = µs FN , where µs is the friction coefficient between the road
and tire, begins to fall as the slip angle increases. The tire begins to slide and becomes
harder to control. It is not a good idea to drive in this region.

The vertical load on a tire affects the lateral forces that are generated. It is informative
to plot the lateral force vs. slip angle for different values of the tire load FN . Such a plot
is given in Figure 13.12. There is wide variation in the lateral force as a function of the
vertical load. This suggests consideration of the ratio of lateral force to the vertical load.
The nondimensional lateral force coefficient is defined as
FL
cL = (13.5)
FN

FL Line of peaks

Increasing
wheel load

1 2 3 4 5 6 7 8
Slip angle (deg.)

FIGURE 13.12
Lateral force for varying vertical loads.

In essence, the lateral force coefficient is a measure of the friction force generated by the
tires, as a coefficient of friction is the ratio of a horizontal force to a normal force (wheel
682 Applied Dynamics

load). Figure 13.13 plots cL vs. the slip angle, for different values of the vertical load FN
and it leads to two interesting observations:
• The curves are closer to each other than in the case of the lateral force vs. slip angle
plots (Figure 13.12),
• The lateral force coefficient cL becomes smaller for higher wheel loads, implying that
the lateral force generation capability of a tire becomes smaller at higher loads. This
phenomenon is known as tire load sensitivity and it is an important design issue. If
a vehicle is overloaded, the lateral forces generated become progressively smaller with
respect to the wheel loads and may not be able to provide the necessary lateral forces
for cornering.

cL Line of peaks
1.1
1.0
0.9 Increasing
wheel load
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
1 2 3 4 5 6 7 8
Slip angle (deg.)

FIGURE 13.13
Lateral force coefficient cL for varying loads.

A generally accepted nonlinear model for the cornering stiffness Cα as a function of the
wheel load FN is

Cα = e1 FN − e2 FN2 (13.6)

in which e1 and e2 are positive coefficients, whose values depend on the tire properties.

13.5 Tire Torques


As discussed earlier, it is preferable to describe the effects of the tire forces acting on the
contact patch in terms of a resultant force going through the center of the patch and an
associated tire moment. In this section, two such moments are analyzed, one arising from
the lateral force and the other from the vertical force. A third resultant moment will be
studied in conjunction with rolling resistance in Section 13.7.
Figure 13.14 shows lateral and vertical force distributions and the resultant forces FL
Vehicle Dynamics—Tire and Aerodynamic Forces 683

v x
Rl
z

y
FL
x
tp
tp x
FL y
-z FN

 x

tr x tr
y
z
FN

FIGURE 13.14
Lateral and vertical force distribution along contact patch.

and FN . Both these forces are not symmetrically distributed about the center of the contact
patch, producing resultant moments. Denoting the distance of the resultant lateral force FL
from the y axis by tp , the moment that this lateral force produces is called the aligning
torque Mz

M1 = −tp i × FL j = −tp FL k = −Mz k (13.7)

in which Mz = tp FL . The resultant normal force (wheel load) FN crosses the xy plane at
r = tr i + j, so the resultant moment generated by the wheel load is

M2 = (tr i + j) × −FN k = −FN i + tr FN j = −Mx i + My j (13.8)

The moment Mx = FN is called the overturning moment and My = tr FN is the rolling
resistance moment.

13.5.1 Aligning Torque


Consider the resultant lateral force FL , which acts through a distance tp from the y axis,
as shown in Figure 13.15. The net effect can also be expressed as FL going through the
center of the contact patch and a resulting moment Mz = tp FL . The resulting moment is
called aligning torque and it has a caster effect. It rotates the tire so that the tire aligns its
heading toward the direction of wheel travel. This is a desirable feature and it is an added
benefit of pneumatic tires.
The resultant moment Mz is counterclockwise when viewed from the top, as seen in
Figure 13.15. The distance tp is known as the pneumatic trail. The magnitude of the pneu-
matic trail depends on the slip angle. Figure 13.16 plots the aligning torque as a function
684 Applied Dynamics


 

  

 
 

FIGURE 13.15
Top view of contact patch and aligning torque.

of the slip angle for various values of wheel loads. The aligning torque reaches a peak value
at around α = 2–4◦ , after which it drops. At high slip angles, the aligning torque may even
become negative, due to a change in the lateral force distribution on the contact patch.

Mz Line of peaks

Increasing
wheel load

1 2 3 4 5 6 8 9
Slip angle (deg.)

FIGURE 13.16
Aligning torque as a function of slip angle.

Denoting the mechanical trail, which is due to the steering geometry, by tm , the cu-
mulative caster effect in a tire is defined as the steer torque = (tp + tm ) FL . The following
observations associated with mechanical and pneumatic trail are of interest:
• In general, the mechanical trail is larger than the pneumatic trail.

• A good driver can sense the reduction in the steer torque when the pneumatic trail
becomes smaller as the result of taking a turn at high speed and the generation of high
slip angles.
• In older cars, where the mechanical trail was smaller than in contemporary cars, the
aligning effect on the front tires was due primarily to the pneumatic trail. When such
a car was driven at high speeds, the steer torque become small or even went negative,
resulting in reduction in stability.
Vehicle Dynamics—Tire and Aerodynamic Forces 685

13.5.2 Overturning Moment


The lateral tire force FL distorts the shape of the tire print in the contact patch. For a right
turn the tire print shifts to the right, which has the effect of shifting the resultant of the
vertical tire force FN to the right by a distance , as shown in Figure 13.17.

Mx

y
=
z

FL

FN FN

FIGURE 13.17
Vertical force shift and overturning moment (rear view of tire).

The effect of this load shift is a resultant force going through the center of the contact
patch, and a moment Mx about the negative x axis, referred to as the overturning moment

Mx = FN (13.9)

The overturning moment has a destabilizing effect, as it induces roll. However, the
overturning moment does not have a large magnitude, and it is generated only when large
lateral forces are applied. It also is generated when there is camber.

13.6 Slip Ratio and Longitudinal Tire Forces


There are three types of forces generated by tires in the longitudinal (tractive, travel)
direction: acceleration, braking, and rolling resistance. Similar to lateral and normal forces,
the load distributions due to tractive forces in the longitudinal and vertical directions along
the contact patch are not uniform. Before studying the longitudinal tire forces, we need
to consider that the shape of a tire is not a perfect circle while the tire is in contact with
the road. The amount of contact, as measured by the area of the contact patch, affects the
tractive as well as the lateral forces.

13.6.1 Wheel Slip


Let us first look at slip ratio. Considering Figure 12.12, two measures of the tire radius are
introduced:
686 Applied Dynamics

• Effective Radius. Denoted by Re , the effective radius is a measure of how much a


wheel travels in one revolution. Denoting the distance travelled by the center of the
d
wheel in one revolution by d, the effective radius is found using Re = 2π .
• Loaded Radius. Denoted by Rl , the loaded radius is the perpendicular distance be-
tween the wheel center and the road surface.

While the effective radius is the mathematical quantity used to define slip ratio, the loaded
radius is a quantity that is easier to measure under operating conditions.
Let us denote the speed of the center of the wheel by V and the angular velocity of the
tire by Ω. The SAE definition for slip ratio, denoted by SR, is
ΩRe
SR = −1 when α = 0 (13.10)
V
Slip occurs because the contact between the tire and the road is a finite area and the
tractive forces that are generated by acceleration and braking may exceed the maximum
forces that the tires can generate. The following slip ratio definitions are commonly used:
SR = 0 =⇒ Free rolling, as V = ΩRe for this case
SR = −1 =⇒ Locked brakes, as Ω = 0 for this case
SR = 1 =⇒ Defined as spin (this is an arbitrary definition). At this value of SR,
V = ΩRe /2.
Slip ratio can also be defined in terms of the loaded radius, and many people prefer to
do this, as force and moment calculations for a tire involve Rl . The Calspan TIRF (Tire
Research Facility) definition of slip ratio is

ΩRl
SR = −1 when α = 0 (13.11)
V
It should be emphasized that, even for free rolling, the distance traversed in one revolu-
tion is less than 2πR. Also, in the presence of a slip angle α, the two slip ratio definitions
above are calculated by replacing V with V cos α.

13.6.2 Tractive and Lateral Tire Forces


Acceleration and braking forces are friction forces generated when a driving or braking
torque is applied to the wheel. The friction force opposes the spin of the wheel and generates
the driving or braking force.
Denoting the acceleration and braking forces by FA and FB , respectively, the tractive
force distribution along the contact patch is shown in Figure 13.18. Acceleration and braking
forces generate shear stresses. Similar to other forces generated inside the contact patch,
the tractive force distribution is not uniform.
Both tractive forces, acceleration and braking, vary with the slip ratio. As the slip ratio
becomes higher, the maximum amount of tractive force that can be generated increases,
until a slip ratio of SR = 0.10–0.15 is reached. The maximum tractive force decreases after
that. An accelerating wheel spins-up (spins in place) for higher slip ratios, as a result of
which traction is reduced. A braking wheel approaches locking for negative slip ratios.
Figure 13.19 is a typical plot of tractive force FT versus the absolute value of the slip
ratio, |SR|. We use the absolute value, because plots for acceleration and braking are nearly
identical. As expected, the tractive forces that can be generated become lower as the slip
angle α increases.
Slip ratio also affects the lateral forces that can be generated. An increase in slip ratio
reduces the lateral forces that can be applied. Figure 13.20 shows the reduction in lateral
Vehicle Dynamics—Tire and Aerodynamic Forces 687

v
Rl

FB FA
FA

x
FB
x

FIGURE 13.18
Acceleration and braking forces in tractive direction.

forces as a function of the slip ratio. Plots are given for different slip angles α. As expected,
increasing the slip angle increases the lateral forces. As the slip angle increases, higher
lateral forces are generated. However, increasing the slip ratio uses up some of the available
friction and hence reduces the lateral forces. Note that the lateral forces are different for
acceleration and braking.

13.6.3 More Comprehensive Tire Models


Tire forces, whether tractive or lateral, depend on a variety of factors. The main quantities
that affect tire forces are the vertical load FN , slip ratio SR, slip angle α, and camber
angle γ. These factors affect the tire forces in a nonlinear way and are difficult to model
analytically. In general, they are ascertained from experiments.
When interpolating experimental data and analyzing how different factors affect the tire
loads, a commonly used empirical relationship is called the magic formula, also known as
the Pacejka model. This model has the form
F = c1 sin c2 tan−1 c3 x (1 − c4 ) + tan−1 c3 x)
 
(13.12)
where F is the force (lateral or tractive); c1 , c2 , c3 , and c4 are parameters; and x is the
variable. The variable x is the slip ratio SR when modeling the tractive force, and it is the
slip angle α when modeling the lateral force. The Pacejka model can also be used to model
tire moments, such as aligning torque. More details about the Pacejka model can be found
in texts on tire dynamics.

13.7 Rolling Resistance


The rolling resistance force acts along the wheel heading and always opposes motion. It
is the result of a variety of factors, and it is a complex force that is difficult to quantify.
688 Applied Dynamics

FT
Increasing slip angle ( )

0.10 |SR|

FIGURE 13.19
Tractive force FT versus the absolute value of the slip ratio, |SR|, for varying values of the
slip angle α.

FL

Increasing
slip angle ( )

SR
Acceleration Braking

FIGURE 13.20
Lateral force FL versus slip ratio for varying values of the slip angle α.

Rolling resistance is usually modeled by empirical relationships. We describe some of these


relationships in Section 13.7.3.
The most important factor that contributes to rolling resistance is energy loss due to
hysteresis during the deformation of a tire as the tire comes into contact with the ground.
Figure 13.21 describes the geometry. Following the motion of a point on the tire, as the
point comes into contact with the road, the tire compresses and the contact point begins
to be pushed back towards the center of the tire. When the point is at the center of the
contact patch, its distance from the center of the wheel is reduced to Rl , the rolling radius.
This is the compression stage of the tire contact with the road.
As the point being followed moves beyond the center of the contact patch, it begins to
decompress. It returns to its undeformed position and leaves the contact patch, where its
distance from the center of the wheel again becomes R. This phase is the expansion (or
restitution) stage.
Because of the hysteresis in the tire and treads, the total vertical force during the expan-
sion stage is less than in the compression stage, similar to what happens in the compression
Vehicle Dynamics—Tire and Aerodynamic Forces 689

v
R Rl R

Smaller force Larger force


Expansion Compression
range range

FIGURE 13.21
Tire deformation geometry and associated forces.

and restitution stages of impact. The vertical force profile is shown in Figure 13.22 (the free
rolling plot) and it is skewed forward. On a bumpy road, there is further distortion of the
tire due to bumps, which adds to the vertical force imbalance.

Torques
Braking

Driving

v
Rl

FN
Free Braking
rolling
x
tr
z

FIGURE 13.22
Vertical force distribution along contact patch.

It follows that the resultant vertical force (wheel load), FN , lies toward the front of the
contact patch, at a distance tr from the center of the contact patch, as shown in Figure
13.23a. The resultant is a force FN going through the center of the contact patch and a
counterclockwise moment My = tr FN , referred to as the rolling resistance moment (Figure
13.23b). The rolling resistance moment acts in the same direction as a braking torque.
The rolling resistance moment leads to a friction force acting on the contact patch to
counter the angular acceleration that the rolling resistance moment causes. This force is in
the same direction as a brake force. It is called the rolling resistance force FRR and it is
690 Applied Dynamics

a) b) My = trFN c)

x v v
y z Re Re My
Re

M
tr
FRR = R y
e
FN FN FN

FIGURE 13.23
a) Vertical force FN , b) resultant moment My , c) rolling resistance force FRR .

shown in Figure 13.23c. We obtain the rolling resistance force as


My FN t r
FRR = = (13.13)
Re Re
Because several other factors also contribute to slowing the vehicle and because it is
small, the rolling resistance force is customarily modeled in terms of a rolling resistance
coefficient fr as

FRR = fr W (13.14)

where W is the weight of the vehicle. While the magnitude of the rolling resistance force is
small, the rolling resistance force acts on the vehicle at all times while the vehicle is moving,
so its cumulative effect is not negligible.

Example 13.1
The rolling resistance coefficient of a 3500 lb vehicle is fr = 0.015. Calculate the rolling
resistance force and the power that is needed to counter rolling resistance at a speed of 60
mph.
The rolling resistance force is

FRR = fr W = 0.015 × 3500 = 52.5 lb [a]

At a speed of 60 mph (88 ft/sec) the power requirement is

P = FRR v = 52.5 × 88 = 4620 ft · lb/sec [b]

Converting to horsepower gives


P 4620
Php = = = 8.4 hp [c]
550 550
While this power requirement is not very large, it is always there, and its cumulative
effect increases fuel consumption.
Vehicle Dynamics—Tire and Aerodynamic Forces 691

13.7.1 Factors Affecting Rolling Resistance


The primary factors that affect rolling resistance, in addition to hysteresis, are
• Tire temperature. A cold tire contributes more to rolling resistance than a hot one.
A tire has less rolling resistance after it has been driven more than 20 miles. In general,
a trip of 20 miles results in a 30◦ F increase in tire temperature.
• Road surface. The surface of a road has a significant effect on rolling resistance. Rolling
resistance is lowest on concrete surfaces. Asphalt surfaces lead to slightly higher levels
of rolling resistance. Soft surfaces, such as sand, result in the highest levels of rolling
resistance. The softer the surface, the more penetration the tire has into the road surface
and the higher the rolling resistance.
The roughness and nonuniformity of the road surface also add to the rolling resistance.
Radial tires have lower rolling resistance than bias-ply tires. Some representative values
are given in Table 13.2. It is interesting to note that for a train wheel traveling on a
clean track, the rolling resistance coefficient is around fr ≈ 0.0002.

TABLE 13.2
Representative values for the rolling resistance coefficient fr

Vehicle Type Road Surface


Hard Medium Soft
Concrete or Asphalt Unpaved Road Sand
Passenger Car 0.014–0.016 0.02–0.08 0.30
Truck 0.01–0.015 0.03–0.06 0.25
Tractor 0.02 0.04 0.20

• Tire pressure. Increasing pressure makes a tire stiffer, which has different effects on
rolling resistance depending on the road surface. On hard surfaces, such as concrete or
asphalt, rolling resistance becomes lower with increased pressure.
On the other hand, higher tire pressure increases rolling resistance on soft surfaces such
as sand. A stiffer tire can penetrate a soft surface more easily. Vehicles that are driven on
sand, for example, on beaches, have their tire inflation reduced to develop lower rolling
resistance and increased traction. However, lowering tire pressure too much results in
larger deformations of the sidewall, which creates stability problems and also adds to
rolling resistance.
• Vehicle velocity. Rolling resistance increases significantly at speeds over 60 mph, sug-
gesting a nonlinear model for rolling resistance. The primary reasons for this increase
are tire vibrations and standing waves in the tires. Tire vibrations are a significant factor
in limiting tire speeds.

• Wheel load. Increasing the wheel loads increases rolling resistance. This increase is
more pronounced at higher speeds. Increasing wheel loads is similar to decreasing pres-
sure.
• Tire quality and wear. A worn tire that has smoother treads will have lower values
of rolling resistance. Race cars, for example, use tires with light tread or no tread at
all. Tires manufactured of high hysteresis material will have higher values of rolling
692 Applied Dynamics

resistance. Tires with higher sulfur content tend to have a lower rolling resistance. The
auto industry refers to the three interrelated variables of rolling resistance, traction,
and tread wear as the magic triangle. Improving the desirable properties of one of these
variables leads to reduction in the desirable properties of the other two variables.
• Tire slip and scrub. The rolling resistance coefficient increases rapidly at slip angles
α > 2◦ . This is a reason why a vehicle loses speed when taking a turn.
Scrub is discussed in Section 15.19; it is the lateral motion of the tire print as the tire
moves up and down due to suspension effects and also due to the turning of the tire.
Someone turning the steering wheel while the vehicle is stationary experiences scrub.
Having more scrub increases rolling resistance.
• Tire material and aerodynamic drag. As discussed earlier, the primary source of
rolling resistance is tire hysteresis. Tires manufactured out of higher hysteresis material
will have higher values of fr . Aerodynamic forces in the vicinity of the tires resist the
forward motion and hence add to the rolling resistance.

13.7.2 Induced Drag

x
Travel FRR Path
direction 
v

FL cos
 y
FL sin
FL

FT

FIGURE 13.24
Top view of tire taking a right turn.

Figure 13.24 shows the free-body diagram of a tire viewed from the top, as the tire
makes a right turn. Of interest are the forces opposing motion in the travel direction. There
are three forces acting on the tire: the tractive force FT , the lateral force FL , and the rolling
resistance force FRR , which acts in the negative x direction. As the tire takes the turn, a
slip angle develops and the wheel heading and wheel velocity are not in the same direction.
Denote the direction of the wheel velocity by t (as in tangential direction) and approximate
the wheel velocity by v = V i − vj ≈ V et , where et is the unit vector in the travel direction.
Summing forces along the travel direction gives
X
Ft = (FT − FRR ) cos α − FL sin α = FT cos α − (FRR cos α + FL sin α) (13.15)
Vehicle Dynamics—Tire and Aerodynamic Forces 693

The first term on the right, FT cos α, is the thrust along the travel direction and, because
the slip angle is small, is almost the same magnitude as the tractive force. The second term,
FRR cos α + FL sin α, is the total value of the force resisting motion. It is composed of
the rolling resistance force and the contribution due to the lateral force. The contribution
from the lateral force, FL sin α, is known as induced drag. Noting that the lateral force is
proportional to the slip angle α and that from the small angle assumption sin α ≈ α, the
induced drag can be approximated as

FL sin α = Cα α sin α ≈ Cα α2 (13.16)

For example, if the cornering stiffness is 900 N/◦ per axle and the slip angle is 2◦ , the
induced drag is
2
Cα α2 = 900 × 2 × ≈ 63 N (13.17)
57.296
per axle. The total induced drag is not a negligible amount. Note that doubling the slip
angle increases the induced drag force fourfold.
The effect of induced drag becomes apparent when we enter a turn and notice the
reduction in the speed of the vehicle. A good driver lowers speed before entering a curve.
Towards the middle of the curve, the driver has to lightly depress the accelerator to make
up for speed lost to induced drag.

13.7.3 Rolling Resistance Models


Because rolling resistance is due to many factors, we usually model it empirically. Several
models to approximate rolling resistance have been proposed. The simplest rolling resis-
tance models assume a constant value for the rolling resistance coefficient fr , as shown in
Table 13.2. More advanced models use more than one coefficient, and they also include the
speed of the vehicle. One such model for cars, which is accurate at low speeds, is
 
V
fr = 0.01 1 + (13.18)
100

where V is the vehicle speed in mph.


Other models of rolling resistance have nonlinear relationships between rolling resistance
and speed, such as

fr = µ0 + µ1 V 2 (13.19)

where V is vehicle speed and µ0 and µ1 are coefficients. For most passenger tires, when V
is expressed in km/h, the range for µ0 is 0.009 to 0.015 (lower value for concrete and higher
for asphalt) and µ1 ≈ 0.4 × 10−7 (h/km)2 .
An empirical model adopted as the SAE standard J2452 for the rolling resistance coef-
ficient takes into consideration tire pressure and wheel load and has the form

fr = pc0 × FNc1 × c2 + c3 V + c4 V 2

(13.20)

where ci (i = 0, 1, 2, 3, 4) are coefficients and p is the tire pressure. Research on the subject
of rolling resistance is continuously evolving, as tire manufacturers seek to reduce rolling
resistance while at the same time increasing traction and reducing tread wear.
694 Applied Dynamics

Example 13.2—Dealing with Runaway Trucks


A truck that needs to slow down while traveling downhill must generate very large braking
forces. Trucks whose brakes cannot provide the necessary braking forces, or trucks whose
brakes fail, become runaway vehicles and can cause severe accidents. A 1989 Federal High-
way Administration report indicated that in mountainous areas, one sixth of the accidents
involving large trucks were due to brake failure which occurred as a result of overheating
brake pads and drums.
To deal with runaway trucks on downgrades, many highways have truck escape ramps.
These ramps are characterized by a) a high amount of uphill, such as 10% (tan θ = 0.1,
deceleration provided becomes 0.1g) and b) arrestor beds, which consist of a road surface
made of deep sand or loose gravel. A sand surface can result in a rolling resistance coefficient
of about fr = 0.25, providing a deceleration level of 0.25g.
The combined deceleration effect of the ramp and arrestor bed is 0.1g + 0.25g = 0.35g.
Consider a truck traveling at 90 kph (25 m/s) as it enters the truck escape ramp. Using a
constant deceleration model, the distance the truck travels before coming to a rest can be
found from
v2 252
s = = ≈ 91 m [a]
2a 2 × 0.35 × 9.807
It is interesting to note that once the truck comes to rest, it will not roll back. This
is because the horizontal component of gravity (0.1g) is smaller than the rolling resistance
(0.25g).

13.8 Camber
Camber is the inclination angle of a wheel relative to the vertical. If the wheel leans in
towards the chassis, it has negative camber; if it leans outward away from the car, it has
positive camber. In vehicle dynamics terminology, the terms inclination and camber are used
interchangeably. Figure 13.25 shows the sign conventions for camber and for inclination.
Camber is used in vehicles for two main purposes: providing an additional lateral force and
to help stabilize the suspension system when the wheel goes over a bump or after a sudden
vertical load.
The lateral force that camber generates is known as camber thrust. Camber thrust
always acts in the direction of lean of the tire. Considering Figure 13.25 as the rear view
of a vehicle taking a right turn, the wheel on the left is the outside wheel, and it carries a
heavier load as result of load shift due to the turn. As shown in Figure 13.12, the lateral
force that can be developed as the vertical loads become higher is not a linear relationship,
thus limiting the lateral forces. The total lateral force that can be generated by the inside
and outside tires becomes lower in the presence of load shift. Having an additional lateral
force due to camber on the outside tire is beneficial and reduces the possibility of sliding.
Because wheel loads shift to the outside wheels when taking a turn, the vehicle rolls
towards the outside wheels and the tires also roll in the same direction. Camber makes the
outside tire more upright. An upright tire has a larger tire print than an inclined tire. This
results in the camber thrust. In addition, an upright tire is more stable.
While camber also reduces the lateral force that is generated by the inside tire, this
reduction is not crucial, as the inside tire has a reduced wheel load and can generate a
small lateral force in the first place. At low camber angles, the camber thrust is linearly
Vehicle Dynamics—Tire and Aerodynamic Forces 695

 

FC FC –
Camber is –

Inclination angle is +

FIGURE 13.25
Camber nomenclature and camber thrust force FC (rear view of vehicle taking a right turn).
Source: Race Car Vehicle Dynamics, by Milliken and Milliken, redrawn with permission.

proportional to the camber angle, having the form

Fcamber = FC = Cγ γ (13.21)

in which Cγ is the camber stiffness.


Most passenger vehicles are designed with negative camber angles of 5◦ or less. In such
cases, the camber stiffness is usually around 10% to 20% of the cornering stiffness for most
tires, so that camber provides a 10% of an increase in the lateral tire force of the outside
wheel. For low slip and low camber angles, the effects of slip and camber can be treated
separately.
Because camber generates a lateral force, it further distorts the tire print. Figure 13.26a
shows the free-body diagram and Figure 13.26b shows the tire print when lateral forces due
to cornering and the camber thrust act in the same direction. The resultant of the vertical
tire force moves further to the right and creates an overturning moment Mx = −tc FN ,
where tc is the camber trail. While this overturning moment is not large for cars, it can have
significant values for motorcycles and landing airplanes, as both can have large inclination
angles.

13.9 Other Tire Effects


The ability of a tire to generate a lateral or tractive force is influenced by a large number of
factors. These factors are rarely independent of each other. It usually is difficult to isolate a
factor and to run tests that can pinpoint the effects of changing that factor. The advantages
and disadvantages of these factors must be balanced against each other.
696 Applied Dynamics


a) b)

y y

FC
tc

FN

FIGURE 13.26
a) Rear view of cambered tire, b) tire print for cambered tire (top view).

13.9.1 Pressure
Both underinflated and overinflated tires lead to overheating. Increasing the air pressure
makes the tire stiffer. A stiffer tire has higher cornering stiffness.2 Also, higher pressure in
a tire leads to less drag and lower rolling resistance. The air inside a tire does extra work
when a tire is overinflated, leading to overheating. In addition, because the contact patch is
smaller, the force/unit length on the contact patch becomes higher, resulting in faster wear,
bouncy ride, and reduced controllability. A smaller contact patch generates lower levels of
lateral forces.
Slightly reducing tire pressure increases the size of the contact patch (which is how some
people deal with icy roads). However, in an underinflated tire, the center of the tire print
does not make complete contact with the road. Thus the contact area between the tire and
road is reduced. Also, the sidewall does more work in supporting the wheel loads, resulting
in overheating, additional hysteresis losses, and loss of stability. Severe loss in cornering
stiffness occurs for underinflated tires. In addition, the air in the tire does less of the work
of carrying the wheel load, resulting in larger forces acting on the tire wall.
In a properly inflated tire, 95% of the wheel load is supported by the air and about 5%
by the tire wall. Over or underinflation by less than 5 psi is not easily recognized by the
naked eye, so it is important to check tire pressure periodically. Tire pressure is also affected
by temperature; a tire that is properly inflated during the summer may be underinflated in
the winter and vice versa.
Figure 13.27 depicts under and overinflated tires. Setting tire pressure is an optimization
process, where the above-mentioned factors need to be weighed against each other. The
goal is to set the air pressure such that the center tread and edges of the tread carry a
proportional share of the vehicle load.
A popular method of checking for proper inflation is as follows: Draw a line across the
tire tread with chalk. Then drive in a straight line and examine the chalk remaining on the
tread. If the entire line of chalk has rubbed off, the tire is properly inflated. If the center
of the line is rubbed off but at the ends chalk is still present, the tire is overinflated. If the
line is rubbed off at the ends but chalk is still present in the center of the tread, the tire is
2 Vehicles whose rear wheel loads are very high sometimes specify higher tire pressure on the rear tires.
Vehicle Dynamics—Tire and Aerodynamic Forces 697

Proper inflation Over inflation Under inflation

FIGURE 13.27
Tire print of under and overinflated tires.

underinflated. If chalk on only one side of the tire has rubbed off, there is an imbalance in
the tire or the vehicle.

13.9.2 Temperature
When discussing temperature effects in a tire, we need to specify to which temperature we
are referring:

• The temperature of the air inside the tire, or


• The temperature of the outside walls of the tire, including sidewall and the tread, or
• The temperature of the inside walls of the tire.

The temperature of a tire affects the material properties of the tire. For example, in-
creased temperature lowers the modulus of elasticity, causing the tire to deform more. On
the other hand, higher tire temperatures increase the coefficient of friction between the tire
and road. Race car drivers will “spin” their tires, known as burnout, before a race to increase
the tire temperature and thus create larger amounts of friction. Change in tire temperature
also affects the rolling resistance, as discussed earlier. An increase in the temperature of the
air inside the tire will lead to an increase in pressure.

13.9.3 Speed
In general, the force generating capability of a tire decreases at higher speeds. Tires get
hotter when driven over high speeds and when large lateral forces are involved. It is difficult
to separate speed-temperature effects, as it is nearly impossible to increase tire speed and
keep temperature unchanged. In general, higher speed results in lower cornering stiffness,
which reduces stability and causes more rapid wear of the tire.

13.9.4 Conicity and Ply Steer


The manufacture of tires involves several steps and components, with plies placed over a
carcass on top of which the treads are attached. Any inaccuracies in the manufacturing
process, as well as forces encountered as a result of impact (severe bumps, hitting a side-
walk, etc.), alter the intended geometry and symmetry of tires. Also, one or more of the
components of a tire may wear out or dislocate with use. It is important to balance tires
periodically.
Conicity is due to having an off-center belt. As a result, the tire looks like a truncated
cone and the vertical and lateral force distribution in the contact patch changes. The tractive
698 Applied Dynamics

force may also no longer be symmetric. The effects of conicity include generating a pull in
the steering system, more rapid wear and tear, and reduction in stability.
Ply steer results in unevenness of the plies in the tire. This can be due to a manufacturing
defect or something that happens over time, when the bonds between the carcass and plies
begin to lose their integrity. Ply steer is a more significant problem than conicity.

13.10 Summary of Tire Force Effects

v
 Mx FT – FRR

tp  FT
x
= My
x

tr

FL+FC FL+FC
FN
y FN Mz
z y
z

FIGURE 13.28
Summary of tire effects.

Figure 13.28 shows the tire forces acting on the contact patch and their resultants. The
tractive (acceleration or braking) forces and the rolling resistance act in the x direction.
The lateral force and camber thrust act in the y direction. Acting in the vertical direction
is the wheel load. As a result of the uneven distribution of the tire forces along the contact
patch, there also are tire moments: aligning torque Mz , overturning moment Mx , and rolling
resistance moment My .
For a defective tire or a tire that is worn out on one side more than the other, the
resultant of the tractive force may not go through the x axis and the resulting moment can
create instabilities.
Note that, because a tire is cambered or may be leaning due to a turn, the xyz coordi-
nates considered here are not attached to the wheel. This property is shown in in Figure
13.29. The wheel plane is the xz 0 plane, where the xy 0 z 0 axes are obtained by rotating the
xyz axes about the x axis by the inclination angle γ. The spin axis of the wheel is the y 0 axis.
Consequently, the rolling resistance moment and overturning moment will have components
along the spin axis in the form

My0 = My cos γ + Mz sin γ Mz0 = Mz cos γ − My sin γ (13.22)


Vehicle Dynamics—Tire and Aerodynamic Forces 699

Wheel plane
x


y'
My
y

 y'
Mz
z' z

FIGURE 13.29
Rear view of wheel plane.

13.11 Nondimensional Analysis of Tire Behavior


Earlier in this chapter, we studied changes in the lateral tire FL as a function of the slip
angle α and for a variety of factors. We also saw that by plotting the lateral force coefficient
cL vs. the slip angle α, the curves for different values of the vertical load FN became much
closer to each other. This suggests looking for a more convenient way to characterize tire
behavior. The discussion in this section follows the developments in Milliken and Milliken.
As discussed in Chapter 1, nondimensionalization is a useful tool for analysis. By de-
veloping dimensionless variables and coefficients, we can simplify the describing equations,
build scale models and, if experimentation is required, conduct a smaller set of experiments.
Consider development of a nondimensional lateral force. In Section 13.4, the lateral load
coefficient was defined by cL = FFNL as a nondimensional ratio between the lateral force and
vertical force. This ratio resembles a friction coefficient. The nondimensional lateral force,
denoted by F̄L , with the bar denoting that the quantity is nondimensional, is defined as
FL
F̄L = (13.23)
µy FN

where µy is the coefficient of friction. Because of tire tread geometry and other factors,
there will be different coefficients of friction in the lateral and longitudinal directions. Also,
the coefficient of friction changes with tire pressure and with the size of the contact patch.
The normalized slip angle ᾱ is defined as
Cα tan α
ᾱ = (13.24)
µy FN

in which Cα is the cornering stiffness.


Nondimensionalizing experimental data according to the ratios above gives a lateral
force vs. slip angle curve in the form shown in Figure 13.30a. It is interesting to note that
the vast majority of the data points in tire experiments fall almost exactly on this curve.
The slope of the linear portion of the curve is the normalized cornering stiffness C̄ = F̄L /ᾱ
700 Applied Dynamics

of the tire. A variety of curve-fitting techniques can be used to quantify the curve, including
the magic formula in Equation (13.12).

a) FL b) Mz
1 0.4

0.2
C
0 0

-0.2

-1 -0.4
 
-5 0 5 -5 0 5

FIGURE 13.30
Normalized tire plots: a) normalized lateral force vs. normalized slip angle, b) normalized
aligning torque vs. normalized slip angle.

The aligning torque can be normalized as


Mz
M̄z = (13.25)
tp µy FN
where tp is the pneumatic trail. A plot of M̄z vs. ᾱ has the form shown in Figure 13.30b.
Again, experimental data for a variety of vertical loads fit this curve quite closely.
Normalized counterparts of the overturning moment and camber angle can be obtained
in a similar way:
Mx Cγ sin γ
M̄x = γ̄ = (13.26)
to µy FN µy FN
in which to is the overturning trail, Cγ is the camber stiffness (ratio of camber thrust to
camber angle). Plots of the above quantities follow the same pattern as the other plots of
normalized tire behavior.
Normalization can be extended to the longitudinal motion where the tractive force and
slip ratio are normalized by
Fx kx SR
F̄x = S̄ = (13.27)
µx FN µx FN
in which µx is the coefficient of friction in the longitudinal direction, SR is the slip ratio,
and kx is the initial slope of longitudinal force with slip ratio, plotted in Figure 13.19.

13.12 Aerodynamic Forces


Chapter 4 discusses the basics of aerodynamics. When air flows over a body, two forces are
generated: a pressure distribution about the body that leads to lift and drag forces, and
Vehicle Dynamics—Tire and Aerodynamic Forces 701

a shear force distribution, due to sliding friction between the air flow and the body, that
results in drag. In Chapter 4, we also saw that as air flows over a body, the air flow may
not follow the body and the flow may separate from the body, forming vortices in the wake
and generating additional drag forces.
The magnitudes of the lift and drag forces were quantified in Chapter 4 in terms of lift
and drag coefficients, which are dimensionless quantities. In the early years of aerodynam-
ics, such coefficients were measured only experimentally. With advances in computational
methods, more complex bodies can be analyzed numerically. This is true not only of aircraft,
but also of land vehicles.
The lift, drag, and side forces act at the center of pressure, a point whose location
changes as a result of a variety of factors. It is customary to express the net effect of the
aerodynamic forces as a set of forces applied through a point and a set of moments.3



 



 





 




 

 



FIGURE 13.31
Vehicle aerodynamic system: a) side view, b) top view.

Shown in Figure 13.31, the aerodynamic coordinate system used for vehicles is the same
as vehicle coordinates. The reference point, or origin O, is selected at the midpoint of the
wheelbase. The magnitudes of the forces and moments are given by
1 1 1
FL = ρAv 2 CL FD = ρAv 2 CD FS = ρAv 2 CS (13.28)
2 2 2
3 Similar to what we do with airfoils, as described in Section 4.10.
702 Applied Dynamics
1 1 1
MP M = ρAv 2 LCP M MRM = ρAv 2 LCRM MY M = ρAv 2 LCY M (13.29)
2 2 2
where ρ is the density of air.
Each aerodynamic effect is characterized by its coefficient. The subscripts L, D, and S
stand for lift, drag, and side force, respectively.4 The moments are the pitch moment, roll
moment, and yaw moment, where the notation is obvious. A is a reference area, usually
taken as the planform area of the front of the vehicle, and v is the relative vehicle speed
(vehicle speed minus wind speed). The magnitude of v also depends on the wind direction.
The term 12 ρv 2 is the dynamic pressure. In the moment expressions above, the parameter
L is usually taken as the wheelbase.
Consider a relative wind speed V∞ and a side wind angle β, as shown in Figure 13.32.
The side force, yaw moment, and roll moment coefficients CS , CY M , and CRM are affected
by the wind angle β substantially. So is the drag coefficient CD . The lift coefficient CL is
not affected as much by the wind angle and neither is the pitching moment coefficient CP M .
In the absence of wind blowing from the side, the aerodynamic effects are lift, drag, and
pitching moment. The other three aerodynamic effects—side force, roll moment, and yaw
moment—become important if there is a persistent strong wind or when there is a wind
gust. Crosswinds primarily affect yawing.

L/2 L/2

V

v x
FD

FS

FIGURE 13.32
Wind velocity V∞ and angle β.

The most significant aerodynamic effect acting on land vehicles is drag. Drag is always
present, it opposes the velocity, and it wastes energy. Drag becomes more significant at
higher speeds. Vehicle designers do their best to minimize drag, especially for race cars.
The main sources of drag in a vehicle are

• Vehicle body. Here the fore body (front of the vehicle), after body (where the flow
separates and forms vortices), underbody, and skin friction contribute to drag. Of these,
the after body has the largest contribution to the drag force. The rear of a vehicle that
is designed to be fuel efficient or to go fast resembles an airfoil more than the front.
• Protuberances. Protuberances consist of the wheels, wheel wells, drip tails, window
4 Please do not confuse the lateral force with the lift force.
Vehicle Dynamics—Tire and Aerodynamic Forces 703

recesses, side view mirrors, air intake for cooling and combustion, door handles, and
antennas. Of these, the wheels contribute to the drag force the most. Driving with an
open window also contributes to drag. At high speeds, it is more efficient to keep the
windows closed and the air conditioning on than to keep the windows open and the air
conditioning off.
For a passenger car, the cumulative effect of drag result is a drag coefficient of CD ≈ 0.3.
Sports cars and race cars have lower drag coefficients, around CD ≈ 0.2 to 0.25. Trucks have
drag coefficients around CD ≈ 0.7 to 0.9.
While lift force maximization is desirable for aircraft, lift is an undesirable force in
land vehicles. Lift forces lower the wheel loads and hence reduce the magnitudes of the
acceleration and braking forces that can be applied. Lift makes it possible for the vehicle
to lose contact with the road surface and thus reduces stability. Some sporty vehicles have
aerodynamic aids such as spoilers (inverted airfoil) to reduce lift. As in any design that
involves lift and drag, a geometry that changes the lift characteristics will also change the
drag force and in every design there will be tradeoffs.
The pitching moment coefficient primarily results from the choice of the origin of the
coordinate system to locate the aerodynamic forces. Consider the actual location of the
resultant drag force. Such a force would be near the geometric center of the planform area
(the front of the vehicle). Moving the resultant drag force along the z axis, as shown in
Figure 13.33, creates the aerodynamic moment MP M about the y axis. The lift force also
contributes to the pitching moment, but to a lesser extent than the drag force. For most
vehicles, the pitching moment coefficient is around CP M ≈ 0.05 to 0.2.

