introduction to dynamics

222
123 Friedrich Pfeiffer Thorsten Schindler Introduction to Dynamics

Upload: ingenieria-mecanica-en-red

Post on 02-Aug-2015

77 views

Category:

Engineering


5 download

TRANSCRIPT

Page 1: Introduction To Dynamics

123

Friedrich PfeifferThorsten Schindler

Introduction toDynamics

Page 2: Introduction To Dynamics

Introduction to Dynamics

Page 3: Introduction To Dynamics

Friedrich Pfeiffer · Thorsten Schindler

Introduction to Dynamics

ABC

Page 4: Introduction To Dynamics

Friedrich PfeifferInstitute of Applied MechanicsTechnische Universität MünchenGarchingGermany

Thorsten SchindlerInstitute of Applied MechanicsTechnische Universität MünchenGarchingGermany

ISBN 978-3-662-46720-6 ISBN 978-3-662-46721-3 (eBook)DOI 10.1007/978-3-662-46721-3

Library of Congress Control Number: 2015934938

Springer Heidelberg New York Dordrecht Londonc© Springer-Verlag Berlin Heidelberg 2015

This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part ofthe material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,broadcasting, reproduction on microfilms or in any other physical way, and transmission or informationstorage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodologynow known or hereafter developed.The use of general descriptive names, registered names, trademarks, service marks, etc. in this publicationdoes not imply, even in the absence of a specific statement, that such names are exempt from the relevantprotective laws and regulations and therefore free for general use.The publisher, the authors and the editors are safe to assume that the advice and information in this bookare believed to be true and accurate at the date of publication. Neither the publisher nor the authors orthe editors give a warranty, express or implied, with respect to the material contained herein or for anyerrors or omissions that may have been made.

Printed on acid-free paper

Springer-Verlag GmbH Berlin Heidelberg is part of Springer Science+Business Media(www.springer.com)

Page 5: Introduction To Dynamics

Preface

As a fundamental science in both physics and engineering, mechanics deals withinteractions of forces resulting in motion and deformation of material bodies. Me-chanics serves in the world of physics and in that of engineering in a particularway, in spite of many increasing interdependencies. For physicists, machines andmechanisms are tools for cognition and research. For engineers, they are the objec-tives of research, according to a famous statement by the Frankfurt physicist andbiologist Friedrich Dessauer. Physicists apply machines to support their questionsto nature with the goal of new insights into the physical world. Engineers applyphysical knowledge to support the realization process of their ideas and their intu-ition. Physics is an analytic science searching for answers to questions concerningthe world around. Engineering is a synthetic science, where the physical and math-ematical fundamentals play the role of reinsurance with respect to a working andefficiently operating machine. Engineering is an iterative science typically resultingin long development periods of its products. However, it is also iterative concern-ing fast improvements of an existing product or fast development of a new one.Every physical or mathematical science has to face these properties by formulatingnew specific methods, new practically approved algorithms up to new fundamen-tals, which are adaptable to new technological developments. This is also true forthe field of mechanics.

With the increasing complexity of technical products, we need more sophis-ticated design methods. Therefore, we have to systematically apply all modernoptions of theoretical and experimental modeling to achieve optimal realization pro-cesses. Product development alone by intuitive design and experimental tests is un-reasonably costly. Theoretical simulations replace more and more experiments, or atleast they reduce tests by organizing them in a sensible way. This definitely requiresgood models, which must be close to reality. Realism needs awareness of the mostimportant aspects, which finally decide the size and quality of models and theory.

This book gives an introduction to dynamics and tries to consider the above rec-ommendations. Thereby, a compromise has to be found between basics and appli-cations. The fundamental principles and laws of kinematics and kinetics are treatedin a