G v
( FD )
actual

O FD
x
L/2 L/2
MPM
FL
y
z

FIGURE 13.33
Lift and drag forces and pitching moment.

Example 13.3
A vehicle accelerates from rest to a speed of 60 mph in 12 seconds. Assuming acceleration is
constant, calculate the power needed to overcome the drag force and the total energy used
to fight drag.
The drag force on the vehicle is FD = 12 ρACD v 2 . Writing the velocity as v = at, where
a is the acceleration, the power needed to counter drag becomes
1 1
P = FD v = ρACD v 3 = ρACD a3 t3 [a]
2 2
704 Applied Dynamics

Denoting the time to reach the desired speed by T , the energy use is obtained by
Z T Z T
1 1
E = P dt = ρACD a3 t3 dt = ρACD a3 T 4 [b]
0 0 2 8

Next, let us assign some numbers to calculate the power and energy. The density of air
at a temperature of 59◦ F is ρ = 0.00238 slug/ft3 . Consider a planform area of A = 22 ft2
and drag coefficient CD = 0.3. The final speed of the vehicle is 60 mph = 88 ft/sec. The
average acceleration of the vehicle is
v 88 2
a = = = 7.333 ft/sec [c]
T 12
The drag force at the end of the run is

FD (t = 12) = 0.5 × 0.00238 × 0.3 × 20 × 882 = 55.29 lb [d]

The power needed to resist drag at the end of the run is

P = 0.5 × 0.00238 × 20 × 0.3 × 883 = 4865 lb · ft/sec = 8.845 hp [e]

From Equation [b], the energy used to counter drag as the vehicle accelerates becomes

E = 0.125 × 0.00238 × 20 × 0.3 × 7.3333 × 124 = 14, 595 ft · lb = 18.756 Btu [f ]

It is interesting to note that the same vehicle traveling at twice the speed would use up
eight times the power, so for a speed of 120 mph, 8.845 × 8 = 70.76 hp would be required,
which would be slightly less than half the maximum power of such a vehicle. Assuming a
roll resistance coefficient of 0.014, and a vehicle weight of 4,000 lb, the roll resistance force
becomes FRR = 0.014 × 4, 000 = 56 lb. So, at a speed of 60 mph, the aerodynamic drag and
rolling resistance forces are roughly of the same magnitude.

13.12.1 Calculation of Aerodynamic Coefficients from Test Data


The lift and drag coefficients are often obtained experimentally. The vehicle is placed in a
wind tunnel and the forces that hold the body in place in the wind tunnel are measured
at different wind speeds. The measurement data are used to calculate the aerodynamic
forces that act on the body. The lift and drag coefficients are then obtained from these
aerodynamic forces.
Figure 13.34 shows a vehicle held in place in a wind tunnel. The vehicle is anchored at
the wheels. The three forces Wf , Wr , and FT are measured for different wind speeds and
the aerodynamic forces of lift (FL ), drag (FD ), and moment (MP M ) are calculated. The
sums of forces are
+
X X
→ F = 0 = FT − FD +↑ F = 0 = Wr + Wf + FL − W (13.30)

from which we calculate the lift and drag forces as

FD = FT FL = W − Wf − Wr (13.31)

Summing moments, say, about point O, gives


 
L L L
MO = Wf − Wr − W c− + MP M (13.32)
2 2 2
Vehicle Dynamics—Tire and Aerodynamic Forces 705

c b V
G
h

A O FD B x
FT L/2 L/2

Wr FL Wf
y MPM
z

FIGURE 13.34
Aerodynamic forces acting on vehicle.

from which we obtain the aerodynamic moment as


 
L L
MP M = (Wr − Wf ) +W c− (13.33)
2 2
Let us denote the measured values of the axle loads and tractive force by Wrm , Wfm ,
and FTm , respectively. By manipulating Equations (13.31) and (13.33), we obtain estimates
of the aerodynamic forces and moments as

FDe = FTm FLe = W − Wrm − Wfm

 
L L
MP M e = (Wrm − Wfm ) +W c− (13.34)
2 2
where the superscript e denotes that the quantities are estimated.
The estimated forces and moment are used to estimate the drag, lift, and pitching
moment coefficients by means of the relationships
1 1 1
FDe = ρAv 2 CDe FLe = ρAv 2 CLe MP M e = ρALv 2 CP M e (13.35)
2 2 2
Each set of data obtained from a particular speed gives a set of values for the aerody-
namic forces. By comparing the results for different values of the data, we can estimate the
aerodynamic coefficients.

Example 13.4
Given a vehicle with wheelbase L = 9.5 ft, static weights Wrs = 1500 lb, Wfs = 1800 lb,
and frontal area 22 ft2 , let us find the lift and drag coefficients given the measured data in
Table 13.3.
The values in Table 13.3 are calculated for the case when CD = 0.3, CL = 0.22, and
CP M = 0.1. These “measured” values are obtained by computer simulation. The aerody-
namic forces and moments are calculated for the given speeds and for the lift, drag, and
moment coefficients. Then, we add a random error (uniform distribution) in the range of
−1% to 1% to each force to ascertain the effects of measurement and roundoff errors.
706 Applied Dynamics

TABLE 13.3
Wind tunnel data for tested vehicle

Speed Front Wheel Load Rear Wheel Load Tractive Force


(ft/sec) (lb) (lb) (lb)
0 1800 1500 0
25 1781.5 1497.9 4.8572
50 1801.7 1497.8 19.501
75 1779.0 1492.7 44.103
100 1744.8 1497.6 78.648

The estimated aerodynamic coefficients in the presence of random measurement error


are given in Table 13.4. The results indicate that the front wheel loads change more than
the rear wheel loads.

TABLE 13.4
Estimated aerodynamic coefficients for 1% measurement error

Speed (ft/sec) CDe CLe MP Me


Actual 0.3 0.22 0.1
25 0.2981 1.2618 0.5028
50 0.2992 0.0091 −0.0299
75 0.3008 0.1931 0.0469
100 0.3017 0.2208 0.1014

The drag coefficient is estimated much more accurately, while the lift and moment
coefficients are not. Both the lift and pitching moment coefficients are estimated more
accurately at higher speeds. This makes sense, as at higher speeds the aerodynamic forces
are larger, having a more profound effect on the wheel loads.
To better understand the effects of measurement error, we now calculate the measured
values of the wheel loads and tractive force by assuming 0.05% measurement error. The
results, shown in Table 13.5, clearly show the improvement in the estimation in the presence
of lower measurement error.5 The effects of random measurement errors disappear when the
error percentage goes down to 0.01%.

13.13 Bibliography
Automobile Ride, Handling, and Suspension Design, http://www.rqriley.com/suspensn.htm
Gillespie, T.D., Fundamentals of Vehicle Dynamics, SAE Publications (R114), 1992.
Jazar, R.N., Vehicle Dynamics: Theory and Application, 2nd Edition, Springer, 2013.
Karnopp, D., Vehicle Dynamics, Stability, and Control, 2nd Edition, CRC Press, 2013.
5 If you solve this example yourself, because the errors are generated by a random number generator, you

will get different results than what is listed here every time you run the simulation.
Vehicle Dynamics—Tire and Aerodynamic Forces 707

TABLE 13.5
Estimated aerodynamic coefficients for 0.5% measurement error

Speed (ft/sec) CDe CLe MP Me


Actual 0.3 0.22 0.1
25 0.3006 0.6385 0.0906
50 0.2992 0.1715 0.1450
75 0.2989 0.2672 0.0755
100 0.2994 0.2231 0.1262

Milliken, W.L., and Milliken, D.F., Race Car Vehicle Dynamics, SAE Publications (R146),
1995.
Pacejka, H., Tire and Vehicle Dynamics, 3rd Edition, Butterworth-Heinemann, 2012.
Tire Tech, http://www.tirerack.com/tires/tiretech/tiretech.jsp
Wong, J.Y., Theory of Ground Vehicles, 4th Edition, Wiley, 2008.

13.14 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 13.4—Lateral Forces and Tire Slip

TABLE 13.6
Lateral force vs. slip angle

Slip Angle α 1◦ 2◦ 3◦ 4◦ 5◦ 6◦ 7◦
Lateral Force (lb) 500 1000 1500 1950 2400 2850 3100

13.1 (M) The lateral force vs. slip angle values for a tire are given in Table 13.6. Expressing
the lateral force as a function of the slip angle as FL = c1 α + c2 α2 , estimate the coefficients
c1 and c2 using a least squares approximation. Make sure that at zero slip angle the lateral
force is also zero.

TABLE 13.7
Lateral force vs. slip angle

Slip Angle α 1◦ 2◦ 3◦ 4◦ 5◦ 6◦ 7◦
Lateral Force (N) 2,000 4,000 6,000 7,950 9,600 11,000 10,400

13.2 (M) The lateral force vs. slip angle values for a tire are given in Table 13.7. Expressing
708 Applied Dynamics

the lateral force as FL = c1 α + c2 α2 + c3 α3 , estimate the coefficients c1 , c2 , c3 using a least


squares approximation. Make sure that at zero slip angle the lateral force is also zero.

TABLE 13.8
Lateral force and vertical force at slip angle α = 3◦

Lateral Force FL (N) 1,200 1,800 2,200 2,500


Vertical Force FN (N) 1,800 2,700 3,600 4,500

13.3 (M) The lateral force vs. vertical force values are given in Table 13.8. Expressing
the cornering stiffness in terms of the vertical force as in Equation (13.6), estimate the
coefficients e1 and e2 using a least squares approximation.

Section 13.7—Rolling Resistance


13.4 (M) A 1700 kg rear wheel drive vehicle with a wheelbase L = 2.70 m, and center of
mass at a distance 1.5 m from the rear and 1 m high is going uphill at a grade of 4◦ . The
rolling resistance coefficient is fr = 0.015 and aerodynamic forces are negligible. For this
vehicle to accelerate at 0.2g, determine the tractive force that is necessary and the resulting
axle loads. Also calculate the minimum coefficient of friction that is needed to sustain this
acceleration.
13.5 (M). A vehicle of weight 4000 lb, with wheel loads distributed equally, is driven with
the front wheels experiencing a slip angle of 2◦ . The rear wheels are not experiencing any
slip. Given that the cornering stiffness on the front axle is Cf = 250 lb/◦ and the rolling
resistance coefficient is fr = 0.013, calculate the total drag on the vehicle due to the slip
and rolling resistance (ignore aerodynamics). Also, calculate the power needed to maintain
a speed of 50 mph.

Section 13.12—Aerodynamic Forces


13.6 (E) Consider a vehicle of mass 1600 kg, rolling resistance coefficient of fr = 0.02,
drag coefficient of CD = 0.3, and cross-sectional area of 1.7 m2 . At what speed will the
aerodynamic drag force be the same as the rolling resistance force?
13.7 (E) A vehicle of weight 3500 lb, frontal area 22 ft2 , and drag coefficient CD = 0.33 is
affected by rolling resistance which is modeled by Equation (13.18). At what speed will the
rolling resistance force be the same as the drag force? Solve this problem for a) sea level,
temp = 59◦ F, b) altitude 5,000 ft, temp = 40◦ F. Hint: You need to consult a handbook on
values of density.
13.8 (E) A 1200 kg vehicle (wheelbase 1.5 m, CG height 70 cm, frontal area 1.8 fm2 , mf
= 650 kg under static conditions) is traveling at a speed of 70 kph and is accelerating at
the rate of a = 0.1g. The values for the drag, lift, and pitching moment coefficients are
CD = 0.25, CL = 0.15, CP M = 0.05. a) Calculate the axle loads and required tractive force
to maintain acceleration when there is a headwind (towards the car) of 25 kph and rolling
resistance coefficient of fr = 0.12. b) How much power is the vehicle consuming (in hp)?
13.9 (E) Consider Problem 13.5 and calculate the total drag force if the vehicle has a frontal
area of 23 sq. ft. and drag coefficient of CD = 0.33.
13.10 (M) A rear wheel drive vehicle has a mass of 1500 kg, wheelbase of 2.6 m, and static
axle loads of 800 kg rear and 700 kg front. When traveling at a speed of 60 km/h, the axle
loads due to the weight and aerodynamic forces are 790 kg rear and 700 kg front. What
Vehicle Dynamics—Tire and Aerodynamic Forces 709

are the wheel loads when the vehicle travels at a speed of 75 km/h? Hint: Calculate the lift
force and pitching moment and express in terms of the speed (FL = c1 v 2 , MP M = c2 v 2 ).
13.11 (E) A front wheel drive vehicle is subjected to wind tunnel testing. The wind speed
is along the length of the vehicle. The following results are obtained and tabulated in
Table 13.9. Calculate the horizontal location of the center of mass and the lift, drag, and

TABLE 13.9
Wind tunnel test results

Velocity (ft/sec) Wf (lb) Wr (lb) FT (lb)


0 2300 1700 0
25 2295 1703 11

pitching moment coefficients. Then, calculate the wheel loads and required tractive force
when the wind speed becomes 40 ft/sec. The vehicle has a wheelbase of 9 ft, center of mass
height (not needed for this problem but will be used later on) of 2 ft. The frontal planform
area of the vehicle is 21 ft2 .

TABLE 13.10
Wind tunnel test results

Velocity (ft/sec) Wo (lb) Wi (lb) Ff (lb)


0 2000 2000 0
25 2005 1998 10

13.12 (M) The vehicle in the previous problem is tested in a wind tunnel, but this time
for a side force, so the wind blows from the side. The following loads are measured and
tabulated in Table 13.10. Here, Ff denotes the total friction force that the wheels apply in
the sideways direction, and o and i denote the outside and inside wheels. Assuming that
the only aerodynamic loads acting on the vehicle are the side force, lift, and roll moment,
and the vehicle has a track of t = 5 ft, calculate the aerodynamic loads and associated
aerodynamic coefficients.
13.13 (M) A 4000 lb vehicle (wheelbase 9.5 ft, Wf = 2200 lb under static conditions), is
placed inside a wind tunnel. The wheel loads at a speed of 30 mph are Wr = 1798.1 lb and
Wf = 2191.8 lb. A second set of measurements are taken at a different speed, with values of
Wr = 1795.7 lb. and Wf = 2181.7 lb. At what speed were the second set of measurements
taken? Calculate this speed using both the vertical force balance and the moment balance
and comment on the difference.
14
Vehicle Dynamics—Lateral Stability

14.1 Introduction
This chapter considers the lateral motion of vehicles during cornering or while other side
forces, such as wind and gravity due to banked roads, act on the vehicle. We first look
at a representative description of steering angles, and study oversteer and understeer. The
transient motion equations are derived and the effects of different types of forcing are
analyzed. Steady-state motion is investigated. The chapter introduces important ratios that
quantify vehicle characteristics. A model of the driver is developed.
When studying lateral stability and deriving stability equations, it is important to be
aware of the units as well as positive directions of the variables that are considered, especially
slip angles and steering angles. Also, lateral stability cannot be separated from roll stability
of a vehicle, as rollover accidents usually are related to lateral stability. The effect of lateral
forces on roll will be discussed in the next chapter.

14.2 Kinematics—Steer Angle Definitions


We begin with the bicycle model of a vehicle in Figure 14.1 (using the SAE coordinate
system) and consider location of the turn center (instant center) of the vehicle in the absence
of tire slip. Ignoring the slip angle is a reasonable assumption when the speeds involved are
low. The steer angle provided by the driver is denoted by δ.
To locate the instant center, because the directions of the velocities at A and B are
known, we draw perpendiculars to the velocity vectors at points A and B. For the case of
no tire slip, the velocities at points A and B are in the x and x0 directions, respectively.
Let us denote the location of the instant center for the case of no slip as IN S . The turn
radius is defined as the distance between the instant center and the center of mass of the
vehicle. The turn radius for the case of no slip is denoted by RN S . Considering that the
lines IN S A, IN S B, and RN S are very close in magnitude, a small angles approximation can
be used, which leads to
L L
= tan δ ≈ δ or RN S ≈ (14.1)
RN S δ
where L = b + c is the wheelbase. All angles in the derivation above are expressed in terms
of radians. The tractive and lateral forces developed by the tires are denoted by FT and
FL , respectively.
Next, consider slip in the front and rear tires and denote the slip angles by αf and αr .
Because the vehicle is making a right turn (clockwise when viewed from top, or towards the
right and in the positive y direction), the slip angles are both counterclockwise (away from
the turn).

711
712 Applied Dynamics
F Tf

vA L
F Tr A B
r G x
F Lf
 
F Lr c b vB y
f

RNS
RS y' x'

P


INS S

IS

FIGURE 14.1
Turn centers and slip angles.

In the coordinate system here, a positive rotation about the z axis is clockwise when
viewed from the top so a positive slip angle should be clockwise. The slip angles are, however,
away from the turn, so the actual slip angles are, according to our sign convention, negative.
The slip angles are drawn in their actual direction in Figure 14.1, and we use their absolute
values to avoid ambiguity (that may arise from not marking the slip angles in the positive
directions associated with the sign convention).
The velocities of the tires are along the direction of slip. To calculate the location of
the instant center, also known as turn center, we draw perpendiculars to the velocities of
the front and rear wheels. The intersection of these lines, which defines the location of
the instant center in the presence of slip, is denoted by IS . The turn radius or radius of
curvature, that is, the distance between the instant center IS and center of mass G in the
presence of slip, is represented by RS .
The angle between the two perpendicular (to the travel directions of the front and rear
wheels) lines is, in essence, the resulting steering angle of the vehicle. This angle is denoted
by δS and referred to as the net steer effect. It also is referred to as the Ackermann steer
angle, not to be confused with Ackermann steering.
The angle δS is the resultant of the cumulative effects of the steer angle provided by the
driver and the slip of both tires.1 For small steer angles, the radius of curvature with slip,
RS , can be related to the net steering effect by
L 180 L L
δS ≈ in radians δS ≈ = 57.296 in degrees (14.2)
RS π RS RS
1 Caution: Many texts use different terminology for the driver’s input and net steer effect.
Vehicle Dynamics—Lateral Stability 713

It is of interest examine the relationship between the net steering effect, δS , and δ, the
steering angle provided by the driver. Consider the intersection of the lines IN S B and IS A.
Denote this point by P and note that P is the instant center that we would obtain in the
presence of no slip in the front tire and slip in the rear tire. Clearly, the angle between the
lines IN S B and IS A is larger than δ. The conclusion is that slip in the rear tire increases the
net steer effect. Similarly, slip in the front tire decreases the net steer effect. Using geometry
and considering small angles, we can relate the two steer angles by
L
δS = δ − |αf | + |αr | = − |αf | + |αr | (14.3)
RN S
The above equation uses radians to express the angles involved. In many cases, and
especially with cornering stiffness plots, it is more common to use degrees as units. In
this case, all terms in the above equation are converted to degrees. Multiplying L/RN S by
180/π = 57.296, the net steering effect becomes

L
δS = δ − |αf | + |αr | = 57.296 − |αf | + |αr | (14.4)
RN S
where all angles are now in degrees.
There are three possibilities for the slip angle magnitudes, depending on the value of
|αf | − |αr |. Consider first the case where the magnitude of the front slip angle is larger than
the rear slip, |αf | > |αr |. It follows that δS < δ and RS > RN S . Shown in Figure 14.2a, this
case is known as understeer, as the net steer effect is less than the steer angle provided by
the driver.
The vehicle understeers the driver’s effort and it tracks a larger circle. The driver feels
the vehicle moving out, rather than following the same path of the front tires, as depicted
in Figure 14.2b. To obtain the desired amount of turn, the driver has to provide a larger
steer effort and increase the steer angle by |αf | − |αr |. Most passenger vehicles are designed
with a small amount of understeer, for reasons that will soon be clear.
Another way of visualizing understeer is to note that the outward (away from the turn)
motion is higher at the front wheel, which has the effect of turning the rear wheels inward
(towards the turn), making the vehicle turn less than the intent of the driver. Alternatively,
since the magnitude of the front slip angle is larger than the rear slip angle, the front wheels
slip (away from the turn) more than the rear wheels and the driver needs to increase the
steer angle, as can be seen from Figure 14.2b.
In the second case, |αf | < |αr |, resulting in δS > δ, as shown in Figure 14.3a. It follows
that RS < RN S . This case is known as oversteer, as the net steer effect is larger than the
steer angle provided by the driver. The vehicle oversteers the driver’s effort and turns more,
and it tracks a smaller circle. To have the desired amount of turn, the driver has to reduce
the steer effort (reduce δ). Another way of visualizing oversteer is to note that the outward
(away from the turn) motion is higher at the rear wheel, which has the effect of turning the
front wheels inward (towards the turn), thus making the vehicle turn more than the intent
of the driver, as can be seen from Figure 14.3b.
Oversteer is more dangerous than understeer, especially at high speeds. Consider taking
a turn. The driver turns the steering wheel by an amount the driver expects will be adequate.
If the vehicle understeers, the driver will have to turn the steering wheel more in the direction
of the turn. For an oversteer vehicle, the driver has to reduce the steer angle so as to move
the vehicle away from the turn, which is harder to do. As will be shown later, an oversteer
vehicle becomes uncontrollable when the vehicle speed exceeds a critical value.
For both oversteer or understeer, we can think of the steering angle provided by the
driver δ as the desired steer angle and the δS as the needed steer angle (needed to take
714 Applied Dynamics

L
A vA
a) r G B b)

 f vB Actual
Intended

INS

S

IS

FIGURE 14.2
a) Turn center IS for an understeer vehicle. b) Path followed by vehicle.

L
vA
a) A r B b)
G

f v Intended
B
Actual

 s

INS IS

FIGURE 14.3
a) Turn center IS for an oversteer vehicle. b) Path followed by vehicle.
Vehicle Dynamics—Lateral Stability 715

the turn properly), with the driver providing (hopefully) the needed steer angle in order to
obtain the desired steer effect. Using this terminology, we can write

δneeded = δdesired + |αf | − |αr | (14.5)

This is the notation that some texts use. It is important to distinguish which value of δ is
being considered.
A third but less frequently encountered case is that of neutral steer, where the slip angles
for the front and rear tires are the same, αf = αr . In this case, the net steering effect is the
same as the driver’s steering effort and the vehicle moves as if there is no slip in either tire.
Table 14.1 summarizes the different steer effects.

TABLE 14.1
Steer effects based on slip angles

Type Slip Angles Net Steer Effect δS Turn Radius RS


Understeer |αf | > |αr | δS < δ RS > RN S
Oversteer |αf | < |αr | δS > δ RS < RN S
Neutral Steer |αf | = |αr | δS = δ R S = RN S
(or No Slip)

The smallness of the steer angle leads to an approximate expression for the lateral
acceleration aGy = ay as
2
vG
ay ≈ (14.6)
RN S
for the case of no slip (neutral steer) and
2
vG
ay ≈ (14.7)
RS
for the case of slip.
Section 14.14 presents another situation when the terms oversteer and understeer are
used. If the driver changes the steering angle rapidly, especially at high speeds, the front
tires may start to slide, resulting in understeer behavior. If a driver enters a turn with high
speed, the rear tires may start sliding, which leads to oversteer behavior.

Example 14.1
This example (which does not involve wheel slip) investigates the accuracy of the assump-
tions regarding lateral acceleration. Denote the distance from A to IN S in Figure 14.1 by
h. The steer angle is δ = 0.02 = 1/50 rad. This corresponds to a steer angle of about 1.2◦ ,
which is large for vehicles driven at high speeds. From the small angle approximation
L
tan δ = tan 0.02 = 0.0200027 = [a]
h
so that h = 49.9933L.
Next, consider the right triangle A-IN S -G, in Figure 14.4 and write
2 2 2
RN S = h +c [b]
716 Applied Dynamics

A L B 
G x
vA c b 
vB y

RNS
h


0

INS

FIGURE 14.4
Turn radius for no slip.

Dividing the above equation by L2 gives


 2  2  
RN S h c 2
= + [c]
L L L
RN S
For a vehicle with c/L = 0.6 (center of mass closer to the front axle), L becomes
RN S p
= 49.99332 + 0.62 = 49.99693 [d]
L
Comparing this value with the approximate value RN S /L = 1/δ = 50, the approximation
in the previous section is quite accurate, to less than 0.01%.
The small angle approximation for δ will be used from here on. The normal acceleration
v2
of the vehicle is aGn = RNGS and the component of the normal acceleration along the y axis
is
aGy = aGn cos δ0 [e]
in which
c L
tan δ0 ≈ δ0 = = 0.6 = 0.6δ = 0.012 [f ]
h h
and cos 0.012 = 0.99993. Denoting the velocity of point A by vA , the angular velocity of the
vehicle becomes ω = vA /h and the velocity of the center of mass becomes
vA c
vG = vA + ω × rG/A = vA i + k × ci = vA i + vA j [g]
h h
so that   c 2 
2 2 2
vG = vA 1+ = 1.000144vA [h]
h
Introducing these values into Equation [e] gives the lateral acceleration as
2 2
vG 1.00144vA v2
aGy = cos δ0 = × 0.99993 = 0.020029 A [i]
RN S 49.99693L L
2 2
which is almost identical to the approximate value aGy = vG /R = vG δ/L.
Vehicle Dynamics—Lateral Stability 717

14.3 Wheel Loads and Slip Angles


This section investigates the relationship between slip angles and wheel loads. The vehicle
is modeled as moving with constant speed, so that the tractive force is there to overcome
rolling resistance, aerodynamic drag, and induced drag.
We also consider steady-state motion of the vehicle. At steady-state, the lateral velocity
of the center of mass v and the angular velocity of the vehicle ω are constant. The only
nonzero acceleration is aGy , which is due to the turn.
We consider two free-body diagrams: side view of the vehicle, showing the axle loads,
and top view, showing the lateral forces. Note that we draw the free-body diagrams with
the slip angles in their positive directions (and not in the directions they are expected to
be).

14.3.1 Free-Body Diagram and Slip Angles


The free-body diagram of the vehicle, viewed from the side and including the inertia force
maGx in the negative x direction and ignoring aerodynamic effects, is given in Figure 14.5.

W=mg x
v maGx
G
c b z
A h FRR
Wr F Tr B F Tf
Wf
L

FIGURE 14.5
Free-body diagram, side view.

The tractive forces are FTf and FTr . Summing forces along the tractive direction gives
+
X
→ F = FTr + FTf − FRR = maGx = 0 (14.8)

where FRR is the rolling resistance force. Recalling that there is no pitch angular velocity,
and noting from the above equation that the acceleration in the tractive direction is zero,
we sum moments about point A, which gives
X
MA = Wf L − W c + maGx h = Wf L − W c = 0 (14.9)

The sum of forces in the vertical direction gives Wr + Wf = W . Using this relation with
the above equation and solving for the axle loads give Wf = W Lc , Wr = W Lb , which are
the same as the static axle loads.
Next, consider the top view of the vehicle, and draw the free-body diagram, which is
shown in Figure 14.6. The free-body diagram includes the inertia forces maGx and maGy
and inertia moment IG ω̇ in the negative x, y, and z directions, respectively, as two forces
718 Applied Dynamics

maGy
F Tf
A c B FRR
F Tr r maGx x
G  
vA f
IG  F Lf x'
F Lr vB y
vG
y'
b
RS

S

IS

FIGURE 14.6
Top view of free-body diagram.

and one moment acting on an equivalent static vehicle. Recall that for steady-state motion
aGx = 0, ω̇ = 0, and aGy is only due to the normal acceleration.
The slip angles are shown in their positive directions according to our sign convention
(when viewed from top, a clockwise rotation is positive) and not in their actual directions.
The mathematical values of the slip angles should always be negative, so |αr | = −αr and
|αf | = −αf . Looking at Figure 14.6 and evaluating the turn center locations, the net steer
effect is

δS = δ − |αf | + |αr | = δ + αf − αr (14.10)

Consider next the moment balance about point A in the z direction, which gives
X
 MA = 0 = −IG ω̇ − cmaGy + LFLf cos δ + LFTf sin δ (14.11)

Without loss of generality, consider a rear wheel vehicle and set FTf = 0 in the above
equation. Using a small angles assumption for sin δ ≈ δ and cos δ ≈ 1 leads to an expression
for the lateral force on the front wheel as
c IG
FLf = maGy + ω̇ (14.12)
L L
Summing moments about the front axle B gives the expression for the lateral force on
the rear wheel as
b IG
FLr = maGy − ω̇ (14.13)
L L
The next step is to relate the lateral forces to the axle loads. Writing the axle loads as
c c b b
Wf = W = mg Wr = W = mg (14.14)
L L L L
Vehicle Dynamics—Lateral Stability 719

and noting that the angular acceleration ω̇ = 0 for steady-state, one can express the lateral
forces in Equations (14.12)–(14.13) in terms of the axle loads as
aGy aGy
FLf = Wf FLr = Wr (14.15)
g g
Because the lateral forces are related to the slip angle, it is now possible to relate the
slip angles to the axle loads. Consider the linear range of the lateral force vs. slip angle plot
discussed in the previous chapter. Denoting the cornering stiffnesses2 for the front and rear
axles by Cf and Cr , respectively, the lateral forces can be expressed as

FLf = Cf αf FLr = Cr αr (14.16)

To express the relationships above so that the lateral forces are in their correct directions,
a sign convention has to be developed for the cornering stiffnesses. In the coordinate system
here, the lateral forces are in the y direction for a right hand (rotation about the z axis)
turn. The sign convention for slip angle is positive clockwise (about the z axis). However,
tire slip is away from the turn and, for a right turn the slip angle is counterclockwise, hence
negative. To have a positive value (in the y direction) for the lateral forces FLf and FLr ,
the cornering stiffnesses need to be defined as negative quantities.
Note that for the bicycle model, the cornering stiffnesses Cf and Cr are the cornering
stiffnesses of the front and rear axles, so they are the total of the cornering stiffnesses of the
two tires on the axles. Hence, if the cornering stiffness of each tire, say, on the rear axle is
C, where C is a positive number, then the cornering stiffness of the rear axle is Cr = −2C.
Introducing Equation (14.15) into Equation (14.16), we can express the slip angles in
terms of the axle loads Wf and Wr and lateral acceleration aGy . Recall again that we are
considering steady-state motion, where the vehicle speed and yaw angular velocity are both
constant, so ω̇ = 0. The expressions for the slip angles then become
Wf aGy Wr aGy
αf ≈ αr ≈ (14.17)
Cf g Cr g

The magnitude of the slip angle is due to two factors: the axle loads and the lateral
acceleration of the vehicle. Hence, the primary factor that dictates the magnitudes of the
slip angles of the rear and front axles is the vehicle load distribution. Assuming that the
cornering stiffnesses for the front and rear tires are equal, Cf = Cr , when the center of
mass of a vehicle is towards the front, c > b, the axle loads are related by Wf > Wr so that
|αf | > |αr |. This corresponds to the understeer case. Most passenger vehicles are designed as
understeer vehicles, and the front rear weight distribution, Wf /Wr , is somewhere between
60/40 and 54/46.
As discussed in the previous chapter, the lateral force that can be generated increases
as the wheel load is increased. However, the amount of increase is less than the increase in
the wheel load. For example, if the cornering stiffness for a given a wheel load W is C, the
cornering stiffness for a wheel load 2W is less than 2C, as illustrated in Figure 14.7. When
the wheel loads are close to each other, such as a load distribution of 55/45 with identical
tires in the front and rear, the cornering stiffnesses can be assumed to be equal.
For a more accurate model, the full relationship between cornering stiffness and wheel
load, Cf = Cf (Wf ), needs to be used. One approximation of the cornering stiffness to wheel
n
load is as follows: if Wf /Wr = 1 + x, then Cf /Cr ≈ 1 + (x/a) , where a and n are correction
2 The term cornering stiffness is somewhat misleading because the slip angles depend on velocities and

hence the lateral forces also oppose velocities. Lateral forces due to slip have the effect of damping forces
and not spring forces.
720 Applied Dynamics

factors, whose values depend on the tire load sensitivity. For example, for a = 4 and n = 1,
if Wf /Wr = 1.2, then Cf /Cr ≈ 1.05. Using this approximation, the axle load to cornering
stiffness ratios become
Wf 1.2Wr Wr
= ≈ 1.14 (14.18)
Cf 1.05Cr Cr

C
2W
1.5W

FIGURE 14.7
Linear part of cornering stiffness vs. slip angle plot for different wheel loads.

14.3.2 Understeer Gradient and Critical Speed


The lateral acceleration is approximated as aGy = ay ≈ V 2 /RS , where V is the vehicle
speed and RS is the turn radius in the presence of slip. A more accurate expression for
the lateral acceleration will be used later on, when modeling the kinetics. Introducing this
approximation to Equation (14.17), we can express the net steer effect as

Wf V 2 V2
 
Wr
δS = δ − − = δ − K0 (14.19)
Cr C f RS g RS g

in which the term


   
Wr Wf W b c
K0 = − = − (14.20)
Cr Cf L Cr Cf

is defined as the understeer gradient K 0 . When K 0 > 0 the vehicle understeers and when
K 0 < 0 the vehicle oversteers.
Most of the examples in this chapter assume that the cornering stiffnesses of the front and
rear tires are the same. As discussed earlier, as the wheel load increases, so does the cornering
stiffness. As a result, the actual understeer gradient will be of smaller magnitude than the
understeer gradient that is obtained by assuming the same cornering stiffnesses. Hence, the
analysis and examples in this chapter are conservative. For example, when considering a
vehicle with Wf /Wr = 1.2, if the cornering stiffnesses are taken as the same, the understeer
gradient becomes (note that Cf = Cr = C is negative)

1 Wr
K0 = (Wr − 1.2Wr ) = −0.2 (14.21)
C C
Vehicle Dynamics—Lateral Stability 721

When the previously discussed approximate model is used, the understeer gradient K 0
becomes
   
Wr Wf Wr Wr Wr
K0 = − = − 1.14 = −0.14 (14.22)
Cr Cf Cr Cr Cr

The actions described in Table 14.2 can be taken (either at the front or at the back axles
or on both axles) to change the amount of understeer.

TABLE 14.2
Actions that increase understeer properties

Vehicle Property Front Rear


Axle load Increase Decrease
Tire pressure Increase Decrease
Tire size Smaller Larger
Camber Increase Decrease
Toe in Increase Decrease
Suspension stiffness Higher Lower
Anti-roll bar Stiffer Softer

The opposite steps can be taken to increase oversteer. For example, an understeer vehicle
that is overloaded at the back can become an oversteer vehicle.
The unit of the understeer gradient is force divided by force/angle = angle. When the
cornering stiffnesses are expressed as force/radians, as in lb/radians, the understeer gradient
has units of radians, and when the cornering stiffness is expressed in terms of degrees, as in
lb/◦ , the understeer gradient is in degrees.
The turn radius for neutral steer (no slip) can be approximated as δ = L/RN S , so we
can write Equation (14.20) as

V2 L V2
δS = δ − K 0 = − K0 (in radians) (14.23)
RS g RN S RS g

V2 L V2
δS = 57.296δ − K 0 = 57.296 − K0 (in degrees) (14.24)
RS g RN S RS g
A preliminary study of instability for an oversteer vehicle can be conducted by means of
the oversteer instability, which is the case where the applied steer angle is zero or infinites-
imally small, δ = 0, but the net steer effect is L/RS . That is, at a certain speed, referred to
as critical speed and denoted by Vc , the slightest steering input by the driver generates slip
angles that make the vehicle turn in the path as if the driver had provided a steer input
L/RS . The vehicle becomes uncontrollable.
We can rewrite Equation (14.24) using radians and also when using degrees as

L V2 L V2
= −K 0 c (in radians) 57.296 = −K 0 c (in degrees) (14.25)
RS RS g RS RS g
which can be solved for the critical speed Vc as
r r
gL gL
Vc = (in radians) Vc = 57.296 (in degrees) (14.26)
−K 0 −K 0
722 Applied Dynamics

200

150

Vc (mph) 100

50

0
0 0.5 1 1.5 2 2.5 3 3.5 4
−K‘ (Degrees)

FIGURE 14.8
Critical speed for an oversteer vehicle, as a function of K 0 . L = 9 ft.

The schematic of oversteer is shown in Figure 14.3b. The front tires have more traction
than the rear tires. The front of the vehicle points toward the inside of the turn, forcing the
rear to slide to the outside. As the vehicle speed approaches critical speed, the rear tires
begin to slide and the vehicle becomes unstable.
A plot of critical speed vs. −K 0 is given in Figure 14.8 for a vehicle of wheelbase L = 9
ft. Recall that K 0 has a negative value for oversteer. While an oversteer vehicle is technically
stable when the speed is less than the critical speed, the vehicle becomes increasingly difficult
to control as the speed increases. Dangerous driving conditions may be reached even at
speeds lower than the critical speed.
For an understeer vehicle, the rear tires have greater traction than the front tires, and
they tend to push the front of the car toward the outside, as can be seen from Figure
14.2b. There is no destabilization issue for understeer vehicles, but the agility of the vehicle
(the quickness with which the vehicle responds to steering inputs) is reduced as the speed
increases. The net steering effect becomes smaller.
When the net steer effect approaches zero, the steering provided by the driver has no
effect on the turning motion of the vehicle and the front tires begin to skid. While the
vehicle does not become unstable, as it is moving along a straight line, it cannot take a turn
since it becomes nearly impossible to change direction. On a curved road, such a vehicle
strays from the lane in which it is supposed to travel.
While there is no critical speed associated with understeer vehicles, the speed at which
δS = δ/2 in Equation (14.24) is usually referred to as the characteristic speed. It follows
that
r r
gL gL
Vc (understeer) = (in radians) = 57.296 0 (in degrees) (14.27)
K0 K
Once the characteristic speed is exceeded, it becomes very difficult to steer the vehicle.
While the above analysis of instability provides insight as to when a vehicle becomes
unstable, it does not provide us with a clear picture as to the effects of external forces and
other parameters on the nature of the response of a vehicle. The next two sections discuss
a more detailed model based on the transient motion.
Vehicle Dynamics—Lateral Stability 723

14.3.3 Bundorf Compliances


As discussed earlier, several factors affect the understeer and oversteer properties of a vehi-
cle, in addition to the wheel loads and cornering stiffnesses. A convenient way to quantify
and add understeer and oversteer effects is by means of the Bundorf compliances. We de-
note the Bundorf compliances of the front and rear axles by Df and Dr , respectively. The
Bundorf compliances due to axle loads and cornering stiffnesses are defined as
Wf Wr
Df = − Dr = − (14.28)
Cf g Cr g
Considering that cornering stiffnesses are negative quantities, the Bundorf compliances
are positive. Their units are degrees per g (or radians per g). For steady-state motion, using
Equation (14.17), the Bundorf compliances and the slip angles are related by
αf = −Df aGy αr = −Dr aGy (14.29)
From Equation (14.20), the understeer gradient can be expressed (in radians) as
K 0 = (Df − Dr ) g (14.30)
Modeling of the factors that change understeer and oversteer properties is complicated
and beyond the scope of this text. These effects are usually measured. By adding their
contributions of these factors to the Bundorf compliances, we can calculate effective Bundorf
compliances. The effective compliances, which can alter the original Bundorf compliances
(calculated from axle loads and cornering stiffnesses) by about 10%, are then used for a
more accurate handling analysis.

14.3.4 Lateral Acceleration Gain


An interesting analysis involves describing lateral acceleration as a function of the steer
angle. The lateral acceleration is given by aGy = ay = V 2 /RS . The turn radius RS can be
obtained from the net steer effect (in radians) as
L
= tan δS = tan (δ + αf − αr ) ≈ δ + αf − αr (14.31)
RS
and, from Equation (14.17), the slip angles are
Wr ay Wf ay
αr = αf = (14.32)
Cr g Cf g
Substituting Equation (14.17) into Equation (14.31) gives
V2 V2
 
ay Wr ay Wf ay
= (δ + αf − αr ) = δ− + (14.33)
g gL gL Cr g Cf g
ay ay
Collecting terms involving g on one side, we can solve the above equation for g as

ay δV 2 /gL δV 2 /gL
=   = 2 (14.34)
g 2
1 + VgL W
Wf 1 + K 0 VgL
Cr −
r
Cf

ay
The expression for g above can be rewritten as the ratio of the lateral acceleration to
the steering input as
ay V 2 /L
= 0V 2 (14.35)
δ 1 + KgL
724 Applied Dynamics

This equation is known as the lateral acceleration gain. It is yet another way of char-
acterizing the critical speed. At critical speed, the smallest steering input δ will cause an
infinite lateral acceleration. Setting ay /δ = ∞, we can solve for the value of the critical
speed, which, of course, comes out as the same as Equation (14.26).
In terms of degrees, the lateral acceleration gain becomes

ay 1 V 2 /L V 2 /L
= K 0V 2 = 0V 2 (14.36)
δ 57.296 1 + 57.296gL 57.296 + KgL

Example 14.2
A vehicle has a wheelbase of L = 2.40 m and a mass distribution of mr = 590 kg, mf = 545
kg. The front and tires tires have the same cornering stiffness of 25,000 N/rad. Calculate
the speed at which this vehicle will become unstable.
Recalling that we need to multiply the cornering stiffness by two to get the axle stiffness,
the understeer gradient becomes
 
0 Wr Wf (590 − 545) × 9.807
K = − = = −8.826 × 10−3 rad [a]
Cr Cf −25, 000 × 2

The vehicle is an oversteer vehicle. From Equation (14.26), the critical speed becomes
r r
gL 9.807 × 2.4
Vc = 0
= = 51.64 m/s [b]
−K 8.826 × 10−3
Converting into miles per hour by noting that 90 kph = 25 m/s, the critical speed in
kph is
51.64 × 90
Vc = = 185.9 kph [c]
25
The vehicle considered here has a near 48/52 weight distribution and yet the critical
speed is not very high (around 115 mph). If this car had a 44/56 weight distribution, the
difference between the axle loads would be mr − mf = 135 kg, that is, three times the
previous difference of 45 kg. The
√ understeer gradient would also be three times higher. The
critical speed would then be 3 times lower, 107.3 kph (66.7 mph), a speed that is not high
at all, illustrating the dangers associated with oversteer.
Even if we consider that for such a weight distribution (44/56) the cornering stiffnesses
would be different, the critical speed will still be low. Vehicles carrying heavy loads have
either larger tires on the rear axle, two tires on each end of the axle, or a double axle. This
not only helps support the vertical loads, but also increases the understeer properties.

Example 14.3
The bicycle model of a vehicle has the following parameters: W = 3000 lb, L = 10 ft, axle
stiffnesses Cf = Cr = −300 lb/◦ , center of mass 4.5 ft from the front wheel. The vehicle
is traveling at 60 mph (88 ft/sec) while executing a turn, with the steer angle δ = 2◦ .
Calculate the slip angles for the front and rear tires.
The procedure is to find the lateral acceleration ratio ay /g and use it to find the slip
angles. To this end, we first calculate

V2 882 b
= = 24.07 ft/sec Wr = W = 0.45W Wf = 0.55W [a]
gL 32.17 × 10 L
and
Vehicle Dynamics—Lateral Stability 725

Wr 3000 × 0.45 Wf 3000 × 0.55


= = −4.50◦ = = −5.50◦ [b]
Cr −300 Cf −300
Introducing these values into Equation (14.34), and recalling that the slip and steer
angles are expressed in degrees, we obtain
ay 24.051ptδ 24.05 × 2/57.296
=  = = 0.591 [c]
g 1 + 24.05 −4.50+5.50
57.296
1 + 0.420
Note that calculation of the lateral acceleration without considering the slip angles would
give ay /g = V 2 δ/gL = 0.84, which is higher than the correct value for ay /g. This, of course,
is to be expected. As the vehicle is an understeer vehicle, the turn radius is larger and the
net steering effect is less than a neutral steer vehicle. The slip angles become
Wr ay Wf ay
αr = = −4.50◦ ×0.591 = −2.66◦ αf = = −5.50◦ ×0.591 = −3.25◦
Cr g Cf g
[d]
Let us next calculate the net steering effect, which becomes
δS = δ + αf − αr = 2 − 3.25 + 2.66 = 1.41◦ [e]
which is less than the steer angle, as the vehicle is an understeer vehicle. Finally, in this
example the steer angle is quite high, resulting in a lateral acceleration about 0.6g, which
is close to the limits of the lateral acceleration of a passenger vehicle on a dry road.

14.4 Slip Angle Kinematics


We next develop a more accurate model that takes into consideration the transient motion.
To do this, it is necessary to relate the slip angles to the vehicle speed and angular velocity.
Also, we need to decide which motion variables to use. The analysis begins by revisiting the
kinematics of slip. Again, because of the need to relate slip to vehicle velocities and angular
velocities, the slip angles are drawn in their positive directions (and not in the directions
we expect them to be), as shown in Figure 14.9.

A G u B
r x
  
vA 
v vG f
c b vB x' y
y'

FIGURE 14.9
Slip angles.

The slip angles describe the direction of the wheel speed. Hence, it is possible to express
them in terms of velocity ratios (lateral speed over horizontal speed). Also of interest is the
velocity ratio of the center of mass of the vehicle. Denote the velocity of the center of mass
of the vehicle by
vG = ui + vj (14.37)
726 Applied Dynamics

and the angular velocity by ω = ωk. The velocities of the rear and front of the vehicle are
expressed as

vA = vAx i + vAy j vB = vBx i + vBy j = vBx0 i0 + vBy0 j0 (14.38)

where the velocity terms are obtained from relative velocity expressions. For the rear wheel,

vA = vG + ωk × rA/G = ui + vj + ωk × (−ci) = ui + (v − cω) j (14.39)

so vAx = u, vAy = v − cω. Similarly, the velocity of the front wheel, vB , is

vB = vG + ωk × rB/G = ui + (v + bω) j (14.40)

so vBx = u, vBy = v + bω.

vA x u vB x
A x G x B x
r  +f
vA y v vB y
vA vG vB

y y y

FIGURE 14.10
Velocities of the tires and of the center of mass (drawn not to scale and drawn in the positive
directions of the slip angles).

The velocities of the tires are in the direction of slip, as illustrated in Figure 14.10.
Hence, we relate the components of the velocities to the slip angles by
vB y 0
     
vAy vBy
αr = tan−1 αf = tan−1 = tan−1 −δ (14.41)
vAx vBx0 vB x

Also of interest is the angle β, referred to as the body slip angle, or sideslip angle, and
defined as the angle that the velocity of the center of mass makes with the wheelbase
v
β = tan−1 (14.42)
u
The body slip angle indicates the direction in which the center of mass is moving (drift-
ing). At low speeds, where the slip angles are small, the sideslip angle is towards the turn
center. At high speeds, the sideslip angle is away from the turn, and the center of mass of
the vehicle drifts outward. This effect is felt by drivers when taking a turn at higher speeds.
It is preferable to use the velocity components of the center of mass to express the slip
angles. We will use the angular velocity ω and sideslip angle β as our variables. Using the
relative velocity relationship in Equation (14.39) and invoking the small angles assumption,
the slip angle for the rear wheel becomes
   
vAy v − cω v − cω
αr = tan−1 = tan−1 ≈ (14.43)
vA x u u

Next, return to Figure 14.9 and consider the slip angle for the rear wheel. For the sake
of argument, let v = 0 for an instant. The above equation indicates that vAy < 0, so the
Vehicle Dynamics—Lateral Stability 727

slip angle is negative, which is what was discussed when considering slip in Figure 14.1. A
similar analysis for the front slip angle and for the sideslip angle results in
v + bω v
αf ≈ −δ β ≈ (14.44)
u u
The acceleration of the center of mass is obtained next, as this term will be necessary for
calculating the force balances. Recalling that the velocity of the center of mass is vG = ui+vj
and the angular velocity of the reference frame is ω = ωk, the acceleration of the center of
mass is obtained by differentiating vG by means of the transport theorem as

aG = (v̇G )rel + ω × vG = u̇i + v̇j + ωk × (ui + vj)

= (u̇ − ωv) i + (v̇ + ωu) j = aGx i + aGy j (14.45)

It is customary to assume that the horizontal speed is constant or nearly constant. Thus,
we can eliminate
√ the u̇ term from the formulation and use the approximation u ≈ V , in
which V = u2 + v 2 is the vehicle speed. The rate of change of the body slip angle is then
approximated as
d v v̇ v u̇ v̇ v̇
β̇ ≈ = − ≈ ≈ (14.46)
dt u u u2 u V
Using the approximation for the body slip angle in Equation (14.44), components of the
acceleration become

aGx = u̇ − ωv ≈ −ωv = −ωβV aGy = v̇ + ωu ≈ V β̇ + V ω (14.47)

The expression for the lateral acceleration in Equation (14.47) is different from aGy =
V 2 /RS , as the motion considered here is not steady-state. It is this expression for aGy that
needs to be used when there are nonzero values for the rate of change of the sideslip angle
and angular velocity. Also, the y direction is not the normal direction for the motion of
the center of mass G. However, the relationship an = V 2 /RS still holds along the normal
direction.
Recall that the relationship between the net steering effect and slip angles is given by
Equation (14.10). This relationship is valid for steady-state as well as transient motion. The
net steer angle δS is related to the radius of curvature by δS = RLS , with RS denoting the
instantaneous radius of curvature and
L
≈ δ + αf − αr (14.48)
RS
which can be solved for the radius of curvature given the slip angles.

Example 14.4
Consider the previous example and calculate the velocity of the center of mass and the
angular velocity of the vehicle.
The angular velocity of the vehicle can be approximated by the relationship ω ≈ V /RS ,
where V = 88 ft/sec is the component of the velocity in the x direction and RS is the turn
radius, which is
L 10 × 57.296
RS = = = 406.4 ft [a]
δ + αf − αr 2 − 3.25 + 2.66
728 Applied Dynamics

so that
V 88
ω ≈ = = 0.217 rad/sec cw [b]
RS 406.4
The lateral velocity and the sideslip angle are found from from Equations (14.43)–
(14.44). Using Equation (14.43) the lateral velocity of the center of mass becomes
−2.66
v = αr u + cω = × 88 + 5.5 × 0.217 = −2.89 ft/sec [c]
57.296
The accuracy of the calculation above can be checked by Equation (14.44), which gives
−3.25 + 2
v = (αf + δ) u − bω = × 88 − 4.5 × 0.217 = −2.89 ft/sec [d]
57.296
The velocity of the center of mass is
vG = ui + vj = 88i − 2.89j ft/sec [e]
As expected, the lateral component of the
√ velocity is much smaller than the horizontal
component, justifying the assumption V = u2 + v 2 ≈ u. The sideslip angle β can be found
from Equation (14.44) as
v 2.89 2.89 × 57.296
β ≈ = − rad = − = −1.882◦ [f ]
u 88 88
which is not very small. Also, the sideslip angle is negative, a result expected for high-speed
turns.

14.5 Transient Motion Equations


We next consider the force and moment balances and derive the transient handling equa-
tions. We assume that the vehicle is moving with constant speed.
Figure 14.11 shows the free-body diagram of the vehicle. Note that the slip angles
have been marked in their positive directions, which is clockwise, and not in their actual
directions. For the SAE coordinate system used here, a positive rotation about the z axis
is clockwise when viewed from the top.
The next step is to write the force and moment balances and the acceleration expressions
in terms of the angular velocity ω and sideslip angle β, as derived in Equation (14.47). As
we are dealing with force and moment balances, all angles are in radians. Summing forces
in the x direction gives
+
X
→ Fx = maGx =⇒ FTr − FLr sin δ + FTf cos δ = −mV ωβ (14.49)

The force balance in the y direction becomes


X  
+↓ Fy = maGy =⇒ FLr + FLf cos δ + FTf sin δ = mV β̇ + ω (14.50)

and the moment balance is


X
 MG = IG ω̇ = Iψ ω̇ =⇒ −cFLr + bFLf cos δ + bFTf sin δ = Iψ ω̇ (14.51)

where FTr and FTf are tractive forces, FLr and FLf are lateral forces, and IG = Iψ is the
yaw moment of inertia about the center of mass.
The following simplifying assumptions are used to derive the transient motion equations:
Vehicle Dynamics—Lateral Stability 729

F Lr c b
F Tf F Lf
B
F Tr G u
x x
r  
A vA v 
V y
y' y
f x'
vB

FIGURE 14.11
Free-body diagram, top view.

• The steer angle δ and the slip angles αr and αf are small.
• The vehicle is not gaining speed, so that the tractive forces are there only to maintain
speed by countering rolling resistance, aerodynamic forces, and induced drag due to
lateral tire forces.
• There is relatively low lateral acceleration (aGy not much larger than 0.3g), otherwise
nonlinear effects become significant.
• Because the slip angles are small, the cornering stiffnesses Cf and Cr are constant.
Also consider a rear wheel drive vehicle, so FTf = 0. The lateral forces are in the positive
y and y 0 directions and are expressed as

FLr = Cr αr FLf = Cf αf (14.52)

where we recall that both the cornering stiffnesses and slip angles are negative quantities.
Introducing Equation (14.52) into the force and moment equations, the forces in the x
and y directions and the moment about the center of mass become
+
X
→ Fx = FTr − Cf αf sin δ ≈ FTr − Cf αf δ (14.53)

X
+↓ Fy = Cr αr + Cf αf cos δ ≈ Cr αr + Cf αf (14.54)

X
 MG = −cCr αr + bCf αf cos δ ≈ −cCr αr + bCf αf (14.55)

Next, the above expressions are introduced into the force and moment balances. The
existence of external forces, such as aerodynamic force, rolling resistance, and side force, as
well as the effect of a banked road, also need to be included in the model. The resultants
of all external forces and moments are denoted as a force Fx i + Fy j acting at the center of
mass and a moment M about the z axis, as shown in Figure 14.12. We also use a small
angles approximation for the steer angle δ.
Using as motion variables the angular velocity ω (yaw rate) and the sideslip angle β,3
as well as the expressions for the slip angles from Equation (14.43) and Equation (14.44),
the equations of motion are obtained as
3 Some prefer to use the sideslip velocity v as a motion variable instead of the sideslip angle β.
730 Applied Dynamics

c b
A B
G M
x
Fx 

Fy y

FIGURE 14.12
Resultant of external forces and moments acting on bicycle model.

 

In x direction: − mV ωβ = FTr + Cf + β − δ δ + Fx (14.56)
V

 
1
In y direction: mV β̇ = (Cf + Cr ) β + (bCf − cCr ) − mV ω − Cf δ + Fy (14.57)
V

1 2
b Cf + c2 Cr ω − bCf δ + M

About z axis: Iψ ω̇ = (bCf − cCr ) β + (14.58)
V
The force balance in the tractive (x) direction involves the external force Fx , as well as
terms that are nonlinear in the motion variables and the steer angle, which arise from the
induced drag expression in Equation (13.16). The tractive force needed to maintain speed
has the form
 

FTr = −mωβV − Cf + β − δ δ − Fx (14.59)
V

The force balance in the y direction and the moment balance about the z direction are
recognized as the two linearized equations of motion, in terms of the two motion variables
β and ω. The steering angle δ is the input provided by the driver.
It is of interest to express the stability equations in matrix form. Let us introduce the
motion variable vector {x} and the input vector {F }, which are defined as

−Cf
     
β Fy
{x} = {F } = δ+ (14.60)
ω −bCf M

and the matrices


" # " #
1
mV 0 Cf + Cr V (bCf − cCr ) − mV
[M ] = [A] = 1 2 2
 (14.61)
0 Iψ bCf − cCr V b Cf + c Cr

The stability equations are written as

[M ]{ẋ} = [A] {x} + {F } (14.62)

In the equations of motion above, all angles and cornering stiffnesses are in terms of
radians.
Vehicle Dynamics—Lateral Stability 731

14.6 Response
After deriving the stability equations for the lateral motion, the next step is to analyze
and solve these equations, as well as extract information from them. We can analyze the
equations of motion in several ways:
• Conduct an eigenvalue analysis to assess stability and decay rates of yaw and sideslip,
• Conduct an analytical study by examining the contribution of each term to the stability
equations and extract information from these terms,
• Solve the stability equations for steady-state motion,
• Do numerical analysis, using a numerical integration routine to integrate the stability
equations and evaluate the response.
The stability equations are coupled and they cannot be solved independent of each other.
An easy way of seeing this is to note that matrices [A] and [B] have off-diagonal terms that
are nonzero. A useful way of qualitatively analyzing the stability equations is by giving each
term in these equations a physical interpretation. Used commonly in fluid mechanics, this
approach is known as the stability derivatives approach. We rewrite the stability equations
for the yaw rate ω and sideslip angle β, Equations (14.57)–(14.58), as
 
mV ω + β̇ = f1 (β, ω, δ) + Fy Iψ ω̇ = f2 (β, ω, δ) + M (14.63)

where f1 and f2 are functions of the yaw rate, sideslip angle, and steer angle, and they
include linear as well as nonlinear effects acting on the vehicle. The stability derivatives are
the derivatives of f1 and f2 with respect to ω, β, and δ that are linearized and have the
forms
∂f1 ∂f1 ∂f1
Yβ = Yω = Yδ =
∂β ∂ω ∂δ

∂f2 ∂f2 ∂f2


Nβ = Nω = Nδ = (14.64)
∂β ∂ω ∂δ
so that the stability equations are expressed as
 
mV ω + β̇ = Yβ β + Yω ω + Yδ δ + Fy (14.65)

Iψ ω̇ = Nβ β + Nω ω + Nδ δ + M (14.66)

Rather than deriving the functions f1 and f2 , we can identify the stability derivatives by
comparing the above equations with Equation (14.57) and Equation (14.58). The stability
derivatives have interesting physical explanations:
Nβ = (bCf − cCr ). This term provides coupling between the two equations. It is a
measure of the rate at which the lateral force is developed and is calleddirectional
stability. It can be thought of as an imaginary spring between the heading of the vehicle
and its velocity. For understeer vehicles this term produces a weathercock effect, similar
to what happens when an arrow traveling in air encounters a side force.
732 Applied Dynamics

Nω = V1 b2 Cf + c2 Cr . This term is always negative and it describes the amount of




decay the yaw angle has, or the damping of the yaw angular velocity. It is referred to
as the damping-in-yaw. Note that the value of this term decreases with higher speed,
so that the faster the speed, the slower the decay of the yaw angular velocity.
Nδ = −bCf . This term is referred to as the control moment. It is always positive and it
describes the effect of steering on the yaw.
Yβ = Cf + Cr . This term is always negative and it provides damping for the sideslip.
It is referred to as damping-in-sideslip. Note that, unlike the damping-in-yaw term, the
level at which the sideslip angle is damped is not affected by speed.
Yω = V1 (bCf − cCr ). This term is similar to Nβ in that it provides coupling to the
stability equations. It is referred to as lateral force/yaw coupling derivative.
Yδ = −Cf . This term is always positive and it describes the effect of steering on the
lateral motion and sideslip. It is known as the control force.
The two coupling terms, Nβ and Yω , are similar and they are indicators of oversteer or
0
understeer. Multiplying Nβ by CfW
Cr L gives the understeer gradient K

W Wb Wc Wr Wf
Nβ = − = − = K0 (14.67)
Cf Cr L LCr LCf Cr Cf
It follows that Nβ and Yω are positive quantities for understeer vehicles, and they are
negative for oversteer vehicles. In each equation there is a control term, a damping term,
and a coupling term. The equations of motion can be written in terms of the stability
derivatives as
Yβ Yω − mV
         
mV 0 β̇ β Yδ Fy
= + δ+ (14.68)
0 Iψ ω̇ Nβ Nω ω Nδ M

14.6.1 Numerical Integration of the Stability Equations


To numerically integrate the stability equations, we need to put them into state form. As
[M ] is positive definite, Equation (14.62) can be cast in state form by left multiplying it by
[M ]−1 , with the result
{ẋ} = [B]{x} + {F 0 } (14.69)
where
 
Cf +Cr 1
mV mV 2 (bCf − cCr ) − 1
[B] = [M ]−1 [A] =  bCf −cCr  
1 2 2
Iψ Iψ V b Cf + c C r

 
C " Fy #
− mVf mV
0 −1
{F } = [M ] {F } =  bCf
δ +
M
(14.70)
− Iψ Iψ

Integration of the stability equations gives the response for the sideslip angle and angular
velocity. It also is of interest to obtain the response of the heading angle ψ, the angle between
the inertial X axis and the x axis that is moving with the vehicle, as shown in Figure 14.13.
The time derivative heading angle is related to the angular velocity (yaw rate) by
ψ̇ = ω (14.71)
If desired, this kinematic relationship can be integrated alongside Equation (14.69),
resulting in a system of three differential equations in state form.
Vehicle Dynamics—Lateral Stability 733

X
A
vA 

B x

y Y

FIGURE 14.13
Vehicle heading angle (top view).

Example 14.5
Consider the bicycle model of a vehicle with W = 1700 kg, Iψ = 4000 kg·m2 , and wheelbase
L = 2.7 m. All tires have the same cornering stiffness of 500 N/deg. Numerically integrate
the stability equations for eight seconds for a steering input δ(t) in the form:
For t < 2 seconds, δ(t) = 1◦
For 2 ≤ t < 4 seconds, δ(t) = 0◦
For 4 ≤ t ≤ 6 seconds, δ(t) = −1◦
For t > 6 seconds, δ(t) = 0◦
Do the simulation for the different speeds and understeer characteristics in Table 14.3
and plot the sideslip angle β, angular velocity ω, and heading angle ψ.

TABLE 14.3
Speed and distance c from rear axle for the different cases

Case No. Speed (m/s) c (m) Plotted in


Case 1 13.4 1.5 Figure 14.14
Case 2 26.8 1.5 Figure 14.15
Case 3 13.4 1.2 Figure 14.16
Case 4 26.8 1.2 Figure 14.17

We use the MATLABr program ode45, integrate the three differential equations in
Equations (14.70) and (14.71), and plot the results. Note that the scales of the plots are
different for each case. Case 1 results, shown in Figure 14.14, are for an understeer vehicle
with relatively low speed. The plots are indicative of a stable vehicle, with the motion
variables β and ω reaching steady-state very quickly.
As expected, the results shown in Figure 14.15 for the faster speed of 26.8 m/s lead to
twice as large sideslip angles and angular velocity, with steady-state taking a bit longer to
reach.
We next consider Case 3, which is an oversteer vehicle, driven at a relatively low speed.
While the results, shown in Figure 14.16, are stable, the vehicle is not nearly as stable as
the understeer vehicle in Case 1. Also, the angular velocity and sideslip angle are much
larger than Case 1.
The dangerous consequences of driving an oversteer vehicle can be observed in Case 4
(Figure (14.17)). The vehicle is very close to becoming destabilized. This can be expected,
because the critical speed for this vehicle can be shown to be Vc = 28.1 m/s. The response
734 Applied Dynamics

0.15
β
0.1 ω
ψ
Amplitudes
0.05

−0.05

−0.1
0 1 2 3 4 5 6 7 8
Time (sec)

FIGURE 14.14
Response for Case 1.

0.3
β
0.2 ω
ψ
Amplitudes

0.1

−0.1
0 1 2 3 4 5 6 7 8
Time (sec)

FIGURE 14.15
Response for Case 2.

0.6
β
0.4 ω
Amplitudes

ψ
0.2

−0.2
0 1 2 3 4 5 6 7 8
Time (sec)

FIGURE 14.16
Response for Case 3.
Vehicle Dynamics—Lateral Stability 735

is slow to stabilize, with very high angular velocity and sideslip angle. The difference in
results between Case 2 and Case 4, both of which involve the same speed, is striking.

1
β
ω
0.5 ψ
Amplitudes

−0.5
0 1 2 3 4 5 6 7 8
Time (sec)

FIGURE 14.17
Response for Case 4.

14.7 Eigenvalue Analysis


This section makes use of the developments in Chapters 6 and 7 and presents an eigenvalue
analysis of the stability equations. To this end, we first consider the free response of the
describing differential equations, that is, no excitations or inputs to the system, so δ = 0,
Fy = 0, M = 0. With no external inputs, the stability equations reduce to

[M ] {ẋ} = [A] {x} (14.72)

Consider a solution {x (t)} = {X} eλt , where {X} denotes the amplitude of the solution
and λ the time dependence, and introduce this to Equation (14.72), with the result

(λ [M ] − [A]) {X} eλt = {0} (14.73)

The term eλt cannot be not zero (otherwise we would obtain the trivial solution of zero)
so that the term

(λ [M ] − [A]) {X} = {0} (14.74)

must vanish for a nontrivial solution.


The above equation represents a set of two linear equations in terms of the unknown
vector {X} and in terms of the parameter λ. Because the right side of the above equation
is zero, the only way a solution other than {X} = {0} (trivial solution) can exist is when
the matrix on the left is singular, so that its determinant is zero

det (λ [M ] − [A]) = 0 (14.75)

Equation (14.74) describes the eigenvalue problem. The determinant is a second order
736 Applied Dynamics

polynomial in λ, known as the characteristic polynomial or characteristic equation. Denoted


by λ1 and λ2 , the two roots of the characteristic polynomial, that is, the values of λ that
make the equality in Equation (14.75) hold, are called the eigenvalues, and they dictate the
nature of the solution. There are three possibilities for the values of λ:
• The roots λ1 and λ2 are real and negative or they are complex conjugates with negative
real parts. In this case, the solution for {x (t)} is in the form of a decaying exponential
or an exponentially decaying sinusoid. The variables β and ω, the sideslip angle and
angular velocity, become zero with time. This case is known as stable.
• λ1 and λ2 are real with at least one positive root or they are complex conjugates with
positive real parts. The solution for {x (t)} is in the form of a growing exponential or an
exponentially growing sinusoid. The angular velocity and sideslip angle grow with time.
An undesirable situation, this case is known as unstable.
• λ1 and λ2 are pure imaginary. In this case, the solution does not grow with time, but
it does not decay either. This case is known as critically stable, and it constitutes a
transition point between stability and instability. Because the equations we are dealing
with are linearization of nonlinear equations, it is necessary to examine the nonlinear
terms to ascertain the actual behavior of the angular velocity and sideslip angle.4
The matrix [A] has the vehicle speed in it as a parameter, so we can analyze the eigenval-
ues as a function of the vehicle speed. For example, oversteer vehicles become unstable when
the speed of the vehicle reaches critical speed. At this point, at least one of the eigenvalues
becomes positive or acquires a positive real part.
Even though the eigenvalue problem in Equation (14.75) is of order two and the solution
can be obtained by hand in closed form, because of the complexity of the terms involved,
a closed form solution may not be descriptive. In such cases, one can conduct a qualitative
analysis of the eigenvalues as follows.
The characteristic polynomial associated with the stability equations can be written as
p0 λ2 + p1 λ + Q = 0 (14.76)
where the coefficients p0 and p1 have the form
p1 = −Iψ (Cf + Cr ) − m b2 Cf + c2 Cr

p0 = mV Iψ (14.77)
Both coefficients are always positive (recall that the cornering stiffnesses are negative).
The third coefficient, Q, can be shown to be the determinant of the matrix [A] and has the
form
 
bCf − cCr
Q = det [A] = − (bCf − cCr ) − mV
V

1 1
(Cf + Cr ) b2 Cf + c2 Cr = Cf Cr L2 + mV (bCf − cCr )

+ (14.78)
V V
where L = b + c is the wheelbase.
The roots of the characteristic polynomial are dictated by the values of p0 , p1 , and Q.
When all these parameters are greater than zero, all roots have negative real parts and
the system is stable. We will quantify the statement above by comparing the characteristic
equation with that of a damped oscillator, whose analysis was carried out in Chapter 6. The
next section discusses an analogy between the stability equations and a mass-spring-damper
system.
4 See Section 5.8 for more details.
Vehicle Dynamics—Lateral Stability 737

Example 14.6
Consider a vehicle with wheelbase of L = 2.35 m, mass 1200 kg, center of mass at 0.55L from
the rear axle, and yaw moment of inertia Iψ = IG = 1250 kg·m2 . The cornering stiffnesses
for all tires is C = 500 N/deg. Calculate and tabulate the roots of the characteristic equation
as a function of the vehicle speed.
This is an understeer vehicle. To find the eigenvalues, we need to calculate the [M ]
and [A] matrices. To this end, it is necessary to express the cornering stiffnesses in terms of
radians. Also, the cornering stiffness of the axles have to be calculated as twice the cornering
stiffnesses of each wheel; each axle has a cornering stiffness of 2 × 500 = 1000 N/deg. We
convert the cornering stiffnesses into radians by

Cf = Cr = −2C = −2 × 500 × 180/π = −57, 296 N/rad [a]

The center of mass is at a distance 0.55L = 0.55 × 2.35 = 1.2925 m from the rear of the
vehicle. The eigenvalues, obtained numerically, are listed in Table 14.4.

TABLE 14.4
Eigenvalues of understeer vehicle, c = 0.55L

Speed (m/s) Eigenvalues for c = 0.55L


5 −20.20, −24.46
6 −18.61 ± i0.389
10 −11.17 ± i2.636
20 −5.583 ± i3.133
50 −2.233 ± i3.256
100 −1.117 ± i3.276

As expected, the eigenvalues all have negative real parts and the real part of the eigen-
values become smaller with increasing speed. This is an indication that the vehicle becomes
less responsive (less agile) with increased speed.
Next, consider this vehicle to be an oversteer vehicle of similar ratio, that is, b = 0.55L.
The understeer ratio in this case becomes
 
Wr Wf W
K0 = − = (b − c)
Cr Cf LC

1, 200 × 9.81
= 0.235 = −0.0205 rad = −1.1772◦ [b]
2.35 × (−57, 296)
From Equation (14.26), the critical speed becomes
r
gL
Vc = 57.296 = 33.50 m/s = 120.6 km/h [c]
−K 0
Let us compare the eigenvalues for the oversteer case for varying speeds. Table 14.5 gives
the results.
One of the eigenvalues has become quite small at a speed of 30 m/s. This confirms
the discussion earlier that as the vehicle speed approaches critical speed, even though the
unstable region has not yet been reached, the stability margin becomes dangerously small.
738 Applied Dynamics

TABLE 14.5
Eigenvalues of oversteer vehicle, b = 0.55L

Speed (m/s) Eigenvalues for b = 0.55L


5 −17.22, −27.436
10 −7.346, −14.99
20 −2.159, −9.008
25 −1.092, −7.840
30 −0.3760, −7.069
35 0.1389, −6.520

Example 14.7
Consider the vehicle in the previous example and calculate the critical speed for oversteer
as a function of the center of mass location. The cornering stiffnesses remain constant.
The distance ratio b/L is varied from 0.5 to 1, as shown in Figure 14.18, in which region
the vehicle becomes an oversteer vehicle. The results are given in Table 14.6.

b
G v
x

y
L

FIGURE 14.18
Varying the center of mass location (top view).

TABLE 14.6
Critical speed of oversteer vehicle, as a function of center of mass location ratio b/L

Location Ratio b/L Critical Speed (m/s)


0.51 74.90
0.55 33.50
0.60 23.68
0.70 16.75
0.80 13.67
0.90 11.84
0.99 10.70

It is interesting to note that as b/L goes from 0.55 to 0.60 the critical speed becomes
2/3 of its value and as b/L goes from 0.55 to 0.70 the critical speed gets reduced by half.
As discussed earlier, one way of converting an understeer vehicle into an oversteer vehicle
(or to increase the amount of oversteer) is to add weight to its back.
This example assumes that the mass moment of inertia for yaw, Iψ , remains the same,
Vehicle Dynamics—Lateral Stability 739

and the cornering stiffnesses do not change, as the center of mass moves towards the rear,
so the results here are conservative and, for large values of b/L, unrealistic. Nevertheless,
the substantial change in the critical speed should serve as a warning.

14.8 Mass-Spring-Damper Analogy


The characteristic equation associated with the stability equations can be derived in a way
that is analogous to the characteristic equation of a single-degree-of-freedom mass-spring-
damper system. This permits an elegant quantification of lateral stability.

a) b)
k kx
m m

c cx
mg N
x

FIGURE 14.19
a) Mass-spring-damper in translational motion, b) free-body diagram.

Consider the mass-spring-damper system in Figure 14.19a. The free-body diagram is


given in Figure 14.19b. The equation of motion in the horizontal direction can be written
for free motion (no external excitations) as

mẍ + cẋ + kx = 0 (14.79)

in which m, c, and k are the mass, stiffness and damping coefficients, respectively. Assuming
a solution in the form x (t) = Xeλt , the characteristic equation has the form

mλ2 + cλ + k = 0 (14.80)

From vibration analysis, when all three coefficients m, c, and k are positive, the system
is called damped. The response of a damped system is in the form of a decaying exponential
or a decaying sinusoid. When the damping constant c is negative or if one of m or k is
negative, the roots of the characteristic equation have positive real parts and the system
becomes unstable.
The characteristic equation associated with the lateral stability equations, Equation
(14.76), is analogous to the characteristic equation of the mass-spring-damper system, with
m = p0 , c = p1 , k = Q. It was observed earlier that p0 and p1 are positive. Hence, the
stability of the transient equations depends on the value of Q. For stability, Q needs to
be positive. We can show that, irrespective of the values of p0 and p1 , the lateral motion
equations are not stable when Q < 0. Setting Q = 0 and solving for the critical speed Vc
gives the limiting condition for stability as

Cf Cr L2
Vc2 = − (14.81)
m (bCf − cCr )
740 Applied Dynamics

By substituting the values mb/L = Wr /g, mc/L = Wf /g, keeping in mind that in the
above equation the cornering stiffnesses are in radians, and by making the proper conversion
to degrees, we can show that expression above for the critical speed is the same as Equation
(14.26).
The above equation for the critical speed can be rewritten as
 
−1 m (bCf − cCr ) mg b/L c/L
= = − (14.82)
Vc2 Cf Cr L2 Lg Cr Cf
Recalling the definitions of the axle loads as Wf = W c/L, Wr = W b/L and ay =
V 2 /RS g is the acceleration in g, as well as αf = Wf ay /Cf g, αr = Wr ay /Cr g, and in-
troducing them into the above equation permits writing of the critical speed expression
as
−1 1 αr − αf 1 αr − αf
= = = K (14.83)
Vc2 L V 2 /RS L ay
where K is known as the stability factor, so the critical speed can be expressed in terms of
the slip angles by
r s
−1 Lay
Vc = = (14.84)
K αf − αr

In the above equation, the slip angles are in terms of radians. The stability factor K
and understeer gradient K 0 are related by
K0
 
1 Wr Wf
K = = − (14.85)
gL gL Cr Cf
When analyzing the eigenvalues and response of a damped system, as discussed in Chap-
ter 6, it is advantageous to express the damping properties in terms of a damping constant
and natural frequency.
p Dividing Equation (14.79) by m and introducing the natural fre-
quency ωn = k/m and damping factor ζ so that c/m = 2ζωn , it is shown in Chapter 6
that the roots of the characteristic equation, or the eigenvalues, can be expressed as
p
For underdamped systems (ζ < 1) =⇒ λ1,2 = −ζωn ± iωn 1 − ζ 2

p
For overdamped systems (ζ ≥ 1) =⇒ λ1,2 = −ζωn ± ωn ζ2 − 1 (14.86)

For an underdamped system, given the eigenvalues


p
λ1,2 = −a ± ib = −ζωn ± iωn 1 − ζ2 (14.87)

so

a = ζ 2 ωn2 b2 = ωn2 1 − ζ 2

(14.88)

We can solve for ζ and ωn from the above equation by


p a
ωn = a2 + b2 ζ = √ (14.89)
a + b2
2

Generic plots of the natural frequency and damping factor as a function of the vehicle
speed V are shown in Figure 14.20.
A similar analysis can be used to calculate ζ and ωn when the eigenvalues are both real
and negative.
Vehicle Dynamics—Lateral Stability 741

n 

"#$%&'()#*+c/L "#$%&'()#* c/L

! V ! V

FIGURE 14.20
a) Natural frequency ωn vs. speed V , b) damping factor ζ vs. speed V for an understeer
vehicle.

TABLE 14.7
Damping factor and natural frequency of understeer vehicle, c = 0.55L

Speed (m/s) Eigenvalues ζ ωn


5 −20.20, −24.46 1.0046 22.23
6 −18.61 ± i0.389 0.9998 18.61
10 −11.17 ± i2.636 0.9732 11.47
20 −5.583 ± i3.133 0.8721 6.402
50 −2.233 ± i3.256 0.5633 3.951
100 −1.117 ± i3.276 0.3226 3.461

Example 14.8
Consider Example 14.6 and calculate the associated damping factors. Using Equation
(14.89), the results are shown in Table 14.7.
Next, consider a vehicle with higher understeer properties, say c = 0.65L and repeat the
calculations to find the eigenvalues, damping factor, and natural frequencies. The results
are shown in Table 14.8.

TABLE 14.8
Damping factor and natural frequency of understeer vehicle, c = 0.65L

Speed (m/s) Eigenvalues ζ ωn


5 −17.94, −28.75 1.0279 22.71
10 −11.67 ± i4.115 0.9431 12.38
20 −5.584 ± i5.336 0.7381 7.908
50 −2.233 ± i5.630 0.3830 6.095
100 −1.117 ± i5.671 0.2016 5.790

A few interesting observations can be made by comparing Tables 14.7 and 14.8.

• Values of the damping factor are lower for a vehicle with more understeer (except at
5 m/s). This is expected, as a lower damping factor is an indication that the transient
742 Applied Dynamics

motion dies out more slowly, and that the vehicle is slower to react to the steering input.
It is less agile.
• The natural frequencies are higher when there is more understeer. This means that
there is more oscillation in the response. This is also expected, because with a vehicle
with larger understeer, the response to a steering input will not reach its desired value
rapidly. Also, it will take longer for the yaw rate amplitudes to subside.

For an oversteer vehicle, the damping factors and natural frequencies are not as good
indicators of instability as the actual eigenvalues. This is because, before critical speed is
reached, both eigenvalues are real and negative, and while the damping factor becomes
larger, the natural frequency approaches zero. Also, one of the eigenvalues approaches zero.
This eigenvalue becomes positive once the speed exceeds the critical speed and it leads to
the unstable behavior.
The damping factors and natural frequencies associated with the oversteer vehicle, whose
eigenvalues are given Table 14.5, (b = 0.55L and Vc = 33.50 m/s for this case) are given in
Table 14.9. Note that as the speed V approaches the critical speed Vc , the natural frequency
ωn approaches zero. This is expected, as at critical speed we have Q = 0.

TABLE 14.9
Eigenvalues and damping factor of oversteer vehicle, b = 0.55L. The critical speed is Vc =
33.50 m/s

Speed (m/s) Eigenvalues ζ ωn


5 −17.22, −27.44 1.0272 21.74
10 −7.346, −14.99 1.0642 10.49
20 −2.159, −9.008 1.2662 4.410
25 −1.092, −7.840 1.5263 2.927
30 −0.3760, −7.069 2.2832 1.6302
33.49 −0.0007, −6.677 50.04 0.0666

14.9 Steady-State Response


In the previous five sections, we derived the describing equations of a vehicle for transient
motion and conducted a stability analysis. This section returns to the steady-state response
discussed in the beginning of the chapter, and it expands the steady-state analysis of earlier
sections in this chapter. Steady-state analysis permits analysis basic phenomena, as well as
the long-term effects of applied forces and moments.
As discussed in Chapter 6, at steady-state the transient response has died out. To obtain
the steady-state solution, the derivative terms in Equation (14.68) are set to zero, ω̇ = 0,
β̇ = 0, and we solve for the values of the angular velocity and sideslip angle, having been
given values of the steer angle and/or applied loading (and vice versa). Note also that for
steady-state, ω = V /RS . The steady-state equations become

Yβ Yω − mV
      
βss Yδ Fy
− = δ+ (14.90)
Nβ Nω ωss Nδ M
Vehicle Dynamics—Lateral Stability 743

The solution can be obtained by inverting [A]


     
βss Yδ Fy
= −[A]−1 δ+ (14.91)
ωss Nδ M
   
a b d −b
Noting that the inverse of a matrix [D] = is [D]−1 = 1
, the
c d ad−bc −c a
inverse of [A] has the form
 −1  
Yβ Yω − mV 1 Nω −Yω + mV
[A]−1 = = (14.92)
Nβ Nω Q −Nβ Yβ

where Q was derived in Equation (14.78) and in terms of the stability derivatives has the
form

Q = −Nβ Yω + mV Nβ + Yβ Nω (14.93)

The steady-state equations above can be used to analyze the effect of a particular type
of input to the system behavior, as demonstrated in the following examples.

Example 14.9
A vehicle traveling along a straight line path with speed V is subjected to a side force
(either banked road or wind force) Fy along the positive y direction. Figure 14.21 shows the
configuration for a banked road, where Fy = mg sin θ. Calculate the steady-state values of
the yaw velocity ω and sideslip angle β, as well as the radius of curvature.

G mgsin
y

Wo
z
mgcos  mg
Wi


FIGURE 14.21
Rear view of vehicle on a banked road.

Here, δ = 0 and M = 0. Substituting these values into Equation (14.91), steady-state


values of the sideslip angle and angular velocity become
βss Nω ωss Nβ
= − = [a]
Fy Q Fy Q

At steady-state, the radius of curvature can be obtained from the relationship V = RS ωss
as
V
RS = [b]
ωss
744 Applied Dynamics

The value of Q is always positive for an understeer vehicle. Q is also positive for an
oversteer vehicle when the speed is lower than the critical speed Vc . For an understeer
vehicle, Nβ = bCf − cCr > 0, so that the steady-state angular velocity ωss is positive for a
positive sideways force, and the vehicle develops a clockwise angular velocity. If the applied
side force is due to a banked road (right side of road lower than left side, so the lateral
component of gravity is in the positive y direction), an understeer vehicle will turn into the
incline (towards the downhill) andan oversteer vehicle will turn away from the incline.
Because Nω = V1 b2 Cf + c2 Cr is always negative (recall that cornering stiffnesses are
negative quantities), the steady-state value of the sideslip angle βss is positive when Fy is
in the positive y direction, so the sideslip angle is always in the direction of the side force.
Since Nω is inversely proportional to the vehicle speed, the sideslip angle will be smaller for
a faster moving vehicle, a phenomenon observed by drivers on banked roads. This occurs for
understeer vehicles only. Note that the sideslip angle is present whether there is oversteer,
understeer, or neutral steer.
For a neutral steer vehicle, Nβ = 0, so the vehicle will not develop an angular velocity,
but it will move along a straight line that is defined by the sideslip angle. An oversteer
vehicle on a bank will have a negative steady-state angular velocity ωss and a positive
sideslip angle βss . Figure 14.22 illustrates the motion of the vehicle.

G Oversteer vehicle
G
Fy
x
Lateral G

component Neutral steer y
of weight vehicle
Understeer vehicle

FIGURE 14.22
Behavior of vehicle on a banked road. Courtesy: Race Car Vehicle Dynamics, by Milliken
and Milliken.

Let us next consider the vehicle in Example 14.6 and calculate the angular velocity and
sideslip angle for a bank angle of 10◦ . The results are given in Table 14.10. Both the sideslip
angle and angular velocity have higher magnitudes for oversteer vehicles than for understeer
vehicles. When b/L = 0.6 the critical speed is Vc = 23.7 m/s. A speed of 22 m/s is quite
close to the critical speed and the values for the sideslip angle and angular velocity become
quite high. This is yet another indication of the dangers of driving an oversteer vehicle at
speeds close to the critical speed.
Let us next calculate the amount of sideways drift for a neutral steer vehicle on a banked
road of 5◦ . At a speed of 25 m/s, the sideslip angle becomes β = 0.009 radians. This implies
that the lateral speed of the vehicle is v = 25 × 0.009 = 0.225 m/s. In 10 seconds, the vehicle
will drift 2.25 m, almost the width of a traffic lane on a highway.
Maurice Olley, a legendary figure in 20th century racing, is credited with is the earliest
definition of understeer and oversteer. Olley defined under/oversteer in terms of the path
taken by a vehicle, moving along a straight line with zero steer angle, that is acted upon by
an external side force applied at the center of mass.
Vehicle Dynamics—Lateral Stability 745

TABLE 14.10
Steady-state values for sideslip angle β and angular velocity ω

b/L Speed (m/s) β (rad) ω (rad/s)


0.4 15 0.0132 0.0325
0.4 25 0.0088 0.0359
0.45 25 0.0116 0.0244
0.5 20 0.0178 0
0.55 25 0.0407 −0.0857
0.6 22 0.1351 −0.4866
0.6 15 0.0310 −0.0760

14.10 Yaw Velocity Gain and Curvature Response


We can get another perspective of instability due to oversteer by examining the steady-state
response as a function of the steer angle.
Setting Fy = 0 and M = 0 and using Equation (14.91), the ratio of the steady-state
angular velocity to the steer angle becomes
ωss 1
= (Nβ Yδ − Yβ Nδ ) (14.94)
δ Q
The algebra involved is quite lengthy (it is left as an exercise). Recall that
0
δdeg 0
Kdeg
δrad = Krad = (14.95)
57.296 57.296
Also, when both ω and δ are measured in degrees (or radians), ω/δ has the unit of
1/time, the yaw velocity gain has the form
ωss V /L ωss V /L
= 0 V 2 (in radians) = K0V 2
(in degrees) (14.96)
δ 1 + KgL δ 1 + 57.296gL
For an oversteer vehicle, K 0 < 0 and instability can be defined as the case where ωδss =
∞, so the smallest steering action will cause an infinite angular velocity and the vehicle will
spin out of control. This situation arises when the denominator of the above equation is set
to zero and solved for V = Vc , where Vc is the critical speed. Doing so gives
r
gL
Vc = 57.296 (in degrees) (14.97)
−K 0
which, of course, is the same expression as Equation (14.26).
Similar to the yaw velocity gain above, there are other quantities that characterize the
motion in terms of the steer angle. We studied one of these quantities, the lateral acceleration
gain, in Section 14.3.4. Other such quantities include
• Curvature response (or curvature gain). Defined by the ratio of the radius of curvature
to the steer angle, 1/R
δ
S
= R1S δ , the curvature response shows how the curvature changes
as a function of the steer angle. This ratio has the form (in degrees)
1 1 1/L
= K0V 2
(14.98)
RS δ 57.296 1 + 57.296gL
746 Applied Dynamics

Using radians and the stability factor K = K 0 /gL, we write the curvature response as
1 1/L
= (14.99)
RS δ 1 + KV 2

• Sideslip angle response (or sideslip gain). Defined by the ratio of the steady-state
value of sideslip angle to the steer angle, βδss , the slip angle response shows how the slip
angle changes as a function of the steer angle. This ratio has the form (in radians)
  c Wr V 2
βss c2 + cb Cf Cr /V + mV bCf L + Cr gL
= = K0V 2
(14.100)
δ Q 1+ gL

14.11 Tangent Speed and Hydroplaning


Previous sections studied critical speeds for understeer and oversteer. This section introduces
an additional critical speed, which has an interesting physical interpretation.
It was shown in Chapter 3 when studying vehicle kinematics that when a vehicle takes
a turn, the front tires track a larger circle than the rear tires. This characteristic changes as
slip angles are considered. The speed when the front and rear wheels track the same circle
takes place when the sideslip angle β = 0. The speed at which the sideslip angle β becomes
zero is known as the tangent speed. Above tangent speed, the rear wheels track a larger
radius than the front wheels.
To find the tangent speed, we take the steady-state equations in Equation (14.90), set
βss = 0, Fy = 0, and M = 0, which yields

− (Yω − mV ) ωss = Yδ δ − Nω ωss = Nδ δ (14.101)

Substituting the second equation into the first, denoting the tangent speed by Vt , and solving
for the tangent speed gives
Yδ Nω
Yω − = mVt (14.102)

After some manipulation, the tangent speed can be shown to be
r
−Cr cg
Vt = (14.103)
Wr
Recall that the cornering stiffnesses are negative. It is interesting to note that the tangent
speed is independent of the steer angle.
An interesting application of tangent speed is when considering hydroplaning, as illus-
trated in Figure 14.23. Hydroplaning occurs when there is standing water on the road (due
to flooding or heavy rain) and the tires of the vehicle are not able to push the water away
completely, so the vehicle ends up moving on a thin film of water. This results in loss of
traction and stability. Hydroplaning becomes worse with increased speed.
Experimental measurements show that hydroplaning speed, the speed beyond which
there remains a permanent film of water between the tire and the road, and hence potential
loss of stability, is primarily affected by tire pressure. The following relation is commonly
used for calculating hydroplaning speed:
√ √
VH = 9 p knots or VH = 10.35 p mph (14.104)
Vehicle Dynamics—Lateral Stability 747

!"#$ 

%&'$#()"*+

FIGURE 14.23
Hydroplaning of a tire. Notice the accumulation of water in front of the contact patch.

where VH is the hydroplaning speed and p is tire pressure in psi.


When a vehicle takes a turn with the tangent speed, the front and rear tires will travel
along the same path. In the case of standing water on the road, because the front tires will
have pushed away the water in the path of the vehicle, the rear tires will go through a path
that is relatively free of (or that has less) water and thus the chances of hydroplaning will
be reduced. It is of interest to calculate the tangent speed of a vehicle and to compare it
with the hydroplaning speed. In general, if the sideslip angle is small, say, less than 0.25◦ ,
the front and rear tires will track similar circles.

Example 14.10
Compare the tangent and hydroplaning speeds for a vehicle of weight 3000 lb, wheelbase 9
ft, distance from rear axle to wheelbase 5 ft, tire pressure 28 psi, cornering stiffness C = 120
lb/deg for each tire.
The hydroplaning speed is

VH = 10.35 28 = 54.77 mph = 54.77 × 88/60 = 80.33 ft/sec [a]

To find the tangent speed the rear axle load is calculated first as

W (L − c) 4
Wr = = 3000 × = 1333.3 lb [b]
L 9
Introducing this value into Equation (14.103), the tangent speed becomes
r
2 × 120 × 57.296 × 4 × 32.17
Vt = = 36.43 ft/sec [c]
1333.3
For this example, driving at tangent speed is desirable as the hydroplaning speed is
not exceeded and the driver reaps the added benefit of the rear tires moving against a
smaller amount of water. It should be noted that, while hydroplaning speed reflects an
unstable condition resulting from substantial reduction of friction between the tire and
road, hydroplaning occurs at speeds higher than tangent speed.

Example 14.11
Consider a vehicle with weight 3600 lb, wheelbase L = 9 ft, Wf = 2100 lb, and cornering
stiffness of C = 150 lb/deg on each tire. Find the tangent speed Vt and calculate the
748 Applied Dynamics

front and rear slip angles, as well as the sideslip angle, when the vehicle is moving with a)
v = Vt /2, b) v = Vt , c) v = 3Vt /2, d) v = 2Vt . The steer angle that is applied is δ = 1.2◦ .
The distance to the center of mass from the rear axle is
LWf 9 × 2100
c = = = 5.25 ft [a]
W 3600
Noting that the axle stiffness is Cr = −2C, the tangent speed is found from Equation
(14.103) as
r r
−Cr cg 2 × 150 × 57.296 × 5.25 × 32.17
Vt = = = 43.99 ft/sec [b]
Wr 1500

We will use the approach in Example 14.3 to find the slip angles and Equation (14.100)
to calculate the steady-state sideslip angle βss . The steady-state slip angles are found using
Equation (14.17)
Wf aGy Wr aGy
αf = αr = [c]
Cf g Cr g
and the steady-state lateral acceleration is given (in radians) in Equation (14.35) as (note
that ay = aGy )
aGy V 2 /L
= 0V 2 [d]
δ 1 + KgL
The sideslip angle is calculated using Equation (14.100). Table 14.11 gives the results
for the slip angles and sideslip angle.

TABLE 14.11
Sideslip and slip angles as a function of the tangent speed Vt = 43.99 ft/sec

Speed aGy /g αf (deg) αr (deg) βss (deg)


0.5Vt 0.033 −0.232 −0.165 0.496
Vt 0.114 −0.795 −0.568 0
1.5Vt 0.207 −1.446 −1.033 −0.574
2Vt 0.230 −2.028 −1.448 −1.086

14.12 Neutral Steer Point


We saw in Example 14.9 that an external lateral force applied at the center of mass of a
vehicle produces a sideslip angle and an angular velocity, with the direction of the angular
velocity depending on the understeer/oversteer characteristics. The question can then be
asked if there is a point on the vehicle such that an external lateral force produces no
angular velocity. This point is called the neutral steer point (NSP).
We find the location of the NSP by applying a lateral force F0 at a point that is a
distance d behind the center of mass, as shown in Figure 14.24. The resultant force and
moment are
Vehicle Dynamics—Lateral Stability 749

F0
A B
G
x
d v 
 y
c b

FIGURE 14.24
Neutral steer point.

Fy = F0 M = −F0 d (14.105)

Considering the case when the steer angle is zero, δ = 0, and substituting the above
values for the resultants into Equation (14.91), the steady-state angular velocity becomes

F0
ωss = (Nβ + Yβ d) (14.106)
Q
Setting ωss = 0 and solving for d gives the location of the neutral steer point as
Nβ bCf − cCr
d = − = − (14.107)
Yβ Cf + Cr

As expected, the neutral steer point location depends on whether the vehicle is an
understeer or oversteer vehicle. The stability derivative Yβ = (Cf + Cr ) < 0 all the time.
For an understeer vehicle, Nβ = bCf − cCr > 0, so the value of d is positive and the neutral
steer point is behind the center of mass. For an oversteer vehicle, the neutral steer point is
in front the center of mass.
The quotient of the distance d divided by the wheelbase L is referred to as the static
margin (SM). From Equation (14.107), the static margin has the value

d 1 Nβ
SM = = − (14.108)
L L Yβ

The static margin is yet another way of visualizing the understeer/oversteer character-
istics of a vehicle. Most vehicles are designed with a certain amount of understeer so the
neutral steer point is behind the center of mass and the static margin is positive.

Example 14.12
Consider the vehicle in Example 14.6 and calculate the static margin as a function of the
understeer properties.
Table 14.12 gives the results. Note that the static margin is a property of the vehicle and
is independent of speed. Also, because this example assumes that the cornering stiffnesses
are the same, the relationship between the static margin and center of mass location is
linear.
750 Applied Dynamics

TABLE 14.12
Static margin

b/L 0.3 0.4 0.5 0.6 0.7


SM = d/L 0.2 0.1 0 −0.1 −0.2

14.13 Modeling the Driver


In the transient and steady-state analyses conducted in the previous sections, the steering
input was considered to be a known function. In such a model, we can consider an instan-
taneous change in steering angle, such as a step input. In real life, a driver cannot apply
a steering input instantaneously. There also is the issue of how a driver reacts to a certain
traffic situation and how the driver adjusts steering. This section presents a simple model
of the driver and analyzes the effects of the driver on the response of the vehicle.
There are three main factors that dictate the steering input that a driver provides:
1. The swiftness (reaction time) with which a driver reacts to a need for change in steering.
One of the most important skills of race car drivers is that they react to situations they
encounter, such as taking a turn, passing another race car, braking and acceleration, or
avoiding an obstacle, much faster than ordinary drivers.
2. The criteria on which the driver bases a decision to change the steering input. For
example, when entering or leaving a turn, how much does the driver steer the vehicle?
The heading angle plays an important role, as well as the vehicle speed, when making
this decision.
3. The accuracy of the steering system. The steering geometry, as well as the alignment
and quality of components in the steering system, contribute to accuracy.
Steering systems are designed (for psychological purposes) so that there is a certain
amount of play in the steering wheel, allowing the driver to lightly manipulate the
steering wheel with no change in the steer angle.
The simple model for the driver described here takes into consideration driver reaction
time. Recall the discussion in Chapter 6 about of response of first-order systems and the
concept of time constant. Given a first order differential equation in the form
x
ẋ + = 0 (14.109)
τ0
where τ0 > 0 and the initial condition is x (0), the response can be shown to be a decaying
exponential of the form
t
x (t) = x (0) e− τ0 (14.110)

At time t = τ0 , the amplitude becomes

x (0)
x(τ0 ) = x (0) e−1 = = 0.3679x (0) (14.111)
e
The quantity τ0 is known as the time constant, which indicates the amount of time it
Vehicle Dynamics—Lateral Stability 751

takes for the amplitude to reduce to e−1 = 0.3679 times its initial value. At time t = 2τ0
the amplitude is down to x (2τ0 ) = e−2 x (0) = 0.1353x (0). As discussed in Chapter 6, the
time constant quantifies the decay rate of an exponential.
The reaction time of a driver can be modeled as a time constant. The time constant of
most drivers is between 0.2 and 1 second, with τ0 = 0.2 sec for race car drivers and τ0 = 1
sec for drivers who are slow to react. The time constant of most drivers begins to increase
after middle age.
The following simple model will be used to describe driver behavior:
1
δ̇ + δ = kd ψ + ke (14.112)
τ0
in which ψ is the heading angle, shown in Figure 14.13. The coefficient kd depends on a
variety of factors, such as driver experience and harshness, as in how abruptly the driver
takes turns, so it is a subjective quantity. The coefficient ke denotes the nature of the
steering. For example, the control action by the driver may be to track a desired heading
angle ψd , so that the right side of Equation (14.112) becomes kd (ψ − ψd ). In this case,
ke = −kd ψd .
Because Equation (14.112) involves the heading angle, it is necessary to include the
kinematic differential equation associated with it. The rate of change of the heading angle
is the yaw rate, so ψ̇ = ω. The result is two first-order differential equations that describe
the lateral motion, coupled to the equation for the driver, plus the kinematic differential
equation for the heading ψ, for a total of four first-order differential equations of the form

 
mV ω + β̇ = Yβ β + Yω ω + Yδ δ + Fy

Iψ ω̇ = Nβ β + Nω ω + Nδ δ + M

1
δ̇ + δ = kd ψ + ke ψ̇ = ω (14.113)
τ0

Defining the state vector {z} = [β ω δ ψ]T , we can write the stability equations in matrix
form as

[M 0 ]{ż} = [A0 ] {z} + {F 0 } (14.114)

in which
   
mV 0 0 0 Yβ Yω − mV Yδ 0
 0 Iψ 0 0  Nβ Nω Nδ 0 
[M 0 ] =  [A0 ] =  (14.115)
   
0 0 1 0 0 0 − τ10 −kd
 
  
0 0 0 1 0 1 0 0

and {F 0 } = [Fy M ke 0]T . The above equation is useful for studying the transient motion,
that is, for obtaining the time response to a steering or force input.
It should be noted that the above equations are based on the assumptions of small slip
angles and low lateral acceleration. A bad driver may make the vehicle develop dangerously
high levels of lateral acceleration. The mathematical model may become invalid and nonlin-
ear effects dominate. The tires may not be able to generate the needed lateral acceleration
forces, sliding may start, and the driver may lose control.
752 Applied Dynamics

14.14 Electronic Stability Control


One of the safety features introduced to vehicles in the latter part of the 20th century is
electronic stability control, or ESC. Based on the success of anti-lock brake systems (ABS)
and using equipment similar to that used in ABS, electronic stability control mechanisms
sense whether a certain wheel of the vehicle is beginning to slide and they take corrective
action by cutting power to the engine, disengaging the transmission, or by applying braking
action to certain wheels so that yawing moments are developed.

0(1*&'(-2%$1
)&$*+,-.

V )&$*+,-/
!"#$%&'(

FIGURE 14.25
Path of vehicle during evasive action.

Suppose all of a sudden the vehicle ahead slows down rapidly or you notice an obstacle
on the road. You do not have enough time or distance to stop or slow down. Your evasive
action will be to immediately steer your vehicle to the left (or right if that’s where there is
room to maneuver) and then restore the orientation by steering to the right. This two-step
action, which needs to take place in a very short period of time, can result in two types of
sliding motions. The path of the vehicle is shown in Figure 14.25.
Shown in Figure 14.26a for an evasive maneuver to the left, the first sliding action that
can occur is when the vehicle is sharply steered left. Here, if the front tires are turned too
rapidly,5 they lose grip and begin to slide. The velocity of the front axle is no longer in the
direction of the front wheels. Similar to what happens in an understeer vehicle, the actual
path of the vehicle is less curved than the desired path. The vehicle understeers the driver’s
action. If this situation persists, then the driver will not be able to evade the obstacle even
if the driver turns the steering wheel more.
ESC alleviates this situation by applying the brake of the inside (for a left turn, the
inside is the left side) rear tire, generating a force FB . This braking action has the effect of
providing a counterclockwise yawing moment (torque T = FB × t/2, where t is the track),
which aligns the vehicle path with the direction of the steer angle. Note that, because sliding
is in the front tires, the restoring moment is generated by applying the brake to the inside
rear wheel.
The second type of sliding that may occur is during the second step of the maneuver.
Illustrated in Figure 14.26b, the vehicle may turn more than what the driver intended it
to, oversteering the driver’s effort. Here, the front wheels do not slide but the rear tires
begin to slide out of the intended path of the vehicle. The ESC system counters this yawing
motion and provides a counterclockwise moment by applying the brake on the outside (left)
front wheel. Because the sliding is in the rear tires, the braking action generated by ESC is
applied to the front tires.
A similar (to the second type discussed above) sliding is encountered when a vehicle
5 Similar phenomenon takes place on an icy road for smaller steering maneuvers.
Vehicle Dynamics—Lateral Stability 753

Obstacle
Intended Actual
a) Intended Actual b) path
path
path path
x x

Sliding Sliding B y
B

 
y G T
G FB
T Resultant
Resultant Torque Sliding
Torque
A
A
t
t
FB Sliding

FIGURE 14.26
Evasive actions for vehicle. a) Sharp turn to left. Vehicle understeers and front wheels slide
away from the intended turn. b) Turn to the right. Vehicle oversteers and the rear wheels
slide out.

Intended
path
Actual x
path

B FB
y
Resultant T
Torque
G

Sliding
A

Sliding

FIGURE 14.27
Sliding out during a turn.
754 Applied Dynamics

enters a turn and the lateral forces needed to prevent sliding cannot be provided by the
tire friction, as shown in Figure 14.27. This type of sliding occurs when the vehicle speed is
high, even if the driver gently steers the vehicle. For a left turn with high speed, the rear
wheels slide to the outside, giving the vehicle a larger counterclockwise angular velocity
than needed, thus oversteering the driver’s effort. A clockwise torque is needed to align the
actual path with the intended path. This torque is generated by applying the brake of the
front outside (in Figure 14.27 front right) wheel.
An ESC system consists of sensors that measure vehicle velocity, steer angle, yaw ve-
locity, angular acceleration, and velocity of each wheel. An ESC system has the capability
of applying different brake forces to each wheel and cutting off power to the engine. By
calculating what the wheel velocities should be for the case of no sliding and by compar-
ing with the measured wheel velocities, ESC determines what type of evasive action to
take. This calculation is made by the ESC computer several times each second (e.g., 25
times per second). Companies that manufacture ESC systems keep the decision making
sequence and algorithm they use proprietary. There are several variants of ESC systems in
the marketplace.

Example 14.13
Velocity measurements taken on the vehicle in Examples 14.3 and 14.4 indicate that the
velocities at the axles are vA = 88i − 7.0j ft/sec and vB = 88i − 1.90j ft/sec. Is this vehicle
sliding? If so, calculate the brake force that needs to be applied by the ESC system. The
yaw moment of inertia is Iψ = 2, 400 slug·ft2 , the track is t = 6 ft, center of mass height
is h = 4 ft, and the brake force needs to be applied for half a second. It is assumed that
the ESC system can measure the steer angle of the vehicle and can calculate the slip angles
and angular velocity.
From Example 14.4, the angular velocity of the vehicle when there is no sliding is ω =
0.217 rad/sec cw, so ω = 0.217k rad/sec. Also, the velocity of the center of mass is vG =
88i − 2.89j ft/sec. The expected velocities of the axle centers A and B, in the absence of
sliding, can be found from the velocity of the center of mass as

vAe = vG + ω × rA/G = 88i − 2.89j + 0.217k × −5.5i = 88i − 4.084j ft/sec [a]

vBe = vG + ω × rB/G = 88i − 2.89j + 0.217k × 4.5i = 88i − 1.914j ft/sec [b]
The difference between the expected velocity of B and the measured value is very small,
so we can assume that there is no sliding of the front axle. However, the rear axle is definitely
sliding to the left and a corrective moment needs to be applied by the ESC system. We will
use the angular impulse-momentum theorem, given by Equation (5.3), and which states
Z ∆
Iψ ω0 + MG (t) dt = Iψ ωf [c]
0

where the time interval ∆ = 0.5 seconds is the duration of the corrective force, ω0 is the
initial (while sliding) angular velocity, and ωf = 0.217 rad/sec is the final (and desired)
angular velocity. The angular impulse is the integral of the external resultant moment MG .
In essence, we want ESC to change the angular velocity from its measured value to what it
should be in the absence of any sliding.
The angular velocity of the vehicle in the presence of sliding, ω0 , can be calculated from
the velocities of the axle centers using

vB = vA + ω 0 × rB/A [d]
Vehicle Dynamics—Lateral Stability 755

Noting that the wheelbase is L = 10 ft and ω 0 = ω0 k, we find the angular velocity as


vBy − vAy −1.914 + 7.0
ω0 = = = 0.509 rad/sec [e]
L 10


x

b
B y


G c
T
FB Resultant
Torque
A

Sliding

FIGURE 14.28
Free-body diagram of sliding vehicle.

The free-body diagram in Figure 14.28 indicates that, because the angular velocity is
positive (clockwise) and we need to reduce it to 0.217 rad/sec, a negative (counterclockwise)
moment needs to be applied. This moment is generated by applying the brake force at the
left front wheel, front because ESC applies the brake force to wheels that are not sliding.
Denoting this force by FB , the moment generated by the force is
   
t t
MG k = bi − j × −FB (cos δi − sin δj) = −FB b sin δ + cos δ k [f ]
2 2
Assuming that the brake force FB is constant, the angular impulse becomes MG ∆ and,
introducing this into the angular impulse-momentum expression in Equation [c], we obtain
MG ∆ = Iψ (ωf − ω0 ) [g]
Combining Equations [g] and [f], simplifying using small angles for the steer angle, and
solving for the brake force we obtain
Iψ (ω0 − ωf ) 2400 × (0.509 − 0.217)
FB =  =  = 845.8 lb [h]
∆ bδ + 2t 4.5×2
0.5 × 57.296 + 1.5
Can the front brakes generate this force? To answer this question, we need to calculate
the wheel loads for the front wheels. We will use the approach in Section 12.9 for this
purpose. The front axle load under static conditions is
c
Wfs = W = 0.55W = 1650 lb [i]
L
The braking force creates a deceleration of ax = −FB g/W , which shifts weight to the front
axle by
h W FB g h h 4
∆Wf = −max = = FB = 845.8 × = 338.32 lb [j]
L g W L L 10
756 Applied Dynamics

so that the load on each front wheel becomes Wfo = 1650/2 + 338.2/2 = 994.16 lb.
The lateral load shift from inside to outside tire can be calculated from Equation (12.86)
as
Fy h
∆Wo = [k]
t
in which the lateral force Fy = mf ay is due to the lateral acceleration, which was calculated
as ay = 0.591g in Example 14.3, and mf = Wfs /g denotes the mass of the front axle. Thus,

Wfs h 4
∆Wfo = ay = 0.591Wfs × = 650.1 lb [l]
g t 6
The wheel load on the front left (outside) wheel is Wfo = 994.2 + 650.1 = 1644.3 lb.
The ratio of the brake force to the wheel load defines the amount of friction that is needed
to exert the needed brake force
FB 845.8
µs = = = 0.514 [m]
Wfo 1644.3

which is a reasonable value on a dry road but a bit high for a wet road. Note that the lateral
acceleration is very high in this problem, so it may be preferable for the ESC to apply a
lower braking force over a longer time period.

14.15 Which Wheels Will Slide First?


With the exception of the previous section, which discusses electronic stability control,
all of the previous sections in this chapter considered no sliding of the wheels, permitting
calculation of lateral forces from the slip angles. This section discusses the answer to the
following question. Suppose you are driving with high speed. To negotiate a turn, you rotate
the steering wheel. If sliding of the vehicle occurs, which wheels will slide first? The front
or the rear?
Accident statistics indicate that the rear wheels will slide first. Let us demonstrate that
this is indeed the case. Consider the free-body diagram of the vehicle in Figure 14.11.
Ignoring the tractive force, the force balance in the y direction and sum or moments about
the center of mass become
X
+↓ Fy = maGy =⇒ FLr + FLf cos δ = may (14.116)

and the moment balance gives


X
 MG = Iψ ω̇ =⇒ −cFLr + bFLf cos δ = Iψ ω̇ (14.117)

Let us express the lateral forces in terms of the axle loads as


c b
FLf = µf Wf = µf W FLr = µr Wr = µr W (14.118)
L L
where µf and µr are effective friction coefficients. When the front wheels slide, µf = µk , and
for sliding of the rear wheels, µr = µk . Using relative acceleration expression, acceleration
of the front axle can be written as

aB = aG + α × rB/G + ω × ω × rB/G (14.119)
Vehicle Dynamics—Lateral Stability 757

where α = ω̇k and ω = 0 as the lateral motion has just started. Also, rB/G = bi. The lateral
acceleration of the center of mass is aGy = ay = V 2 /RS . The acceleration of the front axle
B becomes

aB = aB j = (ay + bω̇) j (14.120)

where, using the small angle assumption cos δ ≈ 1,


1  W
aGy = ay = FLf + FLr = (cµf + bµr )
m mL

1  W bc
ω̇ = −cFLr + bFLf = (µf − µr ) (14.121)
Iψ Iψ L
Next, express the acceleration of B in terms of the lateral forces as
W W b2 c
aB = ay + bω̇ = (cµf + bµr ) + (µf − µr ) (14.122)
mL Iψ L
For no sliding at B, the initial acceleration of B should be zero. Setting aB = 0 in the
above equation, expressing the mass moment of inertia for yaw by Iψ = IG = mκ2 , where
κ is the radius of gyration, dividing all terms by W/mL, and collecting terms give
b2 c b2 c
   
µf c + 2 + µr b − 2 = 0 (14.123)
κ κ
The above equation provides a relationship between the two friction coefficients as

µr = µf R (14.124)

where the ratio R is


2
µr c + b 2c
R = = b2 c κ (14.125)
µf κ2 − b

If a minimum friction coefficient of µf is needed to prevent sliding of the front tires, then
the friction coefficient needed to prevent sliding of the rear axle is µr = Rµf . It turns out
that R is greater than 1 for most configurations, so µr > µf . Higher amounts of friction are
needed at the rear tires than the front to maintain the no sliding condition. This implies that
the rear tires will slip before the front tires when taking a turn. The exception, as discussed
in the previous section, is when the steering wheel is turned rapidly and too much, as when
avoiding an obstacle.
For representative values of the ratio R, consider the commonly encountered values of
b/L = 0.45, c/L = 0.55 and κ/L = 0.35 (the usual range for the radius of gyration for yaw
is 0.3 < κ/L < 0.55). Using these values gives R = 3.178. For example, if the maximum
amount of friction available in the rear is µ = 0.9, then the front tires need a friction
coefficient of µ = 0.3 (or more) to maintain no sliding at the front wheels. Whether the
vehicle is an understeer or oversteer vehicle does not make much difference.

14.16 Bibliography
Genta, G., Motor Vehicle Dynamics: Modeling and Simulation, World Scientific, 1997.
758 Applied Dynamics

Gillespie, T.D., Fundamentals of Vehicle Dynamics, SAE Publications (R114), 1992.


Karnopp, D. Vehicle Dynamics, Stability, and Control, 2nd Edition, CRC Press, 2013.
Milliken, W.L., and Milliken, D.F., Race Car Vehicle Dynamics, SAE Publications (R146),
1995.
Jazar, R.N., Vehicle Dynamics: Theory and Application, 2nd Edition, Springer, 2013.
NSX Primer, http://www.nsxprime.com/FAQ/Track/highperfdriving.htm

14.17 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 14.2—Kinematics–Steer Angle Definitions


14.1 (E) Given a vehicle with wheelbase 2.8 m, steer angle 2.5◦ , and slip angles αf = 2.1◦
and αr = 1.5◦ , calculate the radius of curvature and determine whether the vehicle is
understeer or oversteer. Calculate the lateral acceleration when the speed is 78 km/h.
14.2 (E) A vehicle has a wheelbase of L = 9.2 ft, steer angle δ = 2.3◦ , and front slip angle
αf = 1.5◦ . Given that the radius of curvature is 500 ft, calculate the rear slip angle.

Section 14.3—Wheel Loads and Slip Angles


14.3 (E) What is the conversion factor between the units N/rad and lb/◦ ?
14.4 (E) Solve Example 14.2 using as cornering stiffnesses Cf = Cr = −260 lb/◦ .
14.5 (M) Obtain the lateral acceleration and lateral acceleration gain for the above problem.
Plot the lateral acceleration gain as a function of the vehicle speed for the range 0 to 100
mph. Then, switch the axle loads, so the vehicle becomes an understeer vehicle and plot
the lateral acceleration gain as a function of the vehicle speed.
14.6 (M) A vehicle has a weight of 20 kN and a wheelbase of 3.2 m. The weight distribution
on the front/rear axles is 53/47 under static conditions. The cornering stiffnesses of the front
tires is 39 kN/rad per tire, and of the rear 34 kN/rad. a) Calculate the stability factor. b)
Next, replace the front tires with tires with a cornering stiffness of 42 kN/rad, and calculate
the change in under/oversteer characteristics. c) Next, put the new tires (42 kN/rad) in the
rear, keep the old front tires (39 kN) and repeat the stability analysis.
14.7 (M) A vehicle with dimensions m = 1800 kg, L = 2.80 m, Wr = 850 kg, Cf = Cr =
−1100 N/deg (for each axle) is traveling at a speed of 80 km/h. Calculate the slip angles
and radius of curvature RS of the curve that the driver is taking when a) δ = 1.2◦ and b)
δ = 2◦ .
14.8 (M) The induced drag force due to slip in both axles can be modeled as Finduced =
FLr sin αr + FLf sin αf . Calculate the induced drag force for the vehicle in Problem 14.7
and the deceleration (in g) that the induced drag leads to for both values of the steer angle
δ.
14.9 (M) You are driving the vehicle in Problem 14.7 at a speed of 75 kph. You wish to
impress your passenger, who is holding an accelerometer, that you can generate any desired
lateral acceleration by properly turning the steering. What would your steer angle have to
be if the accelerometer records a lateral acceleration of 0.35g?
14.10 (M) A car of weight W and wheelbase L is to be designed as an oversteer vehicle with
a critical speed of Vc . This is to be done by adjusting the center of mass location. Given that
Vehicle Dynamics—Lateral Stability 759

the cornering stiffnesses at the axles are Cf and Cr , calculate the center of mass location
as a function of W , L, Vc , Cf , Cr , and the resulting wheel loads.
14.11 (M) A vehicle has a weight of 3500 lb and a wheelbase L = 9 ft. The center of mass is
at a distance 4.2 ft in front of the rear axle. The vehicle is moving with constant speed of 45
mph and the steer angle is δ = 2◦ . The cornering stiffnesses of the axles are Cf = Cr = −300
lb/◦ . Calculate the following quantities: a) the axle loads, b) the net steer effect and lateral
acceleration, c) the understeer gradient and critical speed.
14.12 (M) A vehicle has the following properties: wheelbase L = 3.5 m, front axle load
Wf = 75, 000 N, rear axle load Wr = 55, 000 N. The cornering stiffnesses for the tires are
given in Table 14.13. The vehicle is traveling at a speed of 40 km/hr and wishes to take a
turn of RN S = 100 m. Calculate the steering and slip angles and the understeer gradient.

TABLE 14.13
Cornering stiffness as a function of tire load for Problem 14.12

Tire Load 10,000 20,000 30,000 40,000 50,000 60,000 N


Cα 1100 2100 3000 3900 4700 5500 N/deg

m M x

y
L!"#
x

FIGURE 14.29
Figure for Problem 14.13.

14.13 (M) A commonly used contraption in experimental wheel load analysis is a beam of
mass m and length L, over which a point mass M can slide, as shown in Figure 14.29.
By adjusting the location of the mass M , we can vary the center of mass location and
wheel loads, as well as the mass moment of inertia. Let m = 1300 kg and M = 400 kg and
calculate the change in the wheel loads and in the understeer gradient, as x, which denotes
the location of mass M , is varied in the range −L/6 < x < 5L/12. Both axles have the
same cornering stiffness of C = −1600 N/deg.

Section 14.4—Slip Angle Kinematics


14.14 (M) The bicycle model of a vehicle has a wheelbase L = 3 m and the center of
mass lies 1.6 m from the rear. The velocities of points A and B are vA = 22i − 0.62j m/s,
vB = 22i + 0.38j m/s, and the steer angle is δ = 2◦ . Calculate a) the slip angles and sideslip
angle and b) the net steer effect and lateral acceleration. Assume that the derivative of the
sideslip angle is zero, β̇ = 0.
14.15 (M) The bicycle model of a vehicle has a wheelbase L = 3 ft and the center of mass
lies 1.6 m from the rear. The velocities of points A and B are vA = 22i − 0.62j m/s and
vB = 22i m/s and the steer angle is δ = 2◦ . Calculate a) the slip angles and sideslip angle
and b) the net steer effect and lateral acceleration. You are given that β̇ = 0.05 rad/s.
760 Applied Dynamics

14.16 (M) Consider the bicycle model of a vehicle (c = 4.7 ft, b = 3.9 ft, m = 120 slugs,
Iψ = 2500 slug·ft2 ). Both wheels have a cornering stiffness of Cf = Cr = −200 lb/deg. At a
given instant ω = 0.3 rad/sec, β = −2◦ , β̇ = −0.05 rad/sec, V = 100 ft/sec, δ = 0.4◦ ; find
the slip angles and radius of curvature at this instant. Note that this is not a steady-state
problem.
14.17 (M) Consider the bicycle model of a vehicle with wheelbase L = 9 ft, weight W = 3000
lb, Wr = 1200 lb. It is observed during steady-state motion that the center of mass velocity
is v = vG = 45i−0.1522j ft/sec and the angular velocity is ω = 0.1277 rad/sec cw. The steer
angle is δ = 2◦ . Calculate the slip angles. Then, find the lateral acceleration of the center
of mass. Then, calculate the lateral forces that are generated and the cornering stiffnesses
that are needed to allow these lateral forces.
14.18 (M) The bicycle model of a vehicle has been programmed for the transient motion
and the following results are printed out at a certain time step: ω = 0.3 rad/sec, β = -0.12
rad, V = 60 ft/sec, δ = 0.3◦ . The vehicle has the following properties: b = 4.1 ft, c = 3.9 ft,
mass m =110 slugs. a) Find both slip angles and the lateral acceleration. b) Do the same
analysis for V = 120 ft/sec and compare results. c) Comment on your answers, evaluating
the friction that would be necessary to maintain this type of lateral acceleration.

Section 14.6—Response
14.19 (C) Consider the bicycle model of a vehicle (c = 4.7 ft, b = 3.9 ft, m = 120 slugs,
Iψ = 2500 slug·ft2 ). Both axles have a cornering stiffness of Cf = Cr = −200 lb/deg.
Simulate the transient motion equations and plot the response of the angular velocity and
sideslip angle as a function of time for the case when the vehicle is traveling with constant
speed of 60 mph and a steering input of δ = 0.9◦ is applied for one second. Then, do the
same maneuver at a speed of 40 mph and compare the results. Conduct the simulation for
5 seconds.
14.20 (C) Consider the previous problem and now apply a second steering angle of δ = −0.9◦
at time t = 4 seconds for a period of one second. Carry on the simulation for 10 seconds. Do
you expect the steady-state values for the angular velocity and sideslip angle to be zero at
t = 8 seconds? Plot the magnitudes of the lateral forces and the effective friction coefficient
for each lateral force.
14.21 (C) Consider the previous problem and change the vehicle parameters to c = 3.9 ft,
b = 4.7 ft, making the vehicle an oversteer vehicle. Carry out the same maneuver (with a
steer input at t = 0 and one at t = 4 seconds of equal and opposite magnitudes of 0.9◦ )
as the previous problem and compare the response, in the form of angular velocity, sideslip
angle, lateral force magnitudes, and effective friction coefficients.
14.22 (C) Consider Problem 14.20. Now, the maneuver is being conducted on a banked
road, where the right wheels are lower than the left wheels. Compare the angular velocity
and sideslip angles for the following bank angles: 0◦ , 2◦ , 4◦ .

Sections 14.7 and 14.8—Eigenvalue Analysis


14.23 (C) Consider the bicycle model of a vehicle (c = 1.4 m, b = 1.2 m, m = 1650 kg,
Iψ = 2600 kg·m2 ). Both axles have a cornering stiffness of Cf = Cr = −1350 N/deg. Plot
the eigenvalues of this vehicle as a function of the vehicle speed, considering a speed range
of 10 to 120 kph.
14.24 (C) Consider the bicycle model of a vehicle (c = 1.2 m, b = 1.4 m, m = 1650 kg,
Iψ = 2200 kg·m2 ). Both axles have a cornering stiffness of Cf = Cr = −1350 N/deg. Plot
the eigenvalues of this vehicle as a function of the vehicle speed, considering a speed range
of 10 to120 kph. Plot the eigenvalues even if the critical speed is exceeded.
14.25 (C) Consider the two previous problems and calculate and plot the damping constants
Vehicle Dynamics—Lateral Stability 761

and natural frequencies for the ranges of speeds considered. Compare these values for the
understeer and oversteer cases.
14.26 (C) Consider the contraption in Problem 14.13 used to model a vehicle and plot the
eigenvalues as a function of the location of the mass M when the vehicle speed in 92 km/h
and when the vehicle speed is 60 km/h.

Section 14.9—Steady-State Response


14.27 (M) Consider the bicycle model of a vehicle (b = 1.35 m, c = 1.15 m, W = 17,000
N, Iψ = 1600 kg·m2 ). The cornering stiffnesses are Cf = −114, 000 N/rad, Cr = −126, 000
N/rad. a) Calculate the understeer coefficient and critical speed. b) The car is riding at a
speed of V = 90 kph when it encounters a bank on the road of 10◦ . Calculate the steady-state
values of the yaw rate and radius of curvature.
14.28 (M) Consider the bicycle model of a vehicle (b = 1.35 m, c = 1.15 m, W = 17,000
N, Iψ = 1600 kg·m2 ). The cornering stiffnesses are Cf = −114, 000 N/rad, Cr = −126, 000
N/rad. The car is riding at a speed of V = 90 kph on a level road, when the driver turns
the steering to 0.9◦ . Calculate the steady-state values of the yaw rate and sideslip angle.
14.29 (C) Consider a vehicle with mass of 1650 kg, yaw moment of inertia Iψ = 3100 kg·m2 ,
wheelbase 2.75 m, c = 0.56L, and cornering stiffness of 1320 N/deg on each axle. The center
of mass is at a height of h = 57 cm. The vehicle is being driven at a speed of 80 km/h and
is taking a turn with a steer angle of δ = 1.1◦ . The road surface is not flat and makes an
angle of θ with the horizontal. Calculate and plot the steady-state response of the angular
velocity and sideslip angle for the range of the uphill of −10◦ < θ < 10◦ .

Section 14.10—Yaw Velocity Gain and Curvature Response


14.30 (E) Consider the transient motion equations. A lateral force Fy is applied to the body.
The steer angle is zero. Ignore the yaw motion by setting the angular velocity ω and its
derivative equal to zero. Show that the time constant of the resultant first-order equation
is −mV /(Cf + Cr ). Calculate the time constant for a vehicle of mass 2000 kg traveling at
a speed of 72 kph, whose axle cornering stiffnesses are −69, 000 N/rad each.
14.31 (C) Consider a vehicle with mass of 1680 kg, yaw moment of inertia Iψ = 3200
kg·m2 , wheelbase 2.75 m, c = 0.56L, and axle cornering stiffnesses of −76, 000 N/rad in the
front and −74, 000 N/rad in the rear. Calculate and plot the yaw velocity gain, curvature
response, and sideslip angle response of a speed range of 0–135 kph.
14.32 (M) Consider a vehicle with mass of 1650 kg, yaw moment of inertia Iψ = 3200 kg·m2 ,
wheelbase 2.75 m, c = 0.56L, and cornering stiffness of C = −1320 N/deg on each axle.
The vehicle is entering a skidpad. What is the maximum speed the vehicle can attain to
track a circle of radius of 50 m, given that the maximum value of the lateral force that can
be obtained is at a slip angle of 6.1◦ . If this slip angle is exceeded, the lateral force will still
have a magnitude of 6.1C.
14.33 (M) Consider the bicycle model of a vehicle (b = 1.25 m, c = 1.35 m, W = 17,000 N,
Iψ = 2300 kg·m2 ). The axle cornering stiffnesses are Cf = −114, 000 N/rad, Cr = −126, 000
N/rad. a) The car is riding at a speed of V = 90 kph on a level road, when it encounters
a bank on the road of 5◦ . Calculate the steady-state values of the yaw rate and radius of
curvature. b) The car is riding at a speed of V = 90 kph on a level road, when the driver
turns the steering to 0.7◦ . Calculate the steady-state values of the yaw rate and sideslip
angle.
762 Applied Dynamics

Section 14.12—Neutral Steer Point


14.34 (M) Consider the bicycle model of a vehicle (b = 1.4 m, c = 1.3 m, m = 1600 kg,
Iψ = 2900 kg·m2 ). Both axles have a cornering stiffness of Cf = Cr = −1000 N/deg.
Calculate the critical speed, understeer coefficient, and static margin.

Section 14.13—Modeling the Driver


14.35 (C) Consider the vehicle in Problem 14.34. Numerically integrate and plot the response
of the vehicle for time constants of τ0 = 0.2, 0.4, 0.6 seconds and kd = 0. The vehicle
is traveling at 45 mph, when the driver executes a maneuver of turning the steering by
ke = 0.25 for 0.7 seconds and then by ke = −0.25 for 0.7 seconds.
14.36 (C) Consider Problem 14.35. Here, the driver wants to travel along a straight line,
but has strayed from a straight line by 12◦ . Given that ke = 0, kd = −0.8, simulate the
response and pick a maneuver length (based on τ0 ) so that the driver will align the vehicle
with the road (φ will become zero). Do this by trial and error.
14.37 (M) The vehicle in Problem 14.34 is traveling on a banked road of 5◦ . What steering
input does the driver have to give so that the vehicle will track a straight line? The steering
counters the effect of the side force generated by the gravity force on the banked road.
15
Vehicle Dynamics—Bounce, Pitch, and Roll

15.1 Introduction



 
  

 

FIGURE 15.1
Coordinate system for vehicle.

The previous chapter analyzed the lateral motion of vehicles (coordinate system shown
in Figure 15.1), namely translation in the y direction (sideslip) and rotation about the z axis
(yaw), as well as lateral tire forces and yaw moments. Chapter 12 considered translational
motion in the x direction, acceleration and braking, tractive forces, and a simple model of
roll. Tire modeling was discussed in Chapter 13. This chapter considers the remaining three
directions: translation in the vertical direction, referred to as bounce, and rotation about
the x and y axes, referred to as roll and pitch, respectively. These three motions describe
the ride properties of a vehicle.
The study of ride properties requires modeling of the suspension system and its compo-
nents. In general, suspensions are complex mechanisms. This chapter discusses simplified
models of suspensions and analyzes the behavior of vehicles in the presence of restoring
and dissipative forces generated by the suspension. Such analysis involves concepts from
vibration theory.1 Suspension geometry and the roll center, as well as jacking, scrub, and
dive and squat motions, round out the rest of the chapter.
We need to consider three main factors to understand the ride properties of a vehicle:
the sources of excitation, the response to the excitation, and how vibrations of a vehicle are
perceived and tolerated by the vehicle’s occupants.
1 Vibration concepts are discussed in Chapters 6 and 7.

763
764 Applied Dynamics

15.2 Sources of Excitation


The bounce, pitch, and roll motions, which get transmitted to the body of the vehicle
through the suspension system, arise from a number of sources. These sources can be cate-
gorized as a) external to the vehicle and b) internal, generated by components of the vehicle.
Among sources of excitation external to the vehicle are
• Body forces, such as aerodynamic and gravity. Aerodynamic forces may act from
several directions.
• Road roughness, incline, and bank. Roads that are not built well, roads that wear
out, and roads that have been repaired in patches do not have uniform surfaces and excite
vibrations in vehicles that travel over them. Road undulations are usually modeled as
periodic functions. As a result of the incline or bank of a road, the force of gravity will
have components in the lateral or tractive directions.
• Tire forces. Acceleration, braking, lateral forces, and normal forces act on tires. The
resultants of these forces create moments, as discussed in Chapter 13.
Excitations generated internally, that is, by the components of the vehicle, are usually
periodic in nature and they include
• Wheel imbalances and tire vibrations. A tire and wheel assembly usually has
imbalances due to a nonuniform mass distribution. This may be due to manufacturing
defects, eccentricities, nonuniform material properties of the tires, aging and uneven
wear, or impact, such as the tire and wheel hitting a curb and becoming deformed.
Another source of imbalance is worn out bearings. As the vehicle moves and the tires
roll, imbalances lead to periodic excitations. In addition, tires have flexibility themselves
and may start vibrating, depending on road conditions and vehicle speed.
• Driveline excitation. The power generated by the engine is transmitted to the tires
by transmission shafts, universal joints, and differential. These components may have
imperfections due to manufacturing defects, accidents, aging, asymmetry, eccentricities,
or larger clearances than the initial design, as well as elasticity. Shafts with permanent
deformation may no longer be straight. All of these factors create periodic excitations. In
addition, joints that connect shafts are not perfect constant velocity joints (see Chapter
9), resulting in fluctuations in the transmitted power.
• Engine rotation and torque. Engines rotate and they generate torques that are
transmitted to the differential and drive shaft. To reduce excitation due to these effects,
the engine is attached to the vehicle body by engine mounts. As the mounts wear out,
the periodic excitation caused by the engine rotation is felt more and more by the vehicle
and the occupants.
Another way of categorizing excitations is based on the duration of the excitation
• Impulsive forces are applied over very short time periods, a sudden wind gust or
entering a pothole, for example. We use concepts from impulsive motion to analyze
such motions.
• Longer duration forces include braking, cornering, or inclined roads, and they are
applied over longer time periods. The response is best investigated by transient analysis,
using time-domain techniques. Sometimes, we ignore transient effects and look at the
steady-state behavior (as we did in the previous chapter for lateral stability).
Vehicle Dynamics—Bounce, Pitch, and Roll 765

• Persistent excitations continue for long durations of time, and they are usually peri-
odic in nature. The effects of periodic excitations are best analyzed by frequency-domain
techniques. Road imperfections are usually treated as random variables.
In general, excitation frequencies in the range of 0 to 25 Hz affect ride properties.
Higher frequency excitations (25 to 20,000 Hz) affect the noise properties, and they
may also manifest themselves in the form of localized vibrations. Recall that 1 Hz =
1 cycle/second. In many cases, the periodic excitation on a vehicle causes both lower
and higher frequency vibrations and a separate analysis becomes difficult. The study
of NVH (noise, vibration, and harshness) is a significant component of vehicle analysis
and design.
Response of dynamical systems that are subjected to excitations are studied in Chapter 6
for single-degree-of-freedom systems and in Chapter 7 for multi-degrees-of-freedom systems.

15.2.1 Ride Quality and Human Response to Vibration


The way ride is perceived varies for different people, and it is a subjective issue. What
contributes to ride quality or ride comfort depends on
• How vehicle motions, translational as well as rotational, are transferred to the driver and
occupants. Accelerations and changes in accelerations affect ride quality (and harshness)
more than velocity and changes in position.
• Vehicle comfort, including factors such as interior design, seats, ergonomics, tempera-
ture, ventilation, windows, noise and lighting.
The interactions among the different factors that contribute to ride quality are not well
established. Also, humans prefer different ride qualities, such as a soft suspension vs. a stiff
suspension or tight handling vs. slower response. Vehicles are designed to emphasize one
ride quality at the expense of another. For example, sports cars are designed to maximize
performance and handling. From research on human response to vehicle motions, it is known
that:
• The effect of change in acceleration is more pronounced at lower frequencies.
• Humans respond differently to vibrations acting on them from different directions.
• Ride amplitude and frequency of a vehicle both contribute to ride quality.
Research has also shown that a pitch, bounce, or roll frequency of the vehicle in the
range of fn = ω2πn = 1 to 1.5 Hz is perceived as a comfortable ride. Most people think a
ride is harsh if any of these three frequencies approaches 2 Hz. High-performance cars have
frequencies that exceed 2 Hz. Table 15.1 lists a summary of human response to different
frequencies of excitation.
The way vibration is perceived depends also on the duration of the vibration, as well as
on the acceleration amplitudes. Researchers have developed threshold curves for amplitudes
at which vibration is detected by humans. For example, in the frequency range of 1 to 10
Hz, the detection threshold is about 0.001g.
The most important factor in determining the ride quality of a vehicle is the lowest nat-
ural frequency of the vehicle. The suspension system should isolate and damp out impulsive
forces, such as those occurring when a bump or pothole is encountered, as well as wind
gusts. The suspension system should also isolate and damp out excitations due to the road
surface roughness and excitations due to rotating components in the vehicle. The vehicle
should have a sufficient amount of noise reduction capabilities.
766 Applied Dynamics

TABLE 15.1
Human response to vibration at different frequencies

Frequency (Hz) Type of Discomfort or Affected Body Parts


0.5 to 0.8 Motion sickness
3 to 10 Internal organs, abdomen, arm, chest, shoulder
> 10 Spinal column
10 to 25 Visual performance
18 to 20 Neck and head, head usually after 20 Hz
30 to 50 Hands
> 80 Localized vibrations, parts in contact with vibration source
60 to 90 Eyeballs
100 to 220 Jaw

15.3 Unsprung vs. Sprung Mass


The connection between the vehicle body and the ground is achieved through the tires and
the suspension system. Tires provide contact with the road, and the suspension system,
which consists of springs, dampers, and suspension arms, provides the connection between
the vehicle body and tires.
The vehicle body usually vibrates with higher amplitudes than the tires. You can test
this by going near a stationary car (whose owner you know well) and push the car down
with two hands. You’ll notice that the body of the car will move up and down, while the
tires remain almost stationary. Even though tires have their own flexibility, tires are much
stiffer than the suspension.
It is convenient to separate the vehicle components into
• The unsprung mass, which consists of the tires, brakes, and axles (some also include a
portion of the weight of the suspension system), and
• The sprung mass, which consists of the vehicle body, engine, transmission, and suspen-
sion system.
The overall motion can be viewed as the motion of the sprung mass over the unsprung mass.
The complexity of the model depends on how many different motions are considered and
the how unsprung mass is treated. A schematic of unsprung and sprung masses is given in
Figure 15.2. The unsprung mass is usually around 13 to 15% of the total vehicle mass.

15.4 Simple Suspension Models


There are several ways of simplifying the mathematical model of a vehicle. Suspension
modeling is no exception. This chapter begins with a simple suspension model and then
considers more sophisticated systems. All the models here assume that the vehicle body,
that is, the sprung mass, is rigid.
The connection between the vehicle and the road is through two elements that have
flexibility: the tires and the suspension. The degree of complexity of a suspension model
Vehicle Dynamics—Bounce, Pitch, and Roll 767







 





FIGURE 15.2
Unsprung and sprung mass.

depends on how these elements are treated and the resulting numbers of degrees of freedom.
The following distinctions are usually made:

• Treatment of the tire mass and stiffnesses. Simpler models ignore the tire mass.
This simplification reduces the number of degrees of freedom. Such models either ignore
tire stiffness or combine it together with the suspension stiffness. The combined stiffness
is modeled as an equivalent spring constant and is referred to as ride rate. More complex
models treat the tires and suspension system separately.
• Treatment of the different motions of the vehicle. Simplifications are made by
considering certain motions separately. For example, the previous chapter modeled the
sideslip and yaw motions separately from the other motions. Complex models of suspen-
sions consider the pitch, roll, and yaw motions together. Simpler models consider the
roll separately from the pitch and bounce, and the simplest suspension models consider
each corner of the vehicle separately.

Following is a classification of suspension models.

• Full car model, which takes into consideration the entire vehicle, as shown in Figure
15.3. The vehicle is modeled as a rigid box that has bounce, pitch, and roll motions.
This model has three degrees of freedom when the suspension is modeled by the ride
rates of each wheel. The model has seven degrees of freedom when the suspension and
tire stiffnesses are modeled separately; four for the tires and three for the sprung mass.

• Half-car model for bounce and pitch. This model considers the side view of the
vehicle and has two degrees of freedom (bounce and pitch) when the suspension on each
axle is modeled by the spring constants kf and kr (ride rates), with the subscripts f
and r denoting front and rear, respectively, as shown in Figure 15.4a. If the tires are
modeled as separate bodies, as in Figure 15.4b, the model has four degrees of freedom.

• Half-car model for roll. This model, depicted in Figure 15.5, is a rear view of the
vehicle (rear axle) with each suspension component (ride rate) modeled by kS . While
this model resembles the two-degrees-of-freedom half-car model for bounce and pitch, it
768 Applied Dynamics

Tire

y Tire

Tire 
x Tire
z

FIGURE 15.3
Full car model.

a) b) 

z
 mS
x
z kr kf
z m zr mU zf mU

kr kf kt kt

FIGURE 15.4
Half-car model for bounce and pitch. a) Two-degrees-of-freedom model, b) four-degrees-of-
freedom model.

turns out that the equations of motion are uncoupled, and we can treat the roll equation
of motion separately from the bounce motion.
• Quarter-car model for bounce. This model only considers translational motion in
the vertical direction (bounce motion only). When the suspension and tire stiffnesses
are modeled as separate, the quarter-car model has two degrees of freedom, as shown in
Figure 15.6. Here, kt denotes the spring constant of the tire and kS the spring constant
associated with the suspension. The damping constant of the suspension is denoted by
cS . Because the tire is much stiffer than the suspension, damping properties of the tire
are usually ignored in this model.
The quarter-car model can be further simplified, as depicted in Figure 15.7, by using
an equivalent spring keq and combining the effects of the suspension and tire stiffness,
resulting in a one-degree-of-freedom model.

An advantage of using a single or two-degrees-of-freedom model is that both lead to


closed-form solutions or solutions that can be obtained by hand. Such analysis gives tremen-
dous insight. A model with more than two degrees of freedom is usually treated numerically.
Vehicle Dynamics—Bounce, Pitch, and Roll 769


mS y

kS kS

FIGURE 15.5
Half-car model for roll (as viewed from the rear of the vehicle).

mS

z kS cS

mU

zU kt

zR

FIGURE 15.6
Two-degrees-of-freedom quarter-car model.

mS mS

z kS =
cS cS keq
kt

zR

FIGURE 15.7
One-degree-of-freedom quarter-car model.
770 Applied Dynamics

When a suspension system is traveling down, or compressing, its motion is called jounce
or compression. When a suspension system is traveling up, or expanding, the motion is
called rebound or bounce, as illustrated in Figure 15.8.

,%-()*.
!"#$%&"
'()*+%

FIGURE 15.8
Jounce (compression) and rebound (bounce) of a suspension system.

15.5 Quarter-Car Model


The quarter-car model considers only vertical motion and analyzes each corner of the ve-
hicle separately. While this model is simple, it is useful in analyzing static deflections and
frequency ratios, and comparing sprung mass motion with the unsprung mass motion am-
plitudes. Moreover, it provides a simple tool for predicting response characteristics.
The quarter-car model represents a two-degrees-of-freedom system when the tires and
suspension are modeled separately. When the motion of the unsprung mass is ignored, the
tire and suspension springs become springs in series (see Section 4.11.2). The model reduces
to the single-degree-of-freedom system in Figure 15.7, where the equivalent spring constant
is
kt kS
keq = (15.1)
kt + kS
and we usually consider the total mass as the sprung mass, m = mS . From prior knowledge
of springs in series in Chapter 4, the equivalent spring (ride rate) is weaker than the stiffness
of each spring, so keq < kt and keq < kS . The tire stiffness is much higher than the suspension
stiffness, kS << kt , with the ratio kkSt = rS = 5 to 10. We can express the ride rate in terms
of the stiffness ratio as
kt kS rS kS
keq = = (15.2)
kt + kS (1 + rS )
For example, rS = 9 gives the result keq = 0.9kS . Because of this, many times when
conducting a single-degree-of-freedom analysis, we may ignore the tire stiffness.

15.5.1 Single-Degree-of-Freedom Model


The single-degree-of-freedom model ignores the unsprung mass and uses the ride rate keq as
the spring constant. The mass mS is one fourth the sprung mass of the vehicle or it is taken
Vehicle Dynamics—Bounce, Pitch, and Roll 771

a) b)

F F + mS g

mS mS

z
keq cS . .
keq (z–zR) cS (z–zR)

zR

FIGURE 15.9
a) Single-degree-of-freedom quarter-car model (with damping) and b) free-body diagram.

as half the axle load. The free-body diagram is given in Figure 15.9b. The displacement z
of the sprung mass is measured from the undeformed position of the spring.
In suspension systems, the spring is installed as preloaded, that is, already in compres-
sion. The spring always stays compressed to some extent, even when the suspension system
expands it. From Chapter 4, such a spring is referred to as a compression spring. This
property does not affect any of the results. Summing forces in the vertical direction gives
X
+↓ F = mS z̈ = −keq (z − zR ) − cS (ż − żR ) + mS g + F (15.3)

in which zR is the road surface amplitude, and F is the applied external load. Collecting
all of the motion variable terms on the left side, the equation of motion becomes

mS z̈ + cS ż + keq z = keq zR + cS żR + mS g + F (15.4)

The first two terms on the right side of the above equation represent the excitation due to
the road surface and the next term is the influence of gravity.
Consider the static deflection. When the model is at rest and the only force acting on it
is gravity, the vehicle is at static equilibrium. The static deflection, denoted by zst , can be
found by setting all derivatives and external force F in Equation (15.4) to zero

mS g WS
zst = = (15.5)
keq keq

The amount of static deflection is an important design issue, and it is closely related
to the natural frequency of the model. To see this better, divide the equation of motion in
Equationq(15.4) by the sprung mass mS . From Chapter 6, the undamped natural frequency
keq
is ωn = mS and the damping factor is

cS cS
ζ = = √ (15.6)
2mS ωn 2 mS k

Dividing Equation (15.4) by mS , we can rewrite the equation of motion as

F
z̈ + 2ζωn ż + ωn2 z = ωn2 zR + 2ζωn żR + +g (15.7)
mS
772 Applied Dynamics

It follows that the expression for the static equilibrium becomes


WS g
zst = = 2 (15.8)
keq ωn

The above relationship is an important design consideration. The static deflection and
natural frequency have to be considered and designed together. Design of the natural fre-
quency is based on the perception of a comfortable ride. For a passenger car, a frequency
fn = ω2πn in the range of 1 Hz is usually viewed as a comfortable ride. The static deflection
that corresponds to fn = 1 Hz is
g g g
zst = = = = 0.8149 ft = 9.778 in. (15.9)
ωn2 4π 2 fn2 4π 2

While a quarter-car frequency of fn ≈ 1 Hz is desirable, other design considerations


come to mind, such as softness/stiffness of ride, as well as the amplitudes of the position,
velocity, and acceleration of the sprung mass. A softer suspension results in higher position
amplitudes but lower acceleration amplitudes. As the suspension is made stiffer, position
amplitudes go down, but acceleration amplitudes go up. To see this, consider the undamped
model and no external excitations and write z̈ + ωn2 z = 0, or z̈ = −ωn2 z. The acceleration is
proportional to the square of the natural frequency.
Chapter 6 discusses techniques for obtaining the response of single-degree-of-freedom
systems. The analysis is summarized here. Consider the free motion, that is, no external
excitations on the quarter-car, so that F = 0, as well as a level road, so that zR = 0.
After calculating the static equilibrium position, replace z with z + zst in Equation (15.7),
which means that we are now measuring the vertical deflection z from the static equilibrium
position. The equation of motion becomes

z̈ + 2ζωn ż + ωn2 (z + zst ) = g (15.10)

Noting from Equation (15.8) that g = zst ωn2 , the gravity term cancels from both sides
of the above equation, leading to a simpler form of the equation of motion as

z̈ + 2ζωn ż + ωn2 z = 0 (15.11)

To solve the above equation, consider a solution in the form z (t) = Ceλt and introduce
it into the above equation and collect terms, with the result

λ2 + 2ζωn λ + ωn2 Ceλt = 0



(15.12)

Because Ceλt cannot be zero (otherwise there would be no nonzero solution), the only
way for the above equation to be satisfied is if the coefficient of Ceλt vanishes, thus

λ2 + 2ζωn λ + ωn2 = 0 (15.13)

The above equation is known as the characteristic equation and is a polynomial of order
two in λ. The roots of the polynomial are
 p 
λ1,2 = ωn −ζ ± i 1 − ζ 2 (15.14)

From Chapter 6, for underdamped systems, that is, when the damping factor ζ < 1, the
response has the form

z (t) = Ad e−ζωn t cos (ωd t − φd ) (15.15)


Vehicle Dynamics—Bounce, Pitch, and Roll 773

where the amplitude Ad and phase anglep φd depend on the initial conditions and ωd is the
damped natural frequency, ωd = ωn 1 − ζ 2 . The response is in the form of a sinusoidal
enveloped inside a decaying exponential. The damping factor ζ dictates the decay rate. The
period of oscillation is denoted by Td and has the value

Td = (15.16)
ωd
A discussion is presented in Chapter 6 about what damping factors are preferred for
suspension systems. Figure 6.10 shows the time response as a factor of the damping factor.
Once the damping factor exceeds 0.5, the motion is almost aperiodic, as the motion ampli-
tude becomes very small in less than a vibration cycle. While it is desirable to damp out
vibrations, designing a suspension system with a large damping factor does not result in a
comfortable ride. It also requires a powerful shock absorber.
Race cars have damping factors going as high as 0.6, with passenger cars having damping
factors of about 0.25 or higher. Also, as will be discussed later, modern dampers have two
damping factors, one for jounce and another for rebound, where the damping factor for
rebound is higher than the damping factor for jounce, sometimes by a factor of two.

Example 15.1
This example investigates the amplitude drop between two peaks of a linear viscously
damped system. It is assumed that a vibrating system is at rest (as in a vehicle mov-
ing on a level road) when an impulsive or similar type of force is applied to it (as in entering
a pothole) and the amplitude of the ensuing free response is analyzed.
The amplitude ratio between two peaks (one full cycle) is given in Equation (6.85) as

z (tj+1 ) − √2πζ
= e 1−ζ2 j = 0, 1, 2, . . . [a]
z (tj )

in which the two peaks are one period apart and z(tj+1 ) = z(tj ) + Td .
Let the first peak occur at time, t = t0 , so that time t = t1 = t0 + Td denotes the time of
the second peak. Equation [a] shows that the amplitude ratio is independent of the natural
frequency. The amplitude ratio is plotted in Figure 6.11. For ζ = 0.2 the amplitude ratio is
0.28 and for ζ = 0.5 the amplitude ratio becomes 0.027, which indeed is small.
Let us next compare amplitude ratios after half a cycle, that is, at time t1/2 = t0 + Td /2.
From Chapter 6, the response amplitudes are

z (t0 ) = Ad e−ζωn t0 cos (ωd t0 − φd ) z t1/2 = Ad e−ζωn t1/2 cos ωd t1/2 − φd


 
[b]

Recalling that the period of oscillation is Td = 2π/ωd , after half a cycle we have

cos ωd t1/2 − φd = cos (ωd t0 + π − φd ) = − cos (ωd t0 − φd ) [c]

Introducing these values to the response expression, the amplitude ratio for a half cycle
becomes
Ad e−ζωn t1/2 cos ωd t1/2 − φd
 
z t1/2 e−ζωn t1/2 − √ πζ
= −ζω t
= −ζω t
= −e−ζωn Td /2 = −e 1−ζ2 [d]
z (t0 ) Ad e n 0 cos (ωd t0 − φd ) −e n 0

The amplitude ratio after half a cycle is 0.53 when ζ = 0.2 and 0.16 when ζ = 0.5.
Hence, for a race car with damping of ζ = 0.5, the effect of a disturbance is dissipated by
84% after half a cycle.
774 Applied Dynamics

15.5.2 Change in Natural Frequency and Damping Factor Due to Pay-


load Weight
The weight of a vehicle changes with the number of occupants and with the cargo that the
vehicle carries. This change affects the natural frequency and damping factor, as well as
the static deflection. For example, in an extreme case, if the weight is doubled, the natural
frequency and damping factors become
r
0 k cS cS
ωn = = 0.707ωn ζ0 = 0
= √ = 0.707ζ (15.17)
2mS 2mS ωn 2 2kmS

which is quite a change. Furthermore, the static deflection, zst = mS g/k, becomes twice its
value. The previous chapter discussed problems associated with overloading vehicles and
the reduction in lateral stability. Overloading a vehicle also adversely affects the suspension
system and fuel efficiency.
The added payload changes the ride qualities in two additional ways. One is the change
in the sprung mass vs. unsprung mass ratio. The second is in the changes that take place in
the suspension geometry, in the form of changes in the orientation of the suspension links
and roll center height.

Example 15.2
Consider a sprung mass of 450 kg and a static deflection of 20 cm. Calculate a) the spring
rate and b) the change in natural frequency of the quarter-car when the mass is increased
to 600 kg due to the addition of passengers or payload.
The natural frequency, when the mass is 450 kg, is
r r
g 9.807
ωn = = = 7.002 rad/s = 1.114 Hz [a]
zst 0.2

and the spring rate is

keq = mS ωn2 = 450 × 7.0022 = 22, 066 N/m [b]

The natural frequency after the mass is increased to m0 = 600 kg becomes


r r
0 keq 22, 066
ωn = = = 6.064 rad/s [c]
m0 600
and the new static deflection is
0 g 9.807
zst = 2 = = 0.267 m = 26.7 cm [d]
(ωn0 ) 6.0642

As expected, the static deflection is higher when the vehicle is overloaded.

15.5.3 Wheel Hop


Another single-degree-of-freedom analysis that can be conducted on the quarter-car is to
look at the wheel bounce (hop) separately from the sprung mass. This analysis and the pre-
vious analysis of a single ride rate is possible because the spring constants for the suspension
and the tire are quite separate from each other.
Look at a moving vehicle and focus on the vertical motion as the vehicle goes over
a bump or a pothole. The motion induced by the bump or pothole primarily affects the
Vehicle Dynamics—Bounce, Pitch, and Roll 775

tire, and the motion amplitude of the vehicle body is less than the tire. Because the spring
constants of the suspension and tire are quite separate from each other, only a small portion
of the motion is transmitted from the tire to the vehicle body.
Figure 15.10 shows the wheel hop model, which considers the motion of the unsprung
mass by treating the sprung mass as nonmoving.

!" #"

kS mU g kS zU

mU mU

zU kt zU
kt

FIGURE 15.10
a) Wheel hop model and b) free-body diagram.

The effects of the two springs can be viewed as springs in parallel, with an equivalent
spring constant for wheel hop kH = kS + kt . Going back to the spring constant ratio
rS = kt /kS , we can write kH = kt 1 + r1S . For rS = 9, kH = 1.11kS .
The natural frequency of the wheel hop can be calculated as
r
kt + kS
ωH = (15.18)
mU

15.5.4 Two-Degrees-of-Freedom Model


We now consider the two-degrees-of-freedom quarter-car model and obtain the equations of
motion. The model and free-body diagrams are shown in Figure 15.11. Both z and zU are
measured from the undeformed positions of the springs. For completeness, we also include
an external force to the sprung mass and one to the unsprung mass, denoted by FS and FU ,
respectively. These forces represent other excitations that the two masses can experience.
Also added to the model are the effects of an uneven road, with zR denoting the road
unevenness. The force balances for the sprung mass and for the unsprung mass become
X
+↓ F = mS z̈ = −cS (ż − żU ) − kS (z − zU ) + FS + mS g

X
+↓ F = mU z̈U = cS (ż − żU ) + kS (z − zU ) − kt (zU − zR ) + FU + mU g (15.19)

As there are no constraint forces to be eliminated, the above equations are the equations
of motion. Rearranging terms, we can write the equations of motion in standard form
(developed in Chapter 7) as
mS z̈ + cS ż − cS żU + kS z − kS zU = FS + mS g

mS z̈U − cS ż + cS żU − kS z + (kS + kt ) zU = kt zR + FU + mU g (15.20)


776 Applied Dynamics

a) b) c)
FS

mS mS
. .
FU kS (z–zU) cS (z–zU)
kS cS
kS (z–zU) . .
z cS (z–zU)
mU FS+mS g mU
zU kt
mU g+FU kt (zU –zR)
zR

FIGURE 15.11
a) Two-degrees-of-freedom quarter-car model. b) Free-body diagram for sprung mass. c)
Free-body diagram for unsprung mass.

Introducing the position vector {z} = [z zU ]T and the matrices

cS −cS −kS
     
mS 0 kS
[m] = [c] = [k] = (15.21)
0 mU −cS cS −kS kS + kt

we can write the equations of motion in matrix form as

[m] {z̈} + [c] {ż} + [k] {z} = {F } (15.22)

where
 
mS g + FS
{F } = (15.23)
mU g + FU + kt zR

is the external excitation vector. All matrices are symmetric, the mass and stiffness matrices
are positive definite, and the damping matrix is positive semidefinite.2
The static case is of interest. Setting all time derivatives and external forces to zero, and
considering a level road surface, the static equilibrium equations become
 
mS g
[k] {zst } = (15.24)
mU g

whose solution is
   
mS g g kS (mS + mU ) + kt mS
{zst } = [k]−1 = (15.25)
mU g kS kt kS (mS + mU )

Note that, when we measure z and zU from the static equilibrium position, the terms
mS g and mU g disappear from the equations of motion.
The next step is to consider the eigenvalue problem and to calculate the natural fre-
quencies via modal analysis. Consider the free response of the undamped model, set the
excitation to zero {F } = {0}, and measure deflections from the static equilibrium position.
2 See Chapter 7 for the definition of positive semidefinite.
Vehicle Dynamics—Bounce, Pitch, and Roll 777

For synchronous motion discussed in Chapter 7, where the characteristics of the response of
the sprung and unsprung masses have the same time dependance, the response is expressed
as {z (t)} = {u} eλt , which when introduced into the equations of motion leads to

λ2 [m] + [k] {u} eλt = {0}



(15.26)

The above equation can have a nontrivial solution only when the coefficient matrix is
singular and its determinant vanishes. Noting from Chapter 7 that the eigenvalues of this
problem are pure imaginary, introducing the notation ω 2 = −λ2 , where ω denotes a natural
frequency, the characteristic equation becomes

det −ω 2 [m] + [k] = 0



(15.27)

Solution of Equation (15.27) yields the two natural frequencies of the quarter-car model.
Both natural frequencies are greater than zero. For convenience, we write them in ascending
order, ω1 < ω2 . It turns out that the first natural frequency, ω1 , is close
p to the natural
frequency of the one-degree-of-freedom model for the unsprung mass, keq /mS , and the
second natural frequency, ω2 , is close to ωH , the wheel hop frequency.
Because the suspension and tire stiffnesses are quite apart from each other, the natural
frequencies obtained from the two-degrees-of-freedom model are also quite separate from
each other. This can be demonstrated by obtaining closed-form expressions for the natural
frequencies as a function of the ratio rS of the two spring rates.
The modal vectors (eigenvectors) of the quarter-car model closely resemble the motions
of the sprung and unsprung mass. We write the modal vectors as {u1 } and {u2 }, where
       
u11 1 u12 −2
{u1 } = = {u2 } = = (15.28)
u21 1 u22 1

where 1 and 2 are small. This property separates the vibrations of the two masses from
each other and is known as ride isolation. Ride isolation is a desirable property.
Let us return to the discussion on wheel hop and the vehicle entering a pothole (or going
over a bump). The tire itself moves up and down much more than the vehicle body. This is
what is wanted from the suspension design: road disturbances are absorbed by the tire and
suspension system and not felt as much by the sprung mass (the occupants and cargo).
When damping is included in the model, the characteristic equation becomes

det λ2 [m] + λ [c] + [k] = 0



(15.29)

The solution turns out to be a pair of complex roots with negative real parts. The real
parts of the roots indicate the amount of damping, while the complex parts are the damped
natural frequencies. The reader is referred to the discussion on damped multi-degrees-of-
freedom systems in Chapter 7 and to the approximate solution in Section 7.12.

Example 15.3
A quarter-car model has the following properties: mS = 400 kg, mU = 40 kg, kS = 17 N/mm,
kt = 180 N/mm. Calculate the ride rate and the natural frequencies associated with the
one-degree-of-freedom model for the sprung mass and for the wheel hop. Compare these
frequencies with the natural frequencies of the two-degrees-of-freedom model. Calculate the
modal vectors and comment on the ride isolation concept.
The ride rate is
kS kt 17 × 180
keq = = = 15.532 N/mm [a]
kS + kt 17 + 180
778 Applied Dynamics

so the frequency of the single-degree-of-freedom model and of the wheel hop model are (note
that the unit of stiffnesses is converted to N/m from N/mm in the equation below)
r r
1 keq 1 15, 532
fn = = = 0.9918 Hz [b]
2π mS 2π 400
r r
1 kS + kt 1 (17 + 180) × 1000
fH = = = 11.170 Hz [c]
2π mU 2π 40
Next, consider the actual two-degrees-of-freedom system and associated eigensolution.
Defining the position vector as {z} = [z zU ]T , the mass and stiffness matrices become
   
400 0 17 −17
[m] = kg [k] = N/mm [d]
0 40 −17 197

Solving the eigenvalue problem, the frequencies are calculated as f1 = 0.9914 and f2 =
11.173 Hz. These frequencies are remarkably close to their estimates given in Equations [b]
and [c]. The modal vectors associated with the two frequencies can be shown to be
   
1 −0.0087
{u1 } = {u2 } = [e]
0.088 1

When looking at the first modal vector, which describes motion amplitudes when the
quarter-car is oscillating with the first frequency, amplitude of the sprung mass (u11 = 1)
is far larger than the amplitude of the unsprung mass (u21 = 1 = 0.0870).
The reverse is true for the second mode. The second frequency closely resembles the wheel
hop frequency, and the associated modal vector has the amplitude of the unsprung mass
(u22 = 1) much higher than the amplitude of the sprung mass (u12 = −2 = −0.0087). This
almost total separation between the wheel hop and ride frequencies and modes illustrates
the ride isolation property.

15.6 Pitch and Bounce Motions


This section considers the side view of a vehicle and uses a two-degrees-of-freedom math-
ematical model shown in Figure 15.4 to describe the interaction between the translational
(in z direction, or bounce) and rotational (about y axis, or pitch) motions of the vehicle.
We ignore tire flexibility and use the ride rates of the front and rear suspensions.

15.6.1 Equations of Motion


The pitch-bounce model has two degrees of freedom. Two sets of motion variables come
to mind: i) translation of the center of mass zG and rotation about the center of mass θ
(pitch angle), or ii) translations of the wheels zA and zB . The variables zG and θ are more
descriptive, as θ is the magnitude of the pitch angle.
Figure 15.12 illustrates the geometry. Initially, both variables are measured from the
undeformed positions of the springs. Note that the vehicle may not be horizontal at static
equilibrium. Unevenness of the road is included in the model via the amplitudes of the road
surface undulations, denoted by zRf and zRr , for the front and rear.
The free-body diagram is shown in Figure 15.13 for the undamped case. For complete-
ness, a resultant external force F that is acting at the center of mass and a resultant moment
Vehicle Dynamics—Bounce, Pitch, and Roll 779
v
A G B
Initial position
c b
zG – sin
zG
zG + csin s
After translation by zG

Final position after


sr rotation by about G
Initial road surface
zR
f
zR New road surface
r

FIGURE 15.12
Deflection of suspension system of vehicle for pitch and bounce.

M is also included in the free-body diagram. The ride rates (spring constants) for the front
and rear are denoted by kf and kr , respectively. Note that kf and kr are axle ride rates and
their values are twice the spring constants of the individual suspensions.

x
Lc
Initial c b
position A B z
G
F
zG
After rotation
After translation

M kf ( zG – bsin – zR )
f
mS g

kr ( zG + csin – zR )
r

FIGURE 15.13
Free-body diagram for pitch and bounce.

Let us denote the distance between the undeformed position and the road surface by s
and the distance from the final position of the rear spring from the (uneven) road surface
by sr . From the geometry, we can write

s + zRr = zG + c sin θ + sr (15.30)

Denote the deflection of the rear spring by ∆r and observe from Figure 15.13 that
∆r = s − sr . It follows that the deflections of the rear and front springs can be expressed as

∆r = zG + c sin θ − zRr ∆f = zG − b sin θ − zRf (15.31)


780 Applied Dynamics

The pitch angle usually is small, which justifies a small angles approximation of sin θ ≈ θ,
cos θ ≈ 1. Summing forces in the vertical direction and moments about the center of mass
yields
X 
+↓ F = mS z̈G = mS g − kr (zG + cθ − zRr ) − kf zG − bθ − zRf + F (15.32)

X 
MG = Iθ θ̈ = −ckr (zG + cθ − zRr ) + bkf zG − bθ − zRf + M (15.33)

where Iθ is the centroidal mass moment of inertia of the vehicle about the y (pitch) axis,
or pitch moment of inertia. The above equations can be put into standard form by keeping
all motion variable terms on the left and all other terms on the right:

mS z̈G + (kf + kr ) zG + (ckr − bkf ) θ = mS g + F + kf zRf + kr zRr

Iθ θ̈ + (ckr − bkf ) zG + c2 kr + b2 kf θ = M + ckr zRr − bkf zRf



(15.34)

Introducing the vector {z} = [zG θ]T , the equations of motion can be written in matrix
form as

[m] {z̈} + [k] {z} = {F } (15.35)

in which
ckr − bkf
   
mS 0 kf + kr
[m] = [k] = (15.36)
0 Iθ ckr − bkf c kr + b2 kf
2

are the mass and stiffness matrices and


 
mS g + F + kf zRf + kr zRr
{F } = (15.37)
M + ckr zRr − bkf zRf

is the external force vector. The stiffness matrix [k] is not diagonal, so the two equations of
motion are coupled to each other and cannot be solved independent of each other. We can
show that both matrices [m] and [k] are positive definite.

15.6.2 Analysis of the Equations of Motion and Response


Consider the static case. Setting all time derivatives to zero, eliminating all external forces,
and assuming a level road, the static equilibrium equation can be obtained as
 
mS g
[k]{zst } = (15.38)
0

whose solution is
c2 kr + b2 kf
     
mS g zGst mS g
{zst } = [k]−1 = = (15.39)
0 θst bkf − ckr D

where D is the determinant of the stiffness matrix and has the form
2
D = det [k] = (b + c) kr kf = L2 kr kf (15.40)
Vehicle Dynamics—Bounce, Pitch, and Roll 781

Note that at equilibrium, the pitch angle θst may not be zero. Selection of the static
deflections, the vertical deformation as well as pitch angle, is a design issue.
Next, consider the free motion and conduct an eigenvalue analysis. Measuring the motion
variables zG and θ from their static equilibrium position eliminates gravity terms. When
all external excitations are zero and the road is level, zRf = zRr = 0, force vector becomes
{F } = {0}.
Considering synchronous motion, the free motion is expressed as {z (t)} = {Z} eλt .
Introducing this into the equations of motion leads to the conclusion that in order to obtain
a nontrivial solution, the relationship
λ2 [m] + [k] {Z} eλt = {0}

(15.41)
must hold. Hence, a nontrivial equation can exist only if the determinant of the coefficient
matrix vanishes. Noting from Chapter 7 that the eigenvalues of this problem are pure
imaginary, introduction of the notation ω 2 = −λ2 enables us to write the characteristic
equation as
det −ω 2 [m] + [k] = 0

(15.42)
Solution of the above equation yields the natural frequencies of the bounce-pitch model
as ω1 , ω2 . The corresponding eigenvectors (or modal vectors) can be found by substituting
the values of ω12 and ω22 into Equation (15.41). The associated modal vectors are denoted
by {Z1 } and {Z2 }. When the vehicle is vibrating with the i-th (i = 1, 2) natural frequency,
the associated modal vector (eigenvector) is
 
ZGi
{Zi } = (15.43)
Θi
where ZGi denotes the amplitude of the center of mass and Θi is the pitch angle.
Because the equations of motion are coupled, the natural motions, as described by the
natural frequencies and modal vectors, are not pure bounce or pitch motions. Nevertheless,
for most vehicles (that is, for vehicles where the front and rear ride rates are not vastly
different and the center of mass is relatively close to the midpoint of the wheelbase), the
natural modes somewhat resemble pitch and bounce motions.
We can see this more clearly by looking at the modal vectors and the amplitude ratio
ZGi /Θi . This ratio, which has the unit of length, defines the center of oscillation for the
i-th mode of vibration. The i-th mode rotates about the i-th center of oscillation.
Figure 15.14 describes construction of the center of oscillation. Begin with line AB,
which depicts the undeformed position (or the static equilibrium) position of the vehicle.
First translate the line AB by a distance ZGi in the vertical direction, with the resulting
line denoted by A0 B 0 . Then, rotate line A0 B 0 by angle Θi about the center of mass G and
denote the resulting line by A00 B 00 . In this illustration, both ZGi and Θi are positive. The
point where lines AB and A00 B 00 (or their extensions) intersect is the center of oscillation
Ci for the i-th mode. Denoting the distance between the center of mass G and center of
oscillation Ci by di , we can write
ZGi
tan Θi = (15.44)
di
For small angles, di can be expressed as
ZGi ZGi
di = ≈ (15.45)
tan Θi Θi
and the position of the center of oscillation is thus found. There are two possibilities:
782 Applied Dynamics

Initial position B'' x


Ci v
A B z
ZG G
i i
A' B'
G'
di

A'' Final position

After translation by ZG
i

FIGURE 15.14
Center of oscillation for i-th mode.

• When di > 0, the center of oscillation is in front of the center of mass,


• When di < 0, the center of oscillation lies behind the center of mass.
In Chapter 7, we saw that the modal vector associated with the first (or lowest) natural
frequency has no sign changes in its elements. This observation holds true when all motion
variables have the same dimensions, such as when all are displacements or all are angles.
In the problem here, which involves translational as well as rotational coordinates, this rule
is not valid. That is, we cannot know which modal vector has a sign change, or whether
ZGi /Θi > 0 or ZGi /Θi < 0, until the modal vectors are calculated.
The oscillation center for either mode can lie to the front of the center of mass. However,
because there is a sign change in one of the modal vectors, one of the oscillation centers lies
in front of the center of mass, while the other is located behind the center of mass.
When the center of oscillation (COO) of a mode is close to the center of mass or it
is inside the wheelbase, the motion associated with that mode resembles pitching, with
points A and B (the suspension springs) moving in opposite directions. By contrast, when
the center of oscillation is away from the center of mass, and especially when outside the
wheelbase, the motion associated with that mode resembles bouncing motion, and points
A and B move in the same direction.
Figure 15.15a depicts the cases when the COO is close the center of mass, resulting in
motion that resembles pitch, while in Figure 15.15b the COO is far away from the center
of mass, resulting in motion that resembles bounce. In both figures, line AB denotes the
initial position, while A0 B 0 and A00 B 00 denote two of the positions that the sprung mass can
assume.
The distances of the centers of oscillation from the center of mass are used as criteria
when designing suspensions, while keeping in mind the following:
• Experiments with human response to vehicle vibrations indicate that pitch motion cre-
ates more discomfort than bounce, and

• The motion amplitude of the first mode is higher than the motion of the second mode.
The lower modes of a vibrating system have smaller frequencies, and they contribute
more to the motion amplitudes than the higher modes.
Vehicle Dynamics—Bounce, Pitch, and Roll 783

($

#$ %$

%$ #$
($$ "
# % #
( ș (
&' &' %
($ !
%$$ #$$

#$$ %$$

($$

FIGURE 15.15
Center of oscillation location: a) close to center of mass, motion resembles pitch, b) away
from center of mass, motion resembles bounce.

The vehicle should be designed so that the oscillation center associated with the first mode
is farther away from the center of mass than the oscillation center associated with the second
mode. This way, the first mode will resemble bouncing motion and the second mode, which
has a higher frequency and lower amplitude, will resemble pitching.
Is it preferable to have decoupled equations of motion? The equations of motion become
decoupled when ckr − bkf = 0. This case is easier to analyze from a mathematical point
of view. However, the resulting ride is of poor quality. Because the two motions of pitch
and bounce are completely uncoupled, they each die out at times proportional to their
initial amplitudes and an uncomfortable ride ensues. There is no transfer of pitch motion to
bounce motion or vice versa. Complex motions, where pitch and bounce motions alternate,
are perceived as more comfortable than only pitch or only bounce motions.
When analyzing pitch and bounce motions, an interesting dimensionless quantity to
consider is the dynamic index, DI. Writing the pitch moment of inertia as Iθ = mS κ2 ,
where κ is the radius of gyration, the dynamic index is defined as

κ2
DI = (15.46)
bc
We can show that when DI = 1, the centers of oscillation are located at the axles.

Example 15.4
Consider the vehicle with the following parameters: Wf = 1000 lb, Wr = 850 lb, kf = 130
lb/in., kr = 95 lb/in., L = 110 in., DI = 1.04. Find the frequencies and modal vectors of
this vehicle. Then calculate the oscillation centers and identify the predominantly pitch and
bounce modes.
From the weight distribution we can calculate the location of the center of mass as

Wr L 850 × 110/12
b = = = 4.212 ft c = L − b = 4.955 ft [a]
Wr + Wf 1850
784 Applied Dynamics

From Equation (15.46), the radius of gyration becomes


√ √
κ = DI × b × c = 1.04 × 4.212 × 4.955 = 4.659 ft [b]

The mass and stiffness matrices then are


   
1 W 0 57.51 0
[m] = = [c]
32.17 0 W κ2 0 1248

ckr − bkf −921.6


   
kf + kr 2700
[k] = = [d]
ckr − bkf c k r + b2 k f
2
−921.6 55, 661
Note that because the displacement vector has both displacement and angular variables,
elements of the mass and stiffness matrices do not have the same units. In this problem, zG
has units of feet and θ has units of radians. Hence, for example, m11 = 57.5 slugs, while
m22 = 1, 247 slug·ft2 .
Solving the eigenvalue problem yields the natural frequencies and modal vectors as
 
0.9577
ω1 = 6.494 rad/s {Z1 } =
0.2876
 
0.9884
ω2 = 7.033 rad/s {Z2 } = [e]
−0.1516

{Z}

{Z1}

{Z2}

FIGURE 15.16
Plot of modal vectors.

The modal vectors are normalized so that their magnitude is unity, or {ui }T {ui } =
1. The modal vectors are plotted in Figure 15.16. The frequencies and oscillation center
locations are
ω1 ZG1 0.9577
f1 = = 1.037 Hz d1 = = = 3.330 ft
2π Θ1 0.2876

ω2 ZG2 0.9884
f2 = = 1.120 Hz d2 = = = −6.520 ft [f ]
2π Θ2 −0.1516

The oscillation center associated with the first frequency is closer to the center of mass,
hence the motion with the first mode resembles pitching more than bouncing. The second
oscillation center is 6.520 ft behind the center of mass. Considering that the distance between
the rear wheel and center of mass is c = 4.955 ft, the second oscillation center lies behind
the rear wheel. Hence, oscillation with the second frequency resembles bouncing more than
pitching. Figure 15.17 shows the locations of the center of mass and the oscillation centers.
Vehicle Dynamics—Bounce, Pitch, and Roll 785

*+,-+./!
*+,-+./"
# $ !! " &
!"
#$%&% #$"!%
&$'"% ($()%

FIGURE 15.17
Oscillation centers.

Example 15.5
The spring constants in the previous example were deliberately selected so that the first
oscillation center would be closer to the center of mass than the second. The first mode
resembles pitching which leads to a less comfortable ride. Let us now consider the same
vehicle and the same mass and center of mass location, but with the following different
spring rates and dynamic index:

kr = 130 lb/in. kf = 110 lb/in. DI = 0.9 [g]

The frequencies are obtained as f1 = 1.059 Hz and f2 = 1.263 Hz, and the modal vectors
become    
1 1
{Z1 } = {Z2 } = [h]
−0.1555 0.3425
where we normalized the modal vectors so that the top element has a value of 1. Recall from
Chapter 7 that how we normalize or scale the modal vectors is only a matter of convenience.
Using the modal vectors, the oscillation centers are calculated as
1 1
d1 = = −6.431 ft d2 = = 2.920 ft [i]
−0.1555 0.3425
The centers of oscillation are plotted in Figure 15.18. The first mode resembles bouncing
and the second mode resembles pitching, which is what is preferred for a comfortable ride.

Center 2
Center 1
A G C2 B v
C1
4.96' 4.21'
6.43' 2.92'

FIGURE 15.18
Oscillation centers for second case.

15.7 Olley Criteria


Maurice Olley (1889–1972), a legendary figure in vehicle dynamics and racing circles, who
was mentioned earlier in this text, developed a set of criteria that defines a comfortable
786 Applied Dynamics

ride. Even though the Olley criteria were developed over 75 years ago, they are still widely
used. The criteria are
1. The front suspension should have a spring constant kf that is 30% lower than the rear
suspension, kr . Note that the requirement is on the spring constants and not on the
natural frequencies.3 Some modern vehicles do not use this criterion and faster cars
routinely deviate from it. Table 15.2 lists some vehicles whose spring rate selection is
different from the Olley criterion. Note that for every vehicle mentioned in the list there
are several variations of the suspension design.

TABLE 15.2
Some vehicles with spring rates different from the First Olley criterion

Model and Year Front Rate (lb/in.) Rear Rate (lb/in.)


Acura TSX 2008 280 173
Subaru Impreza 2001 157 149
Porsche GT3 2008 200 370
BMW M3 2005 158 354

2. The natural frequencies for the pitch and bounce model should be close to each other.
The natural frequencies should have a ratio of ω
ω1 < 1.2, with ω2 being associated with
2

the mode that resembles pitch motion.


3. The frequencies associated with pitch and bounce should not be higher than 1.3 Hz.
Satisfying this criteria results in a static deflection that is over 6 in.

4. The natural frequency of the roll motion, ωφ (which will be studied later in this chapter),
should be similar to the natural frequencies of pitch and bounce.

15.8 Response to Harmonic Excitation


As discussed in Section 15.2, a significant portion of the forces that act on vehicles are
periodic in nature. The engine, drive shafts, differential, bearings, and tires all rotate at
high speeds. Road surface imperfections, except for potholes, can also be modeled as having
some random periodicity. The vehicle and its components, such as the engine, shafts, and
tires, may not be well balanced. All these effects lead to some form of periodic excitation.
In this section, we make use of the frequency response techniques described in Chapters 6
and 7 to analyze the response of vehicles.

15.8.1 Modeling an Uneven Road


A simplified model of an uneven road is shown in Figure 15.19. In general, an uneven road
can be modeled as a summation of sinusoids and by using concepts from random variables
to describe the amplitudes and periods of the road imperfections. Consider an uneven road
3 For passenger cars it is preferable to have the quarter-car frequencies higher in the rear than in the

front, which occurs when this Olley criterion is followed.


Vehicle Dynamics—Bounce, Pitch, and Roll 787

A
x
zR

FIGURE 15.19
Uneven road modeled as a periodic function.

that has the sinusoidal shape zR (x) = ±A sin 2πxL , where A is the amplitude of the road
unevenness and L is the length of one cycle of unevenness (wavelength). Assume that the
vehicle is moving with constant speed V , so we can write x = V t. Substitution of this
expression into the positive value of the road imperfection equation permits description of
the road imperfection as a function of time as
2πV
zR (t) = A sin t = A sin ωt (15.47)
L
The frequency of excitation, ω = 2πV /L, is dependent on the speed of the vehicle.
For the quarter-car model, we can use this excitation by substituting it directly into
the equations of motion. For the pitch and bounce model, there is a lag in the excitation
received by the front and rear tires. Denoting the location of the front tire by x, the location
of the rear tire is x − L, where L is the wheelbase. Hence, the excitations that the front and
rear tires receive are
2πx 2π (x − L)
zRf (x) = A sin zRr (x) = A sin (15.48)
L L
Introducing the relationships x = V t and ω = 2πV /L for constant speed, the excitations
to the front and rear become

zRf (t) = A sin ωt zRr (t) = A sin (ωt − Ψ) (15.49)

where Ψ = 2πL L denotes the lag angle of the excitation, which is dependent on the ratio of
the wheelbase to the road imperfection period.
The ratio of the wheelbase to the road imperfection wavelength plays a role in the
L
response. When this ratio is an integer, L ≈ 1, 2, 3, . . . , the lag angle Ψ is approximately
a multiple of 2π. Since sin 2jπ = 0 (j = 0, 1, 2, . . . ), it follows that zRr (t) ≈ zRf (t). Hence,
the road surface primarily excites the bounce motion of the vehicle. By contrast, when
L 1 3 5 π
L ≈ 2 , 2 , 2 , . . . , the lag angle is approximately ± 2 so that the excitations in the front
and rear wheels are opposite in sign, exciting pitch motions. Note that the lag angle Ψ is
not a function of the vehicle speed.
Because pitching motions are excited whenever there is lag, it is preferable to have, at
least in passenger vehicles, a higher natural frequency of the rear suspension than the front
(consequence of the first Olley criterion). Because a higher frequency is associated with a
smaller period, the transient motion amplitudes of the rear suspension will oscillate faster
than the front suspension, thereby “catching up” in magnitude to the front suspension. This
helps reduce the pitching motion of the vehicle.
788 Applied Dynamics

15.8.2 Frequency Response for a Single-Degree-of-Freedom Model


Chapter 6 discussed the general form of the frequency response. The harmonic excitation
F (t) can be written as F (t) = F0 sin ωt or as F (t) = F0 cos ωt. When analyzing the response
of a system to harmonic excitation, we are not interested in an expression for the response
amplitude as an explicit function of time or in terms of initial conditions. Rather, our
interest is in frequency-domain analysis and in steady-state amplitudes and phase angles.
Comparing with the homogeneous plus particular solution approach in solving an ordinary
differential equation, the steady-state solution becomes similar to the particular solution.
It is convenient to use complex variable notation in expressing the harmonic excitation,
F (t) √= F0 eiωt , where eiωt = cos ωt + i sin ωt and i is the complex variable defined by
i = −1. Hence, for the single-degree-of-freedom quarter-car model, we can write the
equation of motion as
F0 iωt
z̈ + 2ζωn ż + ωn2 z = e (15.50)
m
Let us seek a steady-state solution to the above equation in the form z (t) = Zeiωt .
Substituting this solution into the above equation and collecting terms gives
F0 iωt
−ω 2 + 2iζωωn + ωn2 Zeiωt =

e (15.51)
m
Canceling the eiωt term from both sides and dividing the entire equation by ωn2 , we can
solve for the amplitude Z, with the result
F
Z = G (iω) (15.52)
mωn2

in which G (iω) is the frequency response. Recalling the nondimensional frequency ratio
ω̄ = ω/ωn introduced in Chapter 6, the frequency response can be rewritten as
1
G (iω̄) = (15.53)
1− ω̄ 2 + 2iζ ω̄

The frequency response G (iω̄) is a complex number. Its magnitude, defined as the
magnification factor in Chapter 6, is
1
|G (iω̄) | = q (15.54)
2 2
(1 − ω̄ 2 ) + (2ζ ω̄)

The magnitude of the frequency response is given in Figure 15.20. As expected, the
amplitudes become larger near resonance, when ω̄ = ωωn = 1, and as the damping factor ζ
becomes a smaller quantity.
When the harmonic excitation is generated by the road surface, so that F (t) =
keq zR (t) + cS żR (t) and zR (t) = Aeiωt , we can use the same approach as above and set the
response as z (t) = Zeiωt . Introducing this expression into the equation of motion gives

−ω 2 + 2iζωωn + ωn2 Zeiωt = A ωn2 + 2iωωn eiωt


 
(15.55)

Solving for the amplitude Z results in

Z (iω̄) = AH (iω̄) (15.56)


Vehicle Dynamics—Bounce, Pitch, and Roll 789

!
"#$ ȟ!()'#&
"
ȟ!()'#%
*$+%Ȧ,* %#$
ȟ!()'#"
%
ȟ!()'#$
&#$
ȟ!()&
&
ȟ!()%
'#$
'
' '#$ & &#$ % %#$ "
Ȧ "!ȦȦ#

FIGURE 15.20
Frequency response |G (iω̄) |.

where H(iω̄) is the frequency response function, also known as the transmissibility, whose
magnitude is
q
2
1 + (2ζ ω̄) q
2
|H (iω̄)| = q = |G (iω̄)| 1 + (2ζ ω̄) (15.57)
2 2 2
(1 − ω̄ ) + (2ζ ω̄)

Transmissibility is a measure of how much the shape of the road imperfection is transmit-
ted to the vehicle. It is preferable to design the suspension system with as little transmitted
motion from the road imperfection as possible. To carry out this design, it is instructive to
investigate transmissibility plots, |H (iω̄)|, which are given in Figure 15.21 with the damping
ratio as a parameter.
The plots are similar to those of |G (iω̄)|, with small differences. The frequency√ratio at
which the transmissibility is the same for all √ damping factors is when ω̄ = ω/ωn = 2. This
value of the excitation frequency of ω = 2ωn is also known as the crossover frequency.
At frequencies less than the crossover frequency the transmissibility is greater that 1. Af
frequencies higher than the crossover frequency, the transmissibility is less than 1. It is
interesting to note that

• When the ratio ω̄ < 2, the transmissibility becomes higher as the damping factor
decreases.

• When the ratio ω̄ > 2, the transmissibility becomes lower as the damping factor
decreases.

For a vehicle on a wavy road, the excitation frequency becomes higher with higher speed.
As discussed in Example 6.17, a vehicle that goes over a pothole experiences lower ampli-
tudes for higher vehicle speeds. It is thus preferable for a vehicle to have higher damping
factors for low speeds and lower damping factors for high speeds.
In addition to different damping properties as a function of speed, it is preferable for
the damper mechanism to have different constants during jounce and rebound. Consider
Figure 15.22. During jounce, the downward motion is resisted by the compressing spring,
790 Applied Dynamics

"#$
ȟ#+,'#&
"
ȟ#+,'#%
(!)ȓȦ*( %#$
ȟ#+,'#"
%
ȟ#+,'#$
&#$

&
ȟ#+,% ȟ#+,&
'#$

'
' '#$ & &#$ % %#$ "
Ȧ "#ȦȦ$

FIGURE 15.21
Transmissibility |H(iω̄)|.

as well as the damper. The spring aids the damper in absorbing the energy of the motion
so a high damping factor is not needed.
By contrast, during rebound, the potential energy stored in the spring is released and
the spring and damping forces act in opposite directions. This requires higher damping to
attenuate the vibration. Combining this analysis with the previous observation about low-
speed and high-speed travel, we can plot a desired force vs. velocity curve of a vehicle, as
shown in Figure 15.23.
The dashed lines indicate the force generated by earlier dampers where the damping
constant has the same value for both compression and rebound. Newer dampers generate
higher forces during rebound, as represented by the dotted line. Represented by the solid
lines are the latest in damper technology, where the damping force changes as a function

Undeformed position

kx cx kx cx

x
x= v Jounce v v Rebound x= v

k c k c

FIGURE 15.22
Restoring forces during jounce and rebound.
Vehicle Dynamics—Bounce, Pitch, and Roll 791

FIGURE 15.23
Force vs. velocity curve for a damper.

of velocity. The locus of points at which the slope of the force vs. velocity curve changes is
called the split line and it is a design criterion.
Recall the discussion earlier that human response to ride is more sensitive to acceleration
than to velocity. This means that when we look at response amplitudes, we should also be
interested the acceleration response. The amplitude of the position is AH (iω̄). By differen-
tiating the response equation twice, the expression for acceleration becomes Aω 2 H (iω̄).

15.8.3 Frequency Response for Two-Degrees-of-Freedom Models


For the two- (or higher-) degrees-of-freedom models, such as the quarter-car model with
unsprung mass or half-car models for pitch and bounce, the excitation can be written in
vector form as

{F (t)} = {F0 } eiωt (15.58)

The steady-state solution has the form {z (t)} = {Z (iω)} eiωt . Introduction of this
expression to the equations of motion, and elimination of the eiωt term from both sides,
gives the frequency response equation
 2 
−ω [m] + iω[c] + [k] {Z (iω)} = {F0 } (15.59)

The matrix inside the brackets on the left side is called the impedance matrix [I (iω)].
The motion amplitudes can be obtained by inverting Equation (15.59), which gives

{Z (iω)} = [I (iω)]−1 {F0 } = [H (iω)] {F0 } (15.60)

where [H (iω)] is the transmissibility matrix. We need to take the absolute value of the
above equation in order to express the magnitudes in terms of real numbers. The interest
is in plotting the motion amplitudes as a function of the excitation for a range of the
frequencies. Also of interest is the response of one coordinate as a function of a certain
excitation. Writing the vector {Z (iω)}, matrix [H (iω)] and force {F0 } as
     
Z1 (iω) H11 (iω) H12 (iω) F1
{Z (iω)} = [H (iω)] = {F0 } = (15.61)
Z2 (iω) H21 (iω) H22 (iω) F2
792 Applied Dynamics

the individual frequency responses become

Z1 (iω) = H11 (iω) F1 + H12 (iω) F2 Z2 (iω) = H21 (iω) F1 + H22 (iω) F2 (15.62)

For example, to analyze the effect of F1 on Z2 (iω), we plot H21 (iω). Each plot of
Hij (iω) (i, j = 1, 2) has two peaks, one at each natural frequency, with the sharpness of
the peaks depending on the damping properties. The peaks are usually higher for the first
natural frequency, an expected observation because the modes associated with the lower
frequencies contribute more to the motion than the modes of the higher frequencies.

Example 15.6
Given the two-degrees-of-freedom quarter-car model in Figure 15.11a, calculate and plot
the magnitude of the frequency response for the suspension spring (Fspring /FU ) and tire
forces (Ftire /FU ), where FU is the force that acts on the tire, due to road imperfection and
over the frequency range from 0 to 25 Hz for the sedan (softer suspension) and the sports
car (stiffer suspension). The suspension data are given in Table 15.3.

TABLE 15.3
Suspension system parameters

Parameter Sedan Sports Car


Sprung mass mS (kg) 375 275
Unsprung mass mU (kg) 40 25
Suspension stiffness kS (N/mm) 16 32
Tire stiffness kt (N/mm) 175 200
Suspension damping cS (N·s/mm) 1.2 2

The spring and tire forces are

Fspring = kS (z − zU ) , Ftire = kt zU [b]

where FU (t) = FU eiωt as the harmonic excitation that acts on the tire. It follows from the
notation above that Z1 = z, Z2 = zU , F1 = 0, and F2 = FU .
We calculate the impedance matrix defined in Equation (15.59) and calculate the values
of Z1 (iω) and Z2 (iω). Hence,

Fspring (iω) = kS (Z1 (iω) − Z2 (iω)) Ftire (iω) = kt Z2 (iω) [b]

The forces as a function of the excitation frequency are given in Figure 15.24 for the
sedan and in Figure 15.25 for the sports car. The undamped frequencies of the vehicles are
0.995 and 11.00 Hz for the sedan, and 1.593 and 15.35 Hz for the sports car.
In both figures, the tire forces are higher, because the spring constant for the tire is
higher than the spring constant for the suspension. Also, the tire force for both vehicles is
quite high when the excitation frequency is close to the second national frequency. This is
because the second frequency is very close to the wheel hop frequency.

Example 15.7
Consider the pitch-bounce model in Example 15.5 and plot the frequency response of zG (t)
and θ(t).
Vehicle Dynamics—Bounce, Pitch, and Roll 793

2.5
Spring Force
2 Tire Force
Amplitude Ratio
1.5

0.5

0
0 5 10 15 20 25
Frequency (Hz)

FIGURE 15.24
Frequency response of quarter-car model of sedan.

1.5
Spring Force
Tire Force
Amplitude Ratio

0.5

0
0 5 10 15 20 25
Frequency (Hz)

FIGURE 15.25
Frequency response of quarter-car model of sports car.

The relevant data is given in Example 15.5. Here, we assume that there is a harmonic
force F = F0 eiωt is acting on the model (see Figure 15.13). Hence, the force vector is
 
F0
{F (t)} = eiωt [a]
0

We calculate the impedance matrix, invert it, and find the amplitudes using
   
ZG (iω) −1 F0
= [I] [b]
Θ (iω) 0

We obtain the values for the range of the excitation frequency of 0 to 3 Hz. Figure
15.26 plots |ZG (iω)| and |Θ (iω)| for F0 = 50. Because we have an undamped system and
the peaks of the frequency response become infinity at resonance, we added proportional
damping to the system in the form of a damping matrix proportional to the stiffness matrix
by
[c] = 0.01 [k] [c]
794 Applied Dynamics

0.25
Displacement
0.2 Angle
Amplitudes
0.15

0.1

0.05

0
0 0.5 1 1.5 2 2.5 3
Frequency (Hz)

FIGURE 15.26
Frequency response of pitch and bounce model.

The plots indicate that, as expected, both the displacement zG and angle θ have lower
amplitudes at the second frequency. This difference in amplitudes at the frequencies f1 =
1.059 Hz and f2 = 1.263 Hz between the z displacement and angle θ correlate well with the
amplitudes of the modal vectors, which are (in normalized form)
   
−0.1094 0.0737
{u1 } = {u2 } = [d]
0.0170 0.0252

15.9 Roll Dynamics


This section considers the roll motion of a vehicle. The vehicle is viewed as a half-car from
behind, as shown in Figure 15.27, and the equations of motion are derived. Then, we consider
simplification of the model.

Sprung mass y
G

kS cS cS kS z

Unsprung mass

s
t

FIGURE 15.27
Rear view of roll dynamics.

To analyze roll, we ignore vibration of the unsprung mass and consider only the sprung
Vehicle Dynamics—Bounce, Pitch, and Roll 795

mass. Also, for the time being, consider a single axle. The motion variables are conveniently
selected as the vertical deformation of the sprung mass, denoted by z (positive downward
and measured from the undeformed position of the rear of the vehicle) and the roll angle φ.
The positive direction of the roll angle is clockwise. Also, the suspension stiffness on both
sides is the same and we assume that the dampers act at the same positions as the springs.
The free-body diagram is given in Figure 15.28.

s/2 s/2 Undeformed


i
= z – 2s sin A position
z
y o
= z + 2s sin
G Deformed
kS i+cS M position
i
z F kS o+cS ˙ o
Ground

FIGURE 15.28
Free-body diagram of sprung mass.

The spring deformations on the outside (right) and inside (left)4 are
s s
∆o = z + sin φ ∆i = z − sin φ (15.63)
2 2
where s is the distance between the two suspension springs (and the dampers) and z is the
deflection of the center of mass. The variable z is the same vertical translation (bounce)
variable that was used when discussing the quarter-car model. The weight of the axle is
denoted by W 0 = m0 g, so that for the front axle W 0 = Wf and for the rear axle W 0 = Wr .
Consider small deformations of the springs and use a small angles assumption, sin φ ≈ φ.
Each suspension spring has a spring constant kS (some use the ride rate of the suspension
system, keq , instead) and damper cS .
The resultant of the external excitations on the axle is a vertical force Fφ that acts
through the center of mass and resultant external roll moment Mφ about the x axis. Sum-
ming forces in the vertical direction and moments about the center of mass yields
X  s   s 
+↓ F = m0 z̈ = W 0 − kS z − φ − kS z + φ
2 2

 s   s 
−cS ż − φ̇ − cS ż + φ̇ + Fφ (15.64)
2 2

X s  s  s  s 
 MG = Iφ0 φ̈ = kS z − φ − kS z + φ
2 2 2 2

s  s  s  s 
+ cS ż − φ̇ − cS ż + φ̇ + Mφ (15.65)
2 2 2 2
in which Iφ0 is the mass moment of inertia (the roll moment of inertia) of the axle sprung
4A clockwise roll angle is generated when a vehicle takes a left turn, making the tire on the left the inside
tire.
796 Applied Dynamics
Wr
mass about the x axis. It follows that for the rear axle Iφ0 = W Iφ , and for the front axle
Wf
Iφ0= W Iφ , with Iφ denoting the roll moment of inertia of the entire vehicle. A low roll
moment of inertia vehicle, as in a sports car, has more responsive handling, but also a
rougher ride. A vehicle with higher Iφ is less nimble, but it offers a smoother ride. The
equations of motion for a given axle simplify to
1 2 1
m0 z̈ + 2cS ż + 2kS z = m0 g + Fφ Iφ0 φ̈ + s cS φ̇ + s2 kS φ = Mφ (15.66)
2 2
The two equations of motion are independent because the spring and damper constants
are the same on both sides of the axle and the center of mass is equidistant from each spring.
The translational equation of motion is the same as for the quarter-car model. In the rear
axle, there are two quarter-cars whose total mass is m0 = W 0 /g and their combined spring
constant is 2kS .
Because the bounce motion has been analyzed earlier, it will not be pursued here. The
focus here is on the roll motion. To this end, we introduce the terms roll stiffness kφ and
roll damping cφ and define them in terms of the front and rear axles as

1 1 1 1
kφr = kS s2 kφf = kS s2 cφr = cS s2 cφf = cS s2 (15.67)
2 r 2 f 2 r 2 f
where different spring and damping constants have been defined for the front (f ) and the
rear (r). It follows that the roll stiffness and damping for the entire vehicle become

kφ = kφr + kφf cφ = cφr + cφf (15.68)

Replacing the Iφ0 term in the roll equation in Equation (15.66) with the roll moment of
inertia of the entire vehicle, Iφ , and introducing roll stiffness and roll damping terms for
the front and rear axles from the above equation, the roll equation of motion for the entire
vehicle takes the form

Iφ φ̈ + cφ φ̇ + kφ φ = Mφ (15.69)

where Mφ is the total external roll moment acting on the vehicle.


The undamped natural frequency of the roll motion can be obtained as
s s
kφ kφr + kφf
ωφ = = (15.70)
Iφ Iφ

and the roll damping factor ζφ can be found from

cφ cφr + cφf
2ζφ ωφ = = (15.71)
Iφ Iφ

The Olley criteria discussed earlier in this chapter call for the roll natural frequency
to be close to the natural frequencies associated with pitch and bounce. This is used as a
guideline for selecting the spring constants and the roll moment of inertia.
We will not discuss here the transient motion or frequency response of the roll vibrations
of a vehicle. Rather, the static equilibrium of the roll motion is of interest, especially in the
case when a lateral force is applied.
Vehicle Dynamics—Bounce, Pitch, and Roll 797

Example 15.8
Consider Example 15.5 and design the damping constant and roll moment of inertia so
that the Olley criterion for the roll frequency is satisfied. Consider the distance between the
springs as s = 0.95t, where t = 60 in. is the track. Also, design the roll damping constant
to be ζφ = 0.25.
The vehicle parameters are Wf = 1000 lb, Wr = 850 lb, kf = 95 lb/in., kr = 130
lb/in., L = 110 in., and DI = 0.9, and the bounce and pitch frequencies were calculated as
f1 = 1.059 Hz and f2 = 1.263 Hz. Let us select the roll frequency as fφ = 1.2 Hz, so that
the roll natural frequency becomes

ωφ = 2π × fφ = 2π × 1.2 = 7.540 rad/sec [a]

Noting that the axle spring constants are twice the individual spring stiffnesses, the roll
stiffnesses for the front and rear become
1 2 1 2
kφr = s kSr = s2 kr = 0.25 × (0.95 × 60) × 130 = 105, 593 lb · in. [b]
2 4
1 2 1
kφf = s kSf = s2 kf = 0.25 × (0.95 × 60)2 × 95 = 77, 164 lb · in. [c]
2 4
From Equation (15.70), the roll moment of inertia becomes
kφr + kφf 105, 593 + 77, 164
Iφ = 2 = = 3214.6 lb · in. · sec2 [d]
ωφ 7.540 2

Converting the units into slug·ft2 , we get


3214.6
3214.6 lb · in. · sec2 = = 267.89 slug · ft2 [e]
12
Let us compare this number with the pitch moment of inertia. Because the pitch moment
of inertia describes the mass distribution of the vehicle from the side and the roll moment
of inertia from the rear (or front), we expect the roll moment of inertia to be much smaller.
Using Equation (15.46), the square of the radius of gyration is

κ2 = DI × b × c = 0.9 × 4.212 × 4.955 = 18.783 ft2 [f ]

so that the pitch moment of inertia becomes


W 2 1850
Iθ = κ = × 18.783 = 1080.1 slug · ft2 [g]
g 32.17
As expected, the pitch moment of inertia is considerably larger than the roll moment of
inertia.
The next step is to design the dampers. For simplicity, let us design all the dampers as
the same. It follows that the roll damping becomes

cφ = 2cφr = cS s2 [h]

and, from Equation (15.71), the roll damping is related to the damping factor by

cφ = 2Iφ ζφ ωφ [i]

Combining Equations [h] and [i], we can solve for the damping constant as
cφ 2Iφ ζφ ωφ 2 × 3214.6 × 0.25 × 7.540
cS = = = 2 = 3.730 lb · sec/in. [j]
s2 s2 (0.95 × 60)
798 Applied Dynamics

15.10 Roll Center Analysis


The sprung mass has so far been modeled as a rigid rod supported by two springs. This
section considers the instantaneous center of zero velocity of the sprung mass. Recall that
the instant center defines the point about which a rigid body rotates at a particular instant.
The location of the instant center is based on the geometry of the body.
The roll center is defined as the instant center about which the sprung mass rotates. The
location of the roll center is determined by the suspension geometry. This location changes
as the sprung mass moves.
The roll center location is an important factor when designing suspensions, and it is
different from the location of the center of mass. In this section, we assume that we know
the location of the roll center and we derive expressions to quantify the amount of roll. Ap-
proaches to locate the roll center as a function of the suspension geometry will be developed
later in this chapter.
While an instantaneous center is only a predictor of velocities and angular velocities at
a particular instant and its location changes as the body moves, we can extend this line of
thought and think of the roll center as a moment center for a short period of time. Hence,
the roll center can be used as a fixed point about which we can sum moments. The validity
of this assumption, that the location of the roll center moves little as the sprung mass rolls,
must be verified at all times.

Sprung mass y

G W'
c , k  z
h

RC
hRC Unsprung mass

FIGURE 15.29
Roll center model, rear view.

Considering the assumption above, the roll geometry (Figure 15.29) can be depicted as a
pivot point, located at the roll center, about which the sprung mass rotates. The roll stiffness
is represented by a torsional spring kφ , and a torsional damper cφ provides damping.
The height of the roll center from the ground is denoted by hRC and the distance between
the roll center and center of mass is h, so the height of the center of mass from the ground
is hG = hRC + h, as shown in Figure 15.30a. Most vehicles have a higher roll center in the
rear and a lower roll center in the front. Also, the roll center maybe below ground, as in
Figure 15.30b.
The free-body diagram shown in Figure 15.31 is for a roll center that is above ground
Vehicle Dynamics—Bounce, Pitch, and Roll 799

a) b)
G
G
h
hG
RC h
hRC Ground

hRC
RC

FIGURE 15.30
Roll center geometry, when φ = 0. a) Roll center above ground, b) roll center below ground.

but below the center of mass. The sprung mass has rotated clockwise, which is the positive
direction for the roll angle φ. This can occur when the vehicle takes a left turn, so that the
0 2
lateral force at the center of mass is the inertia force of Fy = mRV acting to the right. As
the vehicle is taking a left turn, the inside tires are on the left. In such a turn, the lateral
tire forces FLi and FLo act to the left and Fy = FLi + FLo . A roll angle can also develop
by wind forces or by gravity on a banked road.
The reaction forces at the roll center are denoted by Cy and Cz . We can think of the
roll center RC as the point such that a lateral force applied at the roll center will not cause
the sprung mass to roll.
The moments generated by the suspension system are, in terms of the torsional spring
and damper of constants, M = kφ φ + cφ φ̇. The sum of moments about the roll center is
X
 MRC = −kφ φ − cφ φ̇ + Fy (hG − hRC ) cos φ + W 0 (hG − hRC ) sin φ (15.72)

Because we are assuming that the roll center is stationary, we can sum moments about
it, which gives
X
 MRC = IRC φ̈ (15.73)

in which IRC is the mass moment of inertia of the axle sprung mass about the roll center,
related to the roll moment of inertia Iφ0 , which is about the center of mass, by the parallel
axis theorem

IRC = Iφ0 + m0 h2 (15.74)

The roll angle is usually small, justifying a small angles assumption, which leads to the axle
equation of motion

IRC φ̈ + cφ φ̇ + (kφ − W 0 h) φ = Fy (hG − hRC ) = Fy h (15.75)

This equation describes the roll motion of the vehicle more accurately than Equation
(15.69). The two equations become the same when the center of mass and roll center loca-
tions coincide. Also, we can write an equation of motion for each axle and add the two to
obtain the roll equation of motion for the entire vehicle.
We will now make yet another assumption, that equilibrium is reached rapidly, and
consider the state of roll equilibrium. Setting all the time derivatives to zero and substituting
800 Applied Dynamics

W'
Sprung mass
G Fy

k c h
y
Cy
Cz
z
k c Cz WU

RC Cy
hRC Unsprung mass

Wi FL i Wo FL o
t

FIGURE 15.31
Free-body diagrams of the sprung and unsprung masses.

m0 V 2
Fy = R give the roll angle due to a lateral force as

Fy h m0 V 2 h
φ = = (15.76)
kφ − W 0 h R (kφ − W 0 h)
From the numerator of the above equation, the roll angle depends on the lateral force as
well as the distance h (between the center of mass and the roll center). The denominator of
Equation (15.76) indicates that a positive value for h results in a lower effective roll stiffness
kφ − W 0 h.
Increasing roll center height hRC , while keeping the center of mass G at the same loca-
tion, has the effect of reducing h. This creates a smaller roll moment Fy h. As the roll center
height is increased, the effective stiffness kφ − W 0 h increases. As a result, the steady-state
value of the roll angle becomes smaller.
Next, consider the load transfer from the inside to the outside wheels. To this end,
consider the unsprung mass and sum moments about the roll center. We can assume that
the unsprung mass is not rolling itself and that the unsprung mass has little flexibility, so
the tire deformation is negligible. It follows that the angular acceleration of the sprung mass
about the x axis is zero. Summing moments for φ̇ = 0 and φ̈ = 0 yields
X t
 MRC = 0 = (Wi − Wo ) + FL hRC + kφ φ (15.77)
2
where the total lateral force is FL = FLi + FLo = Fy = m0 V 2 /R. Substituting this result
into the above equation, the load transfer is obtained as
2
Wo − Wi = (Fy hRC + kφ φ) (15.78)
t
Vehicle Dynamics—Bounce, Pitch, and Roll 801

Substituting the value of the roll angle calculated in Equation (15.76) to the above equation,
we obtain the weight shift.
Equation (15.78) is more accurate than the expression for the weight shift developed in
Chapter 12 (Equation (12.60)), when no distinction was made between the unsprung and
sprung masses and the roll center and suspension characteristics were not considered.
The term Fy hRC = FL hRC is known as the nonrolling overturning moment and it
affects the jacking of the sprung mass. A smaller value for roll center height hRC (or larger
h) reduces the magnitude of this moment.

15.11 Lateral Force Reduction Due to Weight Shift


The vertical load transfer (weight shift) discussed in the previous section has the effect
of changing the lateral forces. This is because, as described in Chapter 12, a nonlinear
relationship exists between the lateral force and wheel load. While the maximum lateral
force that can be generated by the tires increases for higher wheel loads, the increase is not
linear and the increase in lateral force is less than the increase in wheel loads. A typical
plot of lateral force versus vertical force is given in Figure 15.32 with the slip angle as the
parameter.

"%
!&
!%
$
!"
#

"
!
Į'()*+
#

FIGURE 15.32
Plot of lateral force vs. vertical force (wheel load).

The rate of change in lateral force with respect to vertical load for a given slip angle is
called vertical force sensitivity.
Figure 15.33 depicts the lateral forces that are generated when there is a difference in
the inside and outside wheel loads for a given slip angle. Let the wheel load on each wheel
(before roll occurs) be W . The lateral force FL is obtained from Figure 15.32 for a given
angle of slip α. The outside and inside wheel loads due to roll are Wo and Wi , and the
associated lateral loads are FLo and FLi . It follows that

Wi + Wo
W = (15.79)
2
802 Applied Dynamics

!"

!"&

!"
!"# $ !"&
!"# !

%
%# % %&

FIGURE 15.33
Change in lateral force.

It is clear from Figure 15.33 that the sum of the lateral forces on the inside and outside
wheels is less than the lateral force in the absence of weight shift, or
FLo + FLi
FL > (15.80)
2
The weight shift from the inside to the outside wheels due to cornering results in re-
duction in the overall lateral force that can be generated by the tires. Experimental results
−Wi
show that for a load transfer ratio WoW of about 0.6, which is a commonly encountered
value, the ratio of reduction in lateral force, given by
F +FLi
∆FL FL − Lo 2
= (15.81)
FL FL
is about 20%. In other words, a 60% load transfer results in a 0.6 × 0.2 = 12% reduction in
the lateral force that is generated. This reduction also has an effect on the understeer (or
oversteer) properties of the vehicle.
To further analyze the effect of weight shift and subsequent change in lateral forces,
consider an axle, front or rear, of weight W , so that the wheel load in each tire is W/2
before any weight shift occurs. Let us consider the nonlinear relationship in Equation (13.6)
that relates the cornering stiffness of a tire to the wheel load as
 2
W W
Cα = e1 − e2 (15.82)
2 2
where e1 and e2 are positive constants, whose values depend on the tire properties. Note
that in the above expression the wheel cornering stiffness is treated as a positive quantity.
The cornering stiffness of the axle, which was defined in previous chapters as a negative
quantity, is then
 e2 
C = −2Cα = − e1 W − W 2 (15.83)
2
After weight shift due to roll, the new wheel loads are
W W
Wo = + ∆W Wi = − ∆W (15.84)
2 2
Vehicle Dynamics—Bounce, Pitch, and Roll 803

It follows that the cornering stiffness associated with the axle becomes
 2 !  2 !
0 Wo Wo Wi Wi
C = − e1 − e2 − e1 − e2
2 2 2 2

2
= C + 2e2 (∆W ) (15.85)
As expected, the cornering stiffness of the axle becomes smaller in the presence of weight
shift (remember that C is negative). Because the nonlinear effects are usually small, we can
approximate 1/C 0 as
!
2
1 1 1 1 (∆W )
= ≈ 1 − 2e2 (15.86)
C0 C 1 + 2e2 (∆W )2 C C
C

Let us evaluate the effect of the reduction in the cornering stiffness on the over-
steer/understeer properties. The cornering stiffnesses in the presence of weight shift can
be written for the rear and front axles as
2 2
Cr0 = Cr + 2e2 (∆Wr ) Cf0 = Cf + 2e2 (∆Wf ) (15.87)
Substituting Cr0 and Cf0 for Cr and Cf in Equation (14.20), a revised understeer gradient
K 00 can be defined as
Wr Wf
K 00 = 0
− 0 (15.88)
Cr Cf
Introducing the approximation in Equation (15.86) into the revised understeer gradient
in Equation (15.88) and keeping in mind the sign convention for cornering stiffness, the
understeer gradient becomes
Wr Wf Wr  2
 W 
f 2

K 00 = − = K 0
− 2e 2 (∆W r ) + 2e 2 (∆W f ) (15.89)
Cr0 Cf0 Cr2 Cf2
Noting that the understeer gradient K 0 is a positive number for understeer vehicles, we
conclude that
• Weight shift from inside to outside tires in the front axle increases the understeer gra-
dient,
• Weight shift from inside to outside tires in the rear axle decreases the understeer gradi-
ent.
Because it is preferable not to reduce the understeer gradient, there should be less weight
shift in the rear axle, hence less roll in the rear axle, than the front. Obtaining higher roll
stiffness usually requires suspensions with higher spring constants.5 Installation of anti-roll
bars provides additional roll stiffness without increasing spring constants. Anti-roll bars will
be discussed in Section 15.20.

15.12 Roll Axis


The previous analysis of roll considered one axle of the vehicle, rear or front. The front and
rear axles of a vehicle usually have different suspension characteristics, including suspension
5 This also is stressed in the first of the Olley criteria.
804 Applied Dynamics

hRC f

h
Fy=Wv2 hG
Rg
G
Roll axis W

y
hRC f

v
x z

FIGURE 15.34
Roll axis configuration.

stiffness and roll center height. This means that the amount of roll of the front versus the
rear is different. Because the entire vehicle rolls with one angle, we combine the roll of the
front and rear by considering an axis about which the entire sprung mass rolls.
Denote by kφf , kφr , hRCf , and hRCr , the roll stiffnesses and roll center height of the
front and rear axles. The line joining the front and rear roll centers is called the roll axis,
and the sprung mass can be viewed as rotating about the roll axis. Figure 15.34 depicts
the roll axis, and the side view of the roll axis and front view of the roll motion are shown
in Figure 15.35. Because the roll axis is at an angle with the x axis, an accurate analysis
has to be three-dimensional. However, since the angle the roll axis makes with the x axis is
small, we can make use of the small angles assumption.
The first step is to calculate the angle  that the roll axis makes with the x axis as

hRCr − hRCf
 ≈ tan  = (15.90)
L
Point D in Figure 15.35 is the intersection of the roll axis and the vertical line going
through the center of mass. The vertical distance between the center of mass and the roll
axis is denoted by h, where

h = hG − hRCf + b tan  (15.91)

Summing moments about the roll axis and considering the steady-state case gives
X
M = 0 = −kφf φ − kφr φ + (Fy h cos φ + W gh sin φ) cos  (15.92)
Vehicle Dynamics—Bounce, Pitch, and Roll 805

a) G
RCr
h x
RCf x'
v D hG hRC r
hRC f A B
z z'
b Side view

b) Fy c)
G y Angle view
y

h
W x
Front view
D x' z z'
z

FIGURE 15.35
a) Side view of roll axis, b) front view, and c) angle view.

where W = mS g is the total weight of the sprung mass. Using small angle assumptions for
both φ and , we can solve for the roll angle as
a
Fy h W h gy
φ = = (15.93)
kφf + kφr − W h kφf + kφr − W h

The above equation expresses the roll angle in terms of a normalized lateral acceleration
(ay /g). The derivative of the roll angle with respect to the normalized lateral acceleration
is known as the roll gradient and it has the form
dφ Wh
Rφ = = (15.94)
d (ay /g) kφf + kφr − W h

Solving the above equation for the roll center distance h from the center of mass gives

Rφ kφf + kφr
h = (15.95)
W (1 + Rφ )

The roll gradient Rφ is dimensionless. It is expressed in radians in the above equation.


For passenger vehicles, the roll gradient is in the range of Rφ ≈ 3◦ to 7◦ ≈ 0.05 to 0.13
rad, so it is a small quantity. Hence, Equation (15.95), which is in terms of radians, can be
approximated as

Rφ kφf + kφr
h ≈ (15.96)
W
so a near linear relationship exists between the roll gradient and roll center height.
We can now consider both axles separately and sum moments to calculate the weight
shift due to roll. The results turn out to be
 
ay 2
Wfo − Wfi = Wf hRCf + kφf φ
g tf
806 Applied Dynamics
 
ay 2
W ro − W ri = Wr hRCr + kφr φ (15.97)
g tr
where tf and tr denote the front and rear track length, respectively. Substituting the ex-
pression for φ from Equation (15.93) into the above equation gives the load transfer on each
axle due to roll.
When the suspension geometry is not sufficient to provide the desired roll stiffness, one
way to increase it is by using anti-roll bars. Anti-roll bars are discussed later in this chapter.
The above discussion indicates that as we begin to consider the different motions of a
vehicle in more detail, we see how they are related to each other.

15.12.1 Design for Cornering


The previous chapter dealt with the lateral stability of the vehicle. We discussed in this
chapter the roll motion of the sprung mass and how roll affects weight shift in an axle.
These two analyses are combined in this subsection.
Suppose you know all the properties of a vehicle: wheelbase, weight, wheel loads, track,
spring constants, roll center height, cornering stiffnesses, etc. Given a certain vehicle speed,
and a desired lateral acceleration level (such as maximum acceleration) you wish to calculate
how much you need to turn the steering wheel in order to achieve that acceleration. The
following procedure is one way of accomplishing this task.
1. Use Equation (15.93) to calculate the roll angle φ.
2. Use Equation (15.97) to calculate the weight shift on each axle.
3. Use the calculated weight shift on each axle and Equation (15.87) to calculate the new
axle cornering stiffnesses.
4. Use the new cornering stiffnesses and Equation (15.88) to calculate the new understeer
gradient K 00 .
5. Use the new understeer gradient and the expression for lateral acceleration gain in
Equation (14.35) to calculate the steer angle.
Note that the above procedure is for steady-state values of the accelerations and angular
velocities. The transients need to die out. The transient equations of motion for combined
lateral motion and roll are complicated and beyond the scope of this text.

Example 15.9
Calculate the roll axis height as a function of the roll gradient Rφ for a 3600 lb vehicle. The
front springs have a spring constant of 95 lb/in. and the rear springs have constants of 130
lb/in. each. The springs are s = 40 in. apart.
The roll center height h is given in Equation (15.95) as

Rφ kφf + kφr
h = [a]
W (1 + Rφ )
where the roll stiffnesses are
1 1
kφf = kSf s2 = × 95 × 402 = 76, 000 lb · in.
2 2

1 1
kφr = kS s2 = × 130 × 402 = 104, 000 lb · in. [b]
2 r 2
Vehicle Dynamics—Bounce, Pitch, and Roll 807

where kSf and kSr are the stiffnesses of each front and rear spring, respectively.
For a roll gradient of 5◦ , which in terms of radians is
5
Rφ = = 0.08727 rad [c]
57.296
using Equation [a] and solving for h we get

0.08727 × (76, 000 + 104, 000)


h = = 4.013 in. [d]
3600 × 1.087266

15.13 Introduction to Suspension Systems


We next consider suspension systems and their geometry. Suspension systems are among the
most complex and important parts of vehicles. Suspension systems connect the unsprung
mass to the sprung mass. They are designed to not only provide flexibility and ride isolation
through springs and vibration suppression through dampers, but also
• Keep the tires in continuous contact with the road,
• Provide a smooth and bump-free ride,
• Protect the vehicle from damage and wear,

• Resist roll of the vehicle and provide roll stability,


• Maintain steer angles and proper camber, an
• React to tractive and lateral forces on tires.
The design of a suspension system has to balance the advantages and disadvantages
of different types of suspensions. For example, suspension systems in older cars were in
the form of a solid axle, to which springs and dampers were attached. More modern de-
signs use independent suspensions, which have improved ride quality substantially, but have
introduced problems of their own, such as changes in camber rate, jacking, and scrub.
Suspension systems are modeled as linkages, and especially as four-bar or slider-crank
linkages. A four-bar linkage or a slider-crank mechanism, as discussed in Chapter 3, has one
degree of freedom. The developments in this section follow the notation and terminology of
the texts by Gillespie, Milliken and Milliken, and Jazar.

15.14 Suspension System Terminology and Geometry


There is a large number of suspension systems, and it is not possible to discuss all of
them here. We begin our study with the terminology used in suspension systems. Jounce
and bounce (or, compression and rebound) were discussed earlier. Considering the front
and rear suspensions together, when the rear suspension system compresses (and/or front
suspension goes up), the vehicle is pitching up (positive pitch), and its position is called
squat. When the front suspension compresses (and/or the rear suspension moves up) the
808 Applied Dynamics

!,$)- ./01
!
!
!"#$%&'()** !"#$%&'()**

+%*"#$%&'
()**

FIGURE 15.36
Squat and dive geometry.

vehicle is pitching down (negative pitch) and this position is called dive. Figure 15.36 depicts
the different motions of suspension systems.
Suspension systems can be categorized into two groups: axle suspensions and indepen-
dent suspensions. The schematic of suspension systems in both groups can be explained as
linkages, as we will soon see. This observation holds true when the suspension system is
viewed from the side, as well as from the rear.

I14 at

I13

I13

I24
3 I34
I34 2 I23
I23 3 4
4
2 I12 1
1 I14
I12
I24

FIGURE 15.37
Four-bar linkage and slider-crank mechanisms and their instant centers.

Chapter 3 discussed two basic linkages, the four-bar and the slider-crank mechanism.
Both are one-degree-of-freedom mechanisms. Figure 15.37 depicts these mechanisms, to-
gether with their instant centers. Recall that a mechanism with n links has n (n − 1) /2
instant centers. A mechanism with four links has 4 × 3/2 = 6 instant centers, denoted by
I12 , I13 , I14 , I23 , I24 , I34 .
For a planar mechanism with n links, of which one is grounded (fixed), and J joints,
there are three degrees of freedom for each individual link, minus three degrees of freedom
for the fixed link, minus two degrees of freedom for each joint (as a pin joint brings together
two points on two links and hence imposes two constraints). Mathematically,

d.o.f. = 3 (n − 1) − 2J (15.98)
Vehicle Dynamics—Bounce, Pitch, and Roll 809

The above equation is known as Gruebler’s equation. The four-bar linkage has four joints
and four links, so that Gruebler’s equation states that it has 3 × (4 − 1) − 2 × 4 = 1 d.o.f.
An important theorem concerning instant centers is Kennedy’s rule, which states that
any three bodies in a mechanism have three common instant centers and the three instant
centers lie along a straight line.

FIGURE 15.38
Instant center locating tool.

To find the instant centers of a mechanism, we begin with the instant centers that
can be identified visually. For the four-bar linkage, the four joints are the instant centers
I12 , I14 , I23 , I34 . A useful tool to locate the remaining instant centers is to draw a circle,
with the links marked as notches, as in Figure 15.38. Then, for each instant center observed
from the geometry, we draw a line that connects the joints involved in that instant center.
For example, we mark I14 by drawing a solid line that connects notches for links 1 and 4.
To locate instant center I13 , notice that two triangles in the locating tool have solid
lines that involve links 1 and 3: linkages 1-2-3 and 1-3-4. Considering the 1-2-3 linkage
combination, by Kennedy’s rule, instant centers I12 , I23 , I13 will lie on the same line. The
locations of two of these instant centers, I12 and I23 , is known. Hence, instant center I13
will lie the line that connects and extends from I12 and I23 . Similarly, for the 1-3-4 linkage
combination, the line that goes through I14 and I34 contains I13 . The intersection of these
two lines defines I13 . I24 is located in a similar way. From Figure 15.38, the link combinations
for this case are 1-2-4 and 2-3-4.
The process is slightly different for the slider-crank mechanism because I14 is at infinity.
To locate I13 two lines are needed: one going through I12 and I23 and the other going
through I14 and I34 . The line that goes through I34 is perpendicular to the line of action
of the slider. To locate I24 , consider links 2-3-4 and draw a line going through I34 and I23 .
Considering the 1-2-4 combination, a line needs to be drawn that goes through I12 and I14 .
But, I14 is at infinity. We draw a line that goes through I12 and is parallel to the vertical
line drawn earlier to locate I13 . Intersection of these two lines define I24 .
Suspension systems must provide a one-degree-of-freedom connection (in the vertical
direction) between the sprung mass and unsprung mass, as well as prevent motion of the
sprung mass with respect to the unsprung mass in all other directions (lateral, longitudinal)
and rotations. This one-degree-of-freedom motion is generated by designing the suspensions
as linkages. The linkages are known as suspension arms. Suspensions are designed as three-
dimensional components that provide five degrees of restraint, thus producing a one-degree-
of-freedom system.
Consider the projections of the suspension geometry. For a properly designed suspension,
the projection in each plane involving the vertical (side view or rear view) must be a slider-
810 Applied Dynamics

crank or four-bar mechanism. In addition, suspension systems should provide stiffness and
damping between the unsprung and sprung masses. Suspension systems should also prevent
rotational motion of the sprung mass with respect to the unsprung mass.
When viewed from the top, the suspension system should be a zero-degree-of-freedom
system. It is interesting to note that three links connected to each other by joints form a
triangle, such as the truss element in Figure 3.20c, which has zero degrees of freedom, and
hence, the links cannot move with respect to each other. For n = 3 and J = 3, Gruebler’s
equation states d.o.f. = 3 × (3 − 1) − 2 × 3 = 0.6
The different projections of a suspension system are summarized in Table 15.4, where x
is the motion direction, y is the lateral, and z is the vertical, as shown in Figure 15.39.

TABLE 15.4
Projection of suspension systems onto planes

Projection Plane View Type Degree of Freedom


xz Side view 1
xy Top view 0
yz Rear (or front) view 1

2(3-& !"#$%"$&%'(#
,(-.&/#!0
!"#$%"$& )*+"$&,(-.&/"!0
1-"$-*
!"#$%"$&
1-"$-*
"

FIGURE 15.39
Projection planes of suspension systems.

The line connecting the instant centers of the xz and yz planes defines the instantaneous
axis of rotation of the motion of the sprung mass with respect to the unsprung mass. It
should be noted that, because projections of suspension systems are linkages, the motion
of the sprung mass with respect to the unsprung mass is never purely vertical. It is in the
form of an arc about the instant center. This results in some rotational motion, leading to
camber, jacking, and scrub effects.
6 This is the reason trusses are the fundamental building blocks of structures.
Vehicle Dynamics—Bounce, Pitch, and Roll 811

15.15 Axle Suspensions


In axle suspensions, the suspension arms and components are attached to the axle. Axle
suspensions come in two forms. In solid axle (also known as dead axle) suspensions, the axle
does not transmit any power, as in the rear axle of a front wheel drive vehicle. By contrast,
a live axle transmits power by means of a differential and half shafts that are incorporated
into a single rigid unit.
The process of connecting an axle to the sprung mass is commonly known as locating
the axle. In addition to providing a connection between the axle and the vehicle, locating
an axle should also achieve the goal of providing limitations to the motion of the axle, when
viewed from the side as well as when viewed from the rear.
)2 32
" %
$
!"#$%&'
% ()**
! # #
$

"
!

+,"'-./0 !.1/'-./0

FIGURE 15.40
Parallel trailing arm suspension system: a) top view (xy plane), b) side view (xz plane).

There are several ways of providing the connection between the axle and the body. One
way is to use four trailing arms, also referred to as control arms, as shown in Figure 15.40.
In some versions of this suspension, the trailing arms are parallel. This configuration is not
effective in preventing the lateral motion of the axle with respect to the sprung mass. It
becomes necessary to use another mechanism to control the lateral motion of the axle.
A Panhard rod is a low-cost and simple design to prevent lateral motion between the
axle and sprung mass. Shown in Figure 15.41, a Panhard rod connects the sprung and
unsprung mass. The disadvantage of a Panhard rod is that the motion of the sprung mass
with respect to the axle is an arc, which results in some relative lateral motion. The radius
of the arc is the length of the rod, so that a longer rod is more effective. The Panhard rod
is usually more suitable for larger vehicles.
Later versions of trailing arm suspensions have used trailing arms that are angled, so
that a Panhard rod is not needed to locate the axle laterally, as illustrated by the top view
of the suspension mechanism in Figure 15.42. Extending the trailing arms results in the
triangular links, which provide the zero degrees of freedom.
Another common design of an axle suspension is the Hotchkiss drive, named after the
vehicle in which it was first introduced. Here, leaf springs (see Chapter 4) provide the spring
forces. Figure 15.43 illustrates the Hotchkiss drive and identifies the associated four-bar
linkage, as well as the instant centers.
The leaf spring, because of its compliance (or elastic deformation), acts as two linkages.
812 Applied Dynamics

!"#$%&'
()**

+)%,)#-'#.-
"

FIGURE 15.41
Panhard bar shown from the rear.

!"#$%&'
()**
&
'
!

$ % "

FIGURE 15.42
Angled trailing arm suspension, top view.

Considering the instant center I13 , the motion of the axle with respect to the sprung mass
is v3 . Figure 15.44 shows a picture of a leaf spring suspension system on a U.S. Army Jeep.
Leaf springs are cheaper than trailing arms, but they are heavier, which limits their ability
to locate the axle. Also a leaf spring is prone to twisting, thus leading to roll stability
problems.
Axle suspensions have several deficiencies, which become worse when leaf springs are
used, resulting in an uncomfortable ride and reduced stability. The main deficiencies are
• Large mass. Leaf springs, as well as the associated axles, are heavy. As a result,
– the overall weight of the vehicle is increased, which requires more power to propel
the vehicle, resulting in higher fuel usage, and
– the added weight of the leaf springs and axle makes the unsprung mass larger,
making the vehicle less maneuverable.
• Uncomfortable ride. With axles, the effect of a road imperfection, such as a bump or
pothole, is distributed throughout the vehicle, rather than isolated.
• Excess camber. When a vehicle goes over a bump or pothole, the entire axle is dis-
placed and it rolls, so that the tires, which are perpendicular to the axle, assume a tilted
Vehicle Dynamics—Bounce, Pitch, and Roll 813

v3
4
1
3
2
I13

FIGURE 15.43
Hotchkiss drive with leaf springs, shown as a linkage, together with instant center I13 .

FIGURE 15.44
Picture of Hotchkiss drive with leaf springs on a U.S. Army Jeep. Source: Wikimedia Com-
mons, http://commons.wikimedia.org/wiki/File:Leafs1.jpg (Last accessed August 1, 2014).

geometry. This causes undesired camber and affects the lateral forces generated by the
tires.
Trailing arm suspensions on axles are lighter than leaf springs and they provide better
control of the axle motions, but they have their own issues. The majority of modern cars
are front wheel drive and do not use axle suspensions. Leaf springs are used in heavier
vehicles and in trucks.

Example 15.10—Bump Steer7


Instant center analysis has several applications in the design of linkages, providing a visual
interpretation of the motion. As a mechanism moves, so do its instant centers. It is to the
advantage of the designer to design the associated linkage so that the instant center does
not change location too much (recall discussion earlier on roll center), as a large change in
the instant center location may result in unintended motion characteristics.
An interesting example is the phenomenon of bump steer, which was encountered in
cars of the 1970s. Figure 15.45a shows the side view of the rear suspension. Link 1 is the
7 This example was introduced to the literature in the text by Norton.
814 Applied Dynamics

%
%7

" ( 123'/
! "
%44'&536
#$%&' !!" )

*+,-./,0/$
$ %
57

#
( "
"
! )
#$%&' !!"

FIGURE 15.45
Bump steer geometry: a) travel on level road, b) going over bump.

sprung mass and link 3 is the wheel and axle. Instant center I13 is common to the sprung
mass and the axle. The velocity of the axle with respect to the sprung mass is perpendicular
to the line joining the instant center I13 and the center of the wheel.
Consider the case when the wheel goes over a bump (Figure 15.45b). The change in the
vertical orientation of the wheel causes I13 to move rearward. The line between I13 and the
center of the wheel becomes steeper, so that the velocity of the wheel with respect to the
sprung mass acquires a larger horizontal component than before the bump.
The two wheels have different horizontal velocities, with the wheel going over the bump
having a higher horizontal velocity. As shown in Figure 15.46, this has the effect of steering
the vehicle in a different direction, even though the driver produces no steer input. The
vehicle develops an angular velocity ω = ∆v/t, where ∆v is the difference in the horizontal
velocities and t is the track.
The bump steer problem was corrected by designing the trailing arms and the suspension
points so that the location of the instant center does not vary too much as a wheel goes
over a bump.

15.16 Independent Suspensions


Independent suspensions are by now standard in most automobiles, especially for the front.
Independent suspension systems permit each wheel to move up and down separately, they
weigh less than solid axle suspensions, they provide better isolation of a bump or shock,
they avoid uncontrolled oscillations between the sprung mass and the suspension system,
Vehicle Dynamics—Bounce, Pitch, and Roll 815

v+ v
v x
t

No bump Wheel going


over bump

FIGURE 15.46
Top view of the bump steer effect.

and they have better stability properties. However, they require more accuracy, and they
are more expensive and more difficult to design than solid axle suspensions.
Independent suspension systems can also be analyzed as mechanisms, as has been done
in the texts by Milliken and Milliken and by Jazar. The discussion in this section is based
on those developments.

)0

!"#$%&'
()** !

"

/0

-
3
!"#$%&' 4
. ()** 4
2
, 1

FIGURE 15.47
a) Rear view of SLA suspension, b) kinematic description as linkage.

There are a large number of independent suspension systems around. We consider two
commonly used suspensions, the double-A arm suspension, which is also known as double
wishbone or short-long arm (SLA), and the MacPherson strut suspension.
Figures 15.47 (links 4 and 6 for the left side and 5 and 7 for the right side are the short
816 Applied Dynamics

a) Sprung
mass

b) 4 5
Sprung
mass
8 3
2 3
2
6 7

FIGURE 15.48
a) Rear view of MacPherson strut, b) Kinematic description as slider crank.

and long arms) and 15.48 show these two suspension systems from the rear, together with
associated schematics depicting the equivalent linkages.
The SLA and other linkage suspensions are in essence four-bar linkages, when viewed
from the rear (yz plane) and MacPherson strut suspensions are slider-crank mechanisms.
When viewed from the side, these mechanisms need to have a special geometry or a trailing
arm, in order to complete the linkage on the xz plane and to define the instant center.
In a double wishbone or SLA suspension, if the upper and lower arms are not on parallel
planes, a linkage will not exist for the side view (xz plane). In a MacPherson strut, the
caster angle helps create the linkage viewed from the side. The geometry of the suspension
system determines the instant centers of both the front and side view at the same time.
Consider the rear view (yz plane) of one side of the suspension system, say, the left
(driver side). For the SLA suspension, as shown in Figure 15.49, links 2-4-6-8 make up the
four-bar mechanism, with link 2 denoting the tire and 8 the sprung mass. Links 4 and 6 are
the suspension arms.
The four joints are instant centers. The remaining instant centers I28 and I46 are found
by using Kennedy’s rule. Considering the link combinations 2-6-8 and 2-4-8, we extend the
lines of links 4 and 6 for I28 and we extend lines of links 2 and 8 for I46 . Instant center I28
can be used to evaluate the motion of the tire with respect to the sprung mass and vice
versa. Its counterpart on the right suspension is I38 and it is found the same way as I28 ,
but by considering the linkage 3-5-7-8.
A similar analysis is conducted for the MacPherson strut, as shown in Figure 15.50.
Because the MacPherson strut is a slider-crank mechanism, instant center I24 is at infinity.
Here, instant centers I24 , I26 , I48 , and I68 are identified by visual inspection. Link triplets
Vehicle Dynamics—Bounce, Pitch, and Roll 817
!$"
!!$ $

!!" ! "

!!# #
!#"

FIGURE 15.49
The 2-4-6-8 link and instant center I28 for SLA.

2-4-8 and 2-6-8 are used to locate I28 . The line connecting I24 and I48 goes through I24 and
is parallel to the line locating I24 . Hence, this line is perpendicular to link 4. Link triplets
2-4-6 and 4-6-8 are used to locate I46 . The line that connects I24 and I26 goes through I24
and is parallel to the lines mentioned above.

!"$

!#" ’ "
$
!#$
#
!!$
!
!#!
!"!

FIGURE 15.50
The 2-4-6-8 link and instant center I28 for MacPherson strut.

Table 15.5 lists the advantages and disadvantages of suspension systems. Even though
we did not conduct an analysis, we also include multilink suspension systems in the list, as
they constitute a complicated but desirable suspension system.

15.17 Roll Center Construction


The desired roll center height is obtained by adjusting the suspension arms and their pivot
points. In general, the lower arms have a bigger effect in roll center location. High perfor-
mance and race cars are built so that the roll center height can be changed easily.
It is very important that the sprung mass is distributed symmetrically between the left
and right sides and that the roll center lie on the vertical line going through the center
of mass. The consequence of asymmetry is uneven roll of the vehicle and uneven force
818 Applied Dynamics

TABLE 15.5
Comparison of suspension systems. (A: advantage, D: disadvantage)

Type Characteristics
Axle A. Simple design and low cost. Also simple geometry.
Suspension A. Stronger, can carry heavier loads than independent suspensions.
D. Less camber during normal driving.
D. Less comfortable and stable ride. Heavy. Not very adjustable.
D. More camber when going over bump or pothole.
Double A-arm A. Independent suspension, with better ride characteristics.
(SLA) A. Better isolation of bumps and shocks than axle suspensions.
A. Fewer oscillations than axle suspensions.
A. More adjustability of suspension parameters and roll center
than axle suspension and MacPherson strut.
A. Lighter than axle suspensions.
D. Heavier and bulkier than MacPherson strut.
MacPherson A. Compact. Weighs less and uses less space than double A-arm.
Strut A. Very suitable for use as front suspension.
A. Better ride characteristics than axle suspensions.
A. Lower cost than double A-arm.
D. Less adjustability of roll center and camber than double A-arm.
D. Worse ride characteristics than double A-arm.
Multilink A. Better handling and ride.
A. Better adjustability of suspension parameters.
D. More expensive and more difficult to design and tune.
D. Bulkier than double A-arm, takes more space.

distributions, which creates discomfort, reduces stability, and increases wear and tear. The
analysis here assumes that the sprung mass and suspension geometry are symmetric.
Consider the rear view of an independent suspension, say, SLA, shown in Figure 15.47.
Link 1 is the grounded link (road, in this case) and the sprung mass is link 8. Links 2 and
3 are the tires and the connection between the tires and the road is modeled as a pin joint.
There are eight components to this mechanism, leading to 8 × 7/2 = 28 instant centers.
Of interest are the instant centers I28 and I38 , which locate the instant centers common to
the sprung mass and wheels, as located in the previous section, and I18 , which is the instant
center common to the ground and the sprung mass. This instant center is the roll center
and it describes the point about which the sprung mass rotates.
Noting that the suspension mechanism has eight links (n = 8) and 10 joints (J = 10),
Gruebler’s equation gives the number of degrees of freedom as

d.o.f. = 3 (n − 1) − 2J = 21 − 20 = 1 (15.99)

We do not need to calculate all 28 instant centers to locate the roll center, I18 . Recall that
instant centers I28 and I38 , which are the instant centers common to the unsprung mass
and the left and right wheels, respectively, were calculated in the previous section.
Noting that instant centers I12 and I13 are identified from the geometry, we can use
the link combinations 1-2-8 and 1-3-8 to locate the roll center I18 . These combinations are
shown in the instant center locating tool in Figure 15.51. From the locating tool it is clear
that we can use several other three link combinations, such as 1-4-8 and 1-5-8, but the wiser
Vehicle Dynamics—Bounce, Pitch, and Roll 819

choice is to make use of the instant centers that have already been calculated. The link
combinations 1-2-8 and 1-3-8 are suitable because two of the instant centers in each set are
known already (I28 and I12 in the linkage set 1-2-8, and I38 and I13 in the linkage set 1-3-8).
These known instant centers are drawn with solid lines in Figure 15.51.

( "

' #

& $
%

FIGURE 15.51
Roll center locating tool.

To locate the roll center, we draw two lines, one going through I28 and I12 and another
going through I38 and I13 . The intersection of these two lines defines I18 , the roll center, as
shown in Figure 15.52.
In Figure 15.52 the roll center is below ground. When the short arm is not very small
compared to the long arm, or when the distance of the arm becomes smaller towards the
sprung mass, as shown in Figure 15.53, the roll center, I18 , can be above ground. Compare
the positions of the instant centers I28 and I38 in Figure 15.53 with Figure 15.52.

!#0 !10
!$# !21
$ # 1 2
%&'()*+ !20
$ 2
!$0 ,-..
0
!$" " / !2/
!"0 !/0
!!$ ! !!2

!!0

FIGURE 15.52
Roll center (below ground) of SLA suspension.

A similar approach, consisting of triplet 1-2-8 and the midpoint of the suspension system,
is used to locate the roll center for the MacPherson strut, as shown in Figure 15.54. We
note that the line joining I28 and I48 is parallel to the line that locates I24 . Unlike double-A
arm suspensions, the geometry of a MacPherson strut does not allow for a wide selection
of roll center locations.
820 Applied Dynamics

$ # 1 &
'()*+,- %
$
!&% ./00 !$%
" 2
!!%
!!$ ! !!&

FIGURE 15.53
Roll center (above ground) of SLA suspension.

I48

4
Sprung I28
mass 5
2
8
I24 2 6 3 3
I18 7
I68
I12 1 I13
Mid point

FIGURE 15.54
Roll center (I18 ) of MacPherson strut.

15.18 Jacking
As we saw earlier, the roll center can be viewed as a coupling point between the unsprung
and sprung mass. If the roll center is too high, the sprung mass rolls little when a lateral
load is applied, but the propensity of the unsprung mass to roll increases. This is a design
criterion in selecting the roll center height.
Another consideration when selecting the roll center height is the horizontal-vertical
coupling that the roll center provides. Consider Figure 15.55a, which shows the left tire
(body 2 according to our convention) and the sprung mass (body 8). Considering links 1,
2, and 8, from Kennedy’s rule, instant centers I12 , I18 , and I28 lie on the same line. Thus,
if the roll center (I18 ) is above ground, as in Figure 15.53, then I28 , which is the reaction
point between the sprung mass and the tire, lies to the right of the tire.
Next consider that there is a lateral force acting on the tire and draw the free-body
diagrams of the tire and sprung mass separately, as shown in Figure 15.55b. The force
FJ = FL h0 /d is not the weight, but the vertical force that is generated as a result of the
coupling of the motion of the tire and sprung mass. As can be seen, the force FJ acts in
the upwards direction on the sprung mass and has a tendency to lift the sprung mass.
This effect is known as jacking. Increasing the height of the roll center increases h0 , which
increases the jacking force. Jacking is not desired, especially for performance vehicles, which
Vehicle Dynamics—Bounce, Pitch, and Roll 821

a) Sprung mass b) Sprung mass


8 FJ 8
I18
RC I28
d
2 I28 2 I28 FL FL
h' FJ
1 I12
I12 FL FL
FJ

FIGURE 15.55
a) Tire and sprung mass, and associated instant centers. b) Free-body diagrams and jacking
force FJ .

is one reason the roll centers of performance vehicles are closer to the ground than those of
passenger vehicles.
The tire to the right of the sprung mass produces the opposite result, so that the sprung
mass experiences a clockwise moment. However, Figure 15.55 is descriptive of a vehicle
taking a right turn (lateral acceleration and lateral force to the right), so the outside wheel
is modeled. Because of weight shift to the outside wheel, the lateral force in the outside
wheel will be higher than the lateral force in the inside wheel and the jacking effect in the
inside wheel will be less.
When the roll center is below the road surface (see Figure 15.52), the moment generated
by the lateral forces about the roll center has a tendency to push the sprung mass down.
In general, performance vehicles, as well as other vehicles to a certain extent, prefer to
have less jacking (requires a low roll center height) and less roll (requires higher roll center)
of the sprung mass. The contradiction in the roll center requirement is usually resolved by
having stronger springs, increasing track width, and installing anti-roll bars.

15.19 Scrub
Consider the four-bar linkage in Figure 15.49, describing a suspension viewed from the rear.
Shown in Figure 15.56 is the motion of a wheel with respect to the sprung mass. By using
the instant center I28 , we can view the motion of a wheel with respect to the sprung mass
as rotating about I28 . This relative motion is in the form of an arc, whose radius is the
distance between I12 and I28 . Consequently, as tires go into jounce and rebound, they will
have some lateral motion with respect to the road. This motion is known as scrub.
The amount of scrub depends on the control arms of the suspension, as well as on the
location of the instant center I28 (and I38 for the right suspension). The lowest amount
of scrub is experienced when the instant center is close to the road surface. Scrub is an
undesirable quantity. On rough roads, where the suspension works hard in jounce and
rebound, scrub will reduce stability by changing slip angles. Because of the lateral motion
of the tires and friction between the tire and the road, scrub also dissipates energy and
results in lower mileage.
822 Applied Dynamics

Undeformed tire

I28

Path of
wheel Deformed tire
wrt I28

I12

FIGURE 15.56
Undeformed and deformed tire and resulting scrub.

Scrub also occurs in the front tires when the driver turns the steering wheel and there
is sliding in the contact patch. This type of scrub is significant at very low speeds. Scrub
results in lateral friction forces and wear and tear in the contact patch, as well as in the
steering mechanism.

15.20 Anti-Roll Bar

) ( '

ș %&
! + '
,
* %

$
‫׋‬
"
!"#$%&'()**

#
+%*"#$%&'()**

FIGURE 15.57
Geometry of anti-roll bar ABCD.

One way of generating additional roll stiffness without increasing suspension stiffness
is by installing anti-roll bars, also known as anti-sway bars. Figure 15.57 describes the
schematic of an anti-roll bar for the front end, which is a rod formed into the shape of a
channel section ABCD (some bars have more elaborate shapes in order to accommodate
Vehicle Dynamics—Bounce, Pitch, and Roll 823

other objects in their vicinity). Points A and D are connected to the unsprung mass, while
the long section, BC, is attached to the sprung mass.
Consider a left turn by the vehicle. Weight shifts to the outside wheels (to the right
wheels) and the unsprung and sprung masses come close to each other at point D. While
we can study this situation as the sprung mass moving down, it is more instructive to
treat point D as moving up relative to the sprung mass, to point D0 , by a distance q. This
situation is also encountered when point D goes over a bump and reaches point D0 . Points
A, B, and C remain stationary.
Arm CD0 is now at an angle of θ = tan−1 (q/l) with its initial position. As a result,
rod BC twists by the angle θ. Further, the line AD0 in Figure 15.57 makes an angle of φ
with the original position, AD, where φ = tan−1 (q/d) is the roll angle that results from the
travel of point D to D0 . Hence, φ denotes the roll of the sprung mass with respect to the
unsprung mass. Roll of the sprung mass with respect to the unsprung mass is clockwise and
the roll of the unsprung mass with respect to the sprung mass is counterclockwise.

y
a) F b)
x y
F
F
T F y D'
C F
T T q
v D'
d x C
l q l D
z
x D z

B c) x
x F
F v Mx=Fd
T C
B y
d
F z

FIGURE 15.58
a) Free-body diagram of anti-roll bar. b) Side view of arm CD. c) Rear view of arm BC.

The next step is to analyze the forces and moments that the anti-roll bar generates as
a result of the movement of point D to D0 . The sprung mass generates a force F at D that
acts downwards. Let us consider point C and analyze the internal forces and moments. The
free-body diagrams are shown in Figure 15.58a and a side view of CD0 is given in Figure
15.58b.
Considering arm BC (Figure 15.58c), the twisting of this arm by θ creates a torque T
opposing the twist and a vertical force F at point C. By contrast, in rod CD0 there is an
equal and opposite force F and bending moment of the same magnitude of the torque T .
Thus, there develops a downward force F at point D0 .
Both angles are small and justify a small angles assumption of
q q
θ ≈ φ ≈ (15.100)
l d
From torsion theory, the twist angle θ is related to the torque T by
Td
θ = (15.101)
GJ
824 Applied Dynamics

in which G is the shear modulus and J is the area polar moment of inertia of the cross
section of the anti-roll bar. It follows that the torque is
GJ q
T = θ = GJ (15.102)
d dl
Considering section CD0 , summing moments about C or D around the y axis, and
considering the above equation give an expression for the force F in terms of the torque as
T q
F = = GJ 2 (15.103)
l dl
The force F causes a moment Mx about the x axis on rod BC and hence on the sprung
mass in the form
q
Mx = F d = GJ 2 (15.104)
l
This moment is counterclockwise, opposing the rotation of the front end, which is clockwise
with respect to the unsprung mass. Noting that the roll angle is φ = q/d, we can express
the moment that is generated in terms of the roll angle as
GJd
Mx = F d = φ = kφ0 φ (15.105)
l2
where kφ0 = GJd/l2 is the additional roll stiffness that the anti-roll bar provides.
Anti-roll bars can be added to front or to rear suspensions. They come in handy when
the suspension springs, because of the constraints they need to satisfy, are not stiff enough
to provide needed roll stiffness. On the other hand, having too stiff an anti-roll bar has
disadvantages. Such an anti-roll bar reduces the effectiveness of independent suspensions,
as the bar relates the motions of the left and right wheels to each other.
A stiff anti-roll bar also changes the lateral load and traction properties. For a vehicle
taking a turn, the outside wheel load increases and the inside wheel load decreases. The
anti-roll bar increases the wheel load of the outside wheel (unsprung mass exerts an upward
force to the sprung mass, and the sprung mass exerts a downward force on the unsprung
mass), with the net effect of reducing the wheel load on the inside wheel even more. As
discussed in Section 15.11, shifting the wheel loads also has the effect of further reducing
lateral forces that can be generated. Load redistribution by the anti-roll bar can also change
the under-/oversteer properties of the vehicle.

15.21 Force Analysis for Anti-Squat and Anti-Dive


This section analyzes the dynamics of squat and dive motions, which take place when
a vehicle accelerates and decelerates. We consider the side view geometry of suspension
systems. The developments in this section follow the excellent discussion on anti-squat and
anti-dive in the text by Gillespie.
We will use the following assumptions for both acceleration and deceleration:
• The location of the instant center common to the sprung mass and the axle, denoted
by IC, does not change with squat and dive motions.
• The suspension links connecting the sprung and unsprung masses are lightweight, so
they can be treated as two force members.
Vehicle Dynamics—Bounce, Pitch, and Roll 825

• Steady-state motion is reached rapidly.


Based on these assumptions, we can use a static model to analyze anti-squat geometry.
Begin with a simple model of an axle with trailing arms(control arm) of a rear wheel drive
vehicle. The free-body diagram of the axle due to suspension forces is shown in Figure 15.59.
Note that here Fz denotes the resultant force carried by the suspension arms and not the
weight of the sprung mass. Also, the free-body diagram is of the axle and not the wheel.

ș!
!!
+,
ș (
&
!"
ș"
'
()
#

* "%
"#$

FIGURE 15.59
Anti-squat free-body diagram of rear axle.

The location of the instant center is described by the parameters e and d, which give
the distance between the instant center and the contact point of the wheel with the road.
Both e and d depend on the lengths and orientations of the trailing arms and the location
of the suspension points to which the trailing arms are connected. Summing forces in the
vertical direction gives
X
+↑ F = 0 =⇒ Fz = P1 sin θ1 + P2 sin θ2 (15.106)

where θ1 and θ2 are the angles the suspension arms make with the horizontal. The tractive
force is related to the loads carried by the suspension arms by
+
X
← F = 0 =⇒ FT = −P1 cos θ1 + P2 cos θ2 (15.107)

It follows from the above equation that the lower arm carries a much higher load than
the upper arm. From the free-body diagram of the sprung mass, the force P2 propels the
vehicle forward and the upper arm force, P1 , pushes back, thereby creating the moment
that makes the vehicle pitch up (or squat).
The sum of moments about the instant center gives the relationship between the vertical
and horizontal forces. Recall the assumption that the location of the instant center does not
change, or that it changes little. Because under such circumstances we can take moments
about the instant center, the instant center is also referred to as virtual reaction point or
virtual pivot point. Summing moments about the instant center gives

 MIC = 0 =⇒ FT e − Fz d = 0 (15.108)

which leads to a relationship between the vertical and tractive force as


Fz e
= (15.109)
FT d
826 Applied Dynamics

so that the vertical force generated by the suspension is completely dependent on the sus-
pension geometry. This means that, by properly designing the trailing arm lengths and
suspension points, we can specify the forces generated by the suspension system.

ș!
"#
!!
ș
' &
!"
ș"
)
*+
(

, $&
$%

FIGURE 15.60
Free-body diagram considering wheel loads.

The physical description of the vertical load Fz is of interest. To this end, we redraw in
Figure 15.60 the free-body diagram of the axle, now considering the wheel loads and where
FN denotes the wheel load and W 0 is the axle force, so that Fz = FN − W 0 . The axle load
W 0 consists of the weight (static load) Wrs plus the force associated with the suspension
stiffness, given by kr zr , where kr is the rear suspension stiffness and zr is the deflection of
the rear suspension. Hence,

W 0 = Wrs + kr zr (15.110)

Recall that zr is measured from the static equilibrium position. The normal force that
acts on the wheel, FN , is due to the static load Wrs plus the weight shift due to acceleration,
W hG
gL ax , as given in Equation (12.5). Hence,

W hG
FN = W r s + ax (15.111)
gL
In the solid axle, the drive torque is transmitted through an inner shaft that connects
the differential to the wheel (half-shaft) so the axle is not affected by the drive torque (the
torque T in Figure 15.60 enters the formulation when considering independent suspensions).
The suspension load Fz is the difference between the two vertical loads,
W hG W hG
Fz = FN − W 0 = Wrs + ax − (Wrs + kr zr ) = ax − kr zr (15.112)
gL gL
Fz can be interpreted as the resultant of the weight shift due to the longitudinal ac-
celeration and deflection of the suspension stiffness. Noting from Equation (15.109) the
relationship between Fz and the tractive force FT , and that the tractive force is FT = Wg ax ,
we can write the change in load in the rear axle due to the suspension stiffness as
 
W hG e W hG e
∆Wr = kr zr = ax − FT = ax − (15.113)
gL d g L d
The above equation determines how much the rear suspension will deflect. For example,
Vehicle Dynamics—Bounce, Pitch, and Roll 827

if it is desired that rear suspension should not deflect at all due to acceleration (zr = 0),
the suspension can be designed so that
e hG
= (15.114)
d L
Considering a ratio of hLG ≈ 0.2 for a passenger vehicle, for no deflection of the rear axle
we would need d ≈ 5e, which would require a long trailing arm. Of course, we would not
want to design a vehicle where the rear suspension would not deflect due to acceleration (or
deceleration).

v
x B L A
zf
zr
z

FIGURE 15.61
Pitch angle.

The pitch angle θ, shown in Figure 15.61, can be obtained from the difference of the
deflection of the front and rear axles
zr − zf
θ ≈ tan θ = (15.115)
L
where zf is the deflection of the front suspension. Considering a rear wheel drive vehicle
and conducting an analysis similar to the rear wheels, the front suspension deformation can
be shown to be due only to the weight shift
W hG
kf zf = − ax (15.116)
gL
and the pitch angle becomes
   
zr − zf W hG 1 1 e
θ = = ax + − (15.117)
L gL2 L kr kf kr d

For zero pitch, that is, full anti-squat, the pitch angle is set to zero in the above equation,
with the result
 
e hG kr
= 1+ (15.118)
d L kf
hG kr e
As an illustration, setting L ≈ 0.2 and kf ≈ 0.7 gives d ≈ 0.34, which still requires a long
trailing arm.
828 Applied Dynamics

15.21.1 Independent Suspensions and Rear Wheel Drive


With independent suspensions, the drive torque does affect the suspension system, and the
torque T needs to be included in the free-body diagram in Figure 15.60. Adding to the
formulation the counterclockwise torque of magnitude T = FT Rl , where Rl is the loaded
radius, the sum of moments about the instant center gives

 MIC = 0 = −Fz d + FT e − T = −Fz d + FT (e − Rl ) (15.119)

Hence, the equations in the previous subsection can be used to calculate the pitch angle by
replacing e with e − Rl in Equation (15.117), which gives
   
zr − zf W hG 1 1 e − Rl
θ = = ax + − (15.120)
L gL2 L kr kf kr d
and for zero pitch angle we need
 
e − Rl hG kr
= 1+ (15.121)
d L kf

15.21.2 Front Wheel Drive

"#
!! -.
'
!"
)
*+
( &

$& ,
$%

FIGURE 15.62
Free-body diagram for front wheel drive.

The analysis of a front wheel drive vehicle is similar to the rear wheel drive analysis.
The free-body diagram is given in Figure 15.62. In this case,
W hG
FN = Wfs − ax W 0 = Wfs + kf zf (15.122)
gL
Considering independent suspensions and the free-body diagram in Figure 15.62, we can
calculate the suspension load Fz as
W hG W hG
Fz = FN − W 0 = Wfs − ax − (Wfs + kf zf ) = − ax − kf zf (15.123)
gL gL
Summing moments about the instant center (assuming, as before, that the instant center
moves little), and noting that the drive torque is T = FT Rl , results in

 MIC = 0 = Fz d + FT e − T = Fz d + FT (e − Rl ) (15.124)
Vehicle Dynamics—Bounce, Pitch, and Roll 829

Assuming that e−Rl is greater than zero, and d is positive and so is the tractive force FT ,
the above equation implies that Fz must be negative. This result follows from an inspection
of the expression for Fz in Equation (15.123). It follows that, since the change in vertical
load that acts on the front suspension is ∆Wf = kf zf , we have
 
W hG W hG e W hG e
∆Wf = − ax − Fz = − ax + FT = − ax − (15.125)
gL gL d g L d
and the pitch angle becomes
   
zr − zf W hG 1 1 e − Rl
θ = = ax + − (15.126)
L gL2 L kf kr kf d
which is the same relationship that was obtained for the rear wheel drive case, except that
kr is replaced with kf and vice versa.

15.21.3 Four Wheel Drive


Four wheel drive analysis is a combination of the front and rear wheel drive analyses and
it is accomplished by distributing the tractive force between the front and rear wheels.
Denoting by η a proportioning factor for the tractive force so that

FTf = ηFT FTr = (1 − η) FT (15.127)

the pitch angle is obtained as


   
W hG 1 1 er − Rl ef − Rl
θ = ax + + (1 − η) −η (15.128)
gL2 L kf kr kr dr kf df
with df , ef , dr , and er denoting the location of the front and rear instant centers, respec-
tively.

15.21.4 Anti-Dive for Braking


The relationships for anti-dive are the same as the relationships for four wheel drive, except
that we set Rl = 0, as the brake torque is generated by the wheel and does not affect the
suspension geometry. Also, FT is replaced by FBr which acts in the opposite direction, with
FBr denoting the brake force acting on the rear axle. Let the braking forces on the front
and rear be related to the total braking force FB by a proportioning factor η, so

FBf = ηFB FBr = FBr = (1 − η) FB (15.129)

where
W
FB = FB f + ax (15.130)
g
is the total braking force. Consider the free-body diagram of the rear wheel in Figure 15.63
and sum moments about the instant center, with the result

 MIC = 0 =⇒ −FN d + W 0 d − FBr e = 0 (15.131)

where
hG
FN = Wfs − W Dx (15.132)
L
830 Applied Dynamics

ș! "#
!!
,-
'
!"
)
ș" *+
(

. $&
$%

FIGURE 15.63
Free-body diagram of rear wheel during braking.

in which Dx = −ax /g denotes the deceleration rate. Also, W 0 = Wrs + kr zr . Substituting


these terms in the above equation, noting that

FBr = (1 − η) W Dx (15.133)

and solving for the spring force in the suspension system result in
 
e W hG W e hG
kr zr = FBr − Dx = Dx (1 − η) − (15.134)
d L g d L

which is the negative of Equation (15.126) and accounts for the brake proportioning.
A similar result is obtained for the front wheel and the expression for the pitch angle
becomes exactly the negative of Equation (15.128),
   
W hG 1 1 er ef
θ = − 2 Dx + + (1 − η) −η (15.135)
L L kf kr kr dr kf df

We can calculate the suspension geometry required for zero pitch angle or for no dive
in the front or rear suspensions from the above equations. The results for zero dive are
ef hG er hG
= for front = for rear (15.136)
df ηL dr (1 − η) L

While from a geometric perspective zero dive (or 100% anti-dive) looks attractive, it is
not desirable for a variety of factors, such as human response to ride, as well as oversteer
problems. In general, vehicle designers prefer about 50% anti-dive.

15.22 Bibliography
Automobile Ride, Handling, and Suspension Design, http://www.rqriley.com/suspensn.htm
Genta, G., Motor Vehicle Dynamics: Modeling and Simulation, World Scientific, 1997.
Gillespie, T.D., Fundamentals of Vehicle Dynamics, SAE Publications (R114), 1992.
Griffin, M.J., Human Response to Vibration, Academic Press, 1996.
Vehicle Dynamics—Bounce, Pitch, and Roll 831

International Organization for Standardization, Mechanical Vibration and Shock - Evalu-


ation of Human Exposure to Whole-Body Vibration - Part 1: General Requirements, ISO
standard 2631-1, 1997.
Jazar, R.N., Vehicle Dynamics: Theory and Application, 2nd Edition, Springer, 2013.
Mansfield, N.J., Human Response to Vibration, CRC Press, 2005.
Milliken, W.L., and Milliken, D.F., Race Car Vehicle Dynamics, SAE Publications (R146),
1995.
Norton, R.L., Design of Machinery: An Introduction to the Synthesis and Analysis of Mech-
anisms and Machines, 4th Edition, McGraw-Hill, 2008.
Smith, C., Tune to Win, Aero Publishers, 1978.
Staniforth, A., Competition Car Suspension: A Practical Handbook, 4th Edition, Haynes
Publishing, 2006.

15.23 Problems
Problems are marked by E—easy, M—moderate, and D—difficult. Computer-oriented prob-
lems are marked by C.

Section 15.5—Quarter-Car Model


15.1 (M) Consider the quarter-car with mS = 300 kg, mU = 40 kg, kS = 15 N/mm, kt = 160
N/mm. Calculate the frequencies associated with the two simplified models (ride, wheel hop)
and compare with the actual frequencies of the two-degrees-of-freedom model. Calculate the
modal vectors and comment on the ride isolation. Calculate the static deflections.
15.2 (C) Consider the quarter-car in the previous problem. The sprung mass is increased
by x kg, while the unsprung mass is reduced by x kg, so the total mass remains the same.
Calculate the new frequencies and plot them as a function
of x, where −20 ≤ x ≤ 20 kg.
u11 u22
Also develop a ride isolation parameter RI = u12 u21 and plot it as a function of x.
15.3 (M) Consider the quarter-car in Problem 15.1 with cS = 1.4 N·s/mm. Calculate the
damping factors using the approximate approach outlined in Section 7.12.
15.4 (M) Show in closed-form that the natural frequencies of the quarter-car model are
quite separate from each other when mS /mU > 6 and kt /kS > 10. Formulate the eigenvalue
problem in terms of these ratios and solve the characteristic equation by hand.
15.5 (C) Consider the quarter-car in Example 15.3. Now, in increments of 5 kg, take mass
from the sprung mass and add it to the unsprung mass, so that the total mass remains
constant. Calculate and plot the frequencies of the quarter-car until the unsprung mass
becomes 100 kg.
15.6 (D) You are asked to design a quarter-car suspension system for a sports car. The
quarter-car weight is WS = 600 lb and WU = 80 lb. You are to design the spring constants
so that the quarter-car has the following frequencies: f1 = 1.4 Hz, f2 = 11 Hz. Calculate
the values for kS and kt . Hint: Obtain a general expression for the frequencies in terms of
the spring constants and invert the expressions.

Sections 15.6 and 15.7—Pitch and Bounce Motions, Olley Criteria


15.7 (E) A vehicle has the following properties: wheelbase L = 3.1 m, axle loads mf = 760
kg, mr = 690 kg, front and rear axle ride rates kr = 42 N/mm, kf = 36 N/mm, pitch
832 Applied Dynamics

moment of inertia Iθ = 2600 kg·m2 . Find the static deflections of the center of mass and
pitch angle, as well as the dynamic index DI.
15.8 (M) Calculate the frequencies, mode shapes, and oscillation centers of the vehicle in
Problem 15.7. Check if the Olley criteria are satisfied for pitch and bounce.
15.9 (M) Consider the pitch-bounce model and a vehicle with the following parameters:
kr = 145 lb/in., kf = 111 lb/in., Wf = 1400 lb, Wr = 1100 lb, wheelbase L = 110
in., DI = 0.95. Calculate the following: a) the static equilibrium position, b) the natural
frequencies and modal vectors, c) locations of the oscillation centers, d) whether the Olley
criteria are satisfied?
15.10 (C) A vehicle has the following properties: wheelbase L = 9.25 ft, static weights
Wf = 1650 lb, Wr = 1450 lb, dynamic index DI = 0.92. Considering axle spring rates as
variables, plot the two frequencies f1 and f2 as functions of the spring rates kr and kf .
Identify the ranges for kr and kf where the conditions 0.9 < f1 < 1.1 Hz, 1.2 < f2 < 1.35
Hz are satisfied.
15.11 (C) Consider Example 15.5 and calculate the locations of the oscillation centers as a
function of the rear axle weight. Vary the real axle weight Wr from 0.35W to 0.65W . Plot
the locations of the centers of oscillation as a function of Wr .
15.12 (M) The pitch-bounce model of a vehicle has the following properties: Wheelbase
= 2.7 m, mass = 1350 kg, mr = 540 kg, DI = 0.9. An incompetent engineer decides to
select the suspension stiffnesses so that the pitch and bounce equations will be completely
uncoupled. The desired bounce frequency is 1.1 Hz. What spring rates does the engineer
have to use to accomplish this misguided goal? What is the resulting pitch frequency?
15.13 (D) Consider the pitch and bounce model in Figure 15.4b, which takes into consider-
ation the tire stiffnesses. Using z, θ, zr , and zf as the generalized coordinates, obtain the
equations of motion of this four-degrees-of-freedom system. Linearize the equations of mo-
tion assuming small motions, and write the equations of motion in matrix form. The center
of mass is at a distance c from the rear axle and the mass and centroidal mass moment of
inertia are m and Iφ , respectively.

Section 15.8—Response to Harmonic Excitation


15.14 (C) Consider Problem 15.7. The front wheels are subjected to a harmonic excitation
F eiωt . Plot the steady-state amplitude of the center of mass height zG as a function of the
excitation frequency ω.
15.15 (C) Consider Example 15.7. Follow the same procedure (including the adding of
proportional damping to the system) and plot the frequency response of zG (t) and θ(t), but
now using the vehicle parameters in Example 15.4. Compare the results with Example 15.7.

Section 15.9—Roll Dynamics


15.16 (E) A vehicle with wheelbase L = 9 ft, wheel loads Wf = 1750 lb, Wr = 1450 lb,
DI = 0.9, track t = 5 ft, s = 58 in., and axle stiffnesses kf = 200 lb/in. and kr = 250 lb/in.
Calculate the roll moment of inertia Iφ and radius of gyration κ, given that the roll natural
frequency is ωφ = 7.2 rad/s.
15.17 (E) In Problem 15.16, calculate the damping coefficient for the front tires that would
give a roll damping factor of ζφ = 0.3. The front dampers have constants that are 20%
larger than the rear dampers.
15.18 (M) An axle has wheel load W = 9000 N, s = 1.2 m, and axle stiffness k = 30, 000
N/m. The roll natural frequency is ωφ = 7.3 rad/s. The center of mass is at a height of 75
cm. Calculate the roll moment of inertia for when a) the roll center is exactly on the center
of mass, and b) the roll center is 15 cm below the center of mass.
Vehicle Dynamics—Bounce, Pitch, and Roll 833

Section 15.10—Roll Center Analysis


15.19 (E) What would happen if the roll center height is selected the same as the center of
mass height? How about when roll center is selected higher than the center of mass?
15.20 (M) An axle has wheel load W = 9000 N, s = 1.2 m, and axle stiffness k = 30, 000
N/m. The center of mass is at a height of 75 cm. Given that the track is t = 1.32 m and
that the axle is subject to a lateral acceleration of 0.2g, calculate the roll angle and the shift
in the wheel loads from the inside to the outside wheel. The roll center is at a distance of
15 cm below the center of mass.
15.21 (C) A rear axle weighs 1500 lb and has a roll moment of inertia of Iφ = 220 slug·ft2 .
The suspension springs of kS = 125 lb/in. are 62 in. apart. The center of mass height is 21
in. Calculate and plot the natural frequency (Equation (15.75)) in terms of the roll center
height h. Then, generate a three-dimensional plot of the wheel load shift as a function of
the distance h and lateral acceleration of the vehicle ay , so Fy = may .
15.22 (M) Consider Problem 15.16, and calculate the roll angles of the front and rear axles,
as well as the lateral weight shift in each axle, given that the roll center heights are hRCf = 3
in. and hRCr = 12 in. The center of mass height is hG = 30 in. The vehicle is traveling at
a speed of 45 mph while taking a turn of radius ρ = 500 ft.

Section 15.11—Lateral Force Reduction Due to Weight Shift


15.23 (M) Consider the vehicle in Problems 15.16 and 15.22 and calculate the reduction in
the lateral load in each axle and the change in the understeer gradient, given axle cornering
stiffnesses Cf = Cr = −300 lb/deg, track t = 62 in., s = 58 in., and nonlinearity factor
e2 = 0.03.

Section 15.12—Roll Axis


15.24 (M) A vehicle has an independent front suspension of roll stiffness kφf = 140 N·m/deg
and a solid rear axle with leaf springs. The leaf springs have a stiffness of 20 N/mm each
and they are separated by 135 cm. a) Calculate the rear suspension roll stiffness. b) The
sprung mass is 1100 kg and its center of mass is at a height of 20 cm above the roll axis.
Find the roll gradient for the vehicle and how much the sprung mass would roll while taking
a turn of radius 80 ft at a speed of 70 kph.

Section 15.17—Roll Center Construction


15.25 (M) Consider the double A-arm suspension systems in Figure 15.64 and graphically
determine the location of the roll centers.
15.26 (M) Consider the double A-arm suspension systems in Figure 15.65. Given the desired
location of the roll center, graphically construct the rest of the linkage. Note: There are
multiple solutions to this problem and for both parts.

Section 15.20—Anti-Roll Bar


15.27 (M) Consider an axle suspension of track t = 1.4 m, s = 1.2 m, and two suspension
springs of kS = 150 N/cm. It is desired to increase the roll stiffness by 20% by an anti-roll
bar. The arms of the anti-roll bar are l = 0.2 m. Design the anti-roll bar as a hollow bar of
cold-rolled steel. The length of the roll bar has to be less than s.
834 Applied Dynamics

*&

# (
+,-./01
$ 2*33 )
4
" '

%&

# (
+,-./01
$ )
2*33 4
" '

FIGURE 15.64
Figure for Problem 15.25.

#$

!!"

%$

!!"

FIGURE 15.65
Figure for Problem 15.26.
16
Appendix—Common Inertia Properties

Note: xyz always denote centroidal coordinates.

Circular Cylinder
Volume = πR2 L
1
Izz = 2 mR2
1 1
Ixx = Iyy = 4 mR2 + 12 mL2
1
Ix0 x0 = Iy0 y0 = 4 mR2 + 13 mL2

Thin Disk (L/R very small)


Volume = πR2 L
1
Izz = 2 mR2
1
Ixx = Iyy = 4 mR2

Slender Rod (R/L very small)


Volume = AL
Izz ≈ 0
1
Ixx = Iyy = 12 mL2
1
Ix0 x0 = Iy0 y0 = 3 mL2

Semicylinder
Volume = 12 πR2 L
9π 2 −64 1
Ixx = 36π 2 mR2 + 12 mL2
1 1
Iyy = Ix0 x0 = 4 mR2 + 12 mL2
2
Izz = 9π18π−32
2 mR2
Iz0 z0 = 12 mR2

835
836 Applied Dynamics
Rectangular Prism
Volume = abc
1

Ixx = 12 m b2 + c 2
1

Iyy = 12 m a2 + c2
1

Izz = 12 m a2 + b2
1

Ix0 x0 = 12 m b2 + 4c2
1

Ix0 x0 = 12 m a2 + 4c2

Thin Plate (One of a, b, or c small. Say, c)


Volume = abc
1 2
Ixx = 12 mb
1 2
Iyy = 12 ma
1 2

Izz = 12 m a + b2

Sphere
4
Volume = 3 πR3
2
Ixx = Iyy = Izz = 5 mR2

Hemisphere
Volume = 23 πR3
83
Ixx = Iyy = 320 mR2
2 2
Izz = 5 mR
2
Ix0 x0 = Iy0 y0 = 5 mR2

Right Circular Cone


Volume = 13 πR2 L
3 3
Ixx = Iyy = 20 mR2 + 80 mL2
3
Izz = 10 R2
3
Ix0 x0 = Iy0 y0 = 20 mR2 + 35 mL2
Appendix—Common Inertia Properties 837
Half Cone
Volume = 16 πR2 L
3
− π12 mR2 = 3
 2
Ixx = 20 80 mL
3 3
Iyy = 20 mR2 + 80 mL2
3
− π12 mR2 , Iz0 z0 = 10
3

Izz = 10 mR2
1
Iyz = − 20π mRL, Ixy = Ixz = 0
3 1
Ix0 x0 = Iy0 y0 = 20 mR2 + 10 mL2

Ellipsoid
4
Volume = 3 πabc
1

Ixx = 5m b + c2
2

1

Iyy = 5m a 2 + c2
1

Izz = 5m a2 + b2
x2 y2 z2
Surface defined by a2 + b2 + c2 =1

Right Triangular Prism


Volume = 12 abc
1 2 1 2
Ixx = 18 mb + 12 mc
1 2 1 2
Iyy = 18 ma + 12 mc
1 2 2

Izz = 18 m a + b
1
Ixy = − 36 mab Ixz = Iyz = 0

Triangular Plate (c << a, c << b)


Volume = 12 abc
1 2
Ixx = 18 mb
1 2
Iyy = 18 ma
1 2

Izz = 18 m a + b2

Half Torus
Volume = πa2 R
1
Ix0 x0 = Iy0 y0 = 2 mR2 + 58 ma2
3
Iz0 z0 = mR2 + ma2 4
838 Applied Dynamics
Half Circular Rod (a/R small)
Volume = πRA
Ixx = m 12 − π42 R2

1
Iyy = 2 mR2
4

Izz = m 1 − π2 R2
1 2
Ix0 x0 = 2 mR Iz0 z0 = mR2

Circular Cylindrical Shell


Ixx = Iyy = 12 mR2 + 12
1
mL2
Izz = mR2
1
Ix0 x0 = Iy0 y0 = 2 mR2 + 13 mL2

Half Cylindrical Shell


Ixx = 12 − π42 mR2 + 12
1

mL2
Izz = 1 − π42 mR2

1 1
Ix0 x0 = Iy0 y0 = 2 mR2 + 12 mL2
Iz0 z0 = mR2

Spherical Shell
2
Ixx = Iyy = Izz = 3 mR2

Hemispherical Shell
5
Ixx = Iyy = 12 mR2
2
Izz = 3 mR2
2
Ix0 x0 = Iy0 y0 = 3 mR2
Mechanical Engineering
“Overall, this is an excellent book and highly recommended. The coverage of the topics is wide
ranging, which makes it suitable for both undergraduate and graduate courses on dynamics. Baruh

APPLIED
What makes this book truly different from the rest are the applications of the dynamics prin-
ciples to real-world systems, such as vibrating systems and vehicles.”
—Ilhan Tuzcu, California State University, Sacramento, USA

APPLIED DYNAMICS
“The combination of applications with theory without compromising either one is excellently

DYNAMICS
done! Also, the unified and fresh approach to dynamics is excellent. …This book is like a breath
of fresh air…”
—Sorin Siegler, Drexel University, Philadelphia, Pennsylvania, USA

GAIN A GREATER UNDERSTANDING OF HOW KEY COMPONENTS WORK

Using realistic examples from everyday life, including sports (motion of balls in the air or during
impact) and vehicle motions, Applied Dynamics emphasizes the applications of dynamics in en-
gineering without sacrificing the fundamentals or rigor. The text provides a detailed analysis of the
principles of dynamics and vehicle motions analysis. An example included in the topic of collisions
is the famous “Immaculate Reception,” whose 40th anniversary was recently celebrated by the
Pittsburgh Steelers.

COVERS STABILITY AND RESPONSE ANALYSIS IN DEPTH

The book addresses two- and three-dimensional Newtonian mechanics, it covers analytical me-
chanics, and describes Lagrange’s and Kane’s equations. It also examines stability and response
analysis, and vibrations of dynamical systems. In addition, the text highlights a developing interest
in the industry—the dynamics and stability of land vehicles.

CONTAINS LOTS OF ILLUSTRATIVE EXAMPLES

In addition to the detailed coverage of dynamics applications, over 180 examples and nearly 600
problems richly illustrate the concepts developed in the text.

TOPICS COVERED INCLUDE


• General kinematics and kinetics
• Expanded study of two- and three-dimensional motion, as well as of impact dynamics
• Analytical mechanics, including Lagrange’s and Kane’s equations
• The stability and response of dynamical systems, including vibration analysis
• Dynamics and stability of ground vehicles

Designed for classroom instruction appealing to undergraduate and graduate students taking
intermediate and advanced dynamics courses, as well as vibration study and analysis of land ve-
hicles, Applied Dynamics can also be used as an up-to-date reference in engineering dynamics for

Haim Baruh
researchers and professional engineers.
K23786
6000 Broken Sound Parkway, NW
Suite 300, Boca Raton, FL 33487
711 Third Avenue
an informa business New York, NY 10017
2 Park Square, Milton Park
www.taylorandfrancisgroup.com Abingdon, Oxon OX14 4RN, UK

You might also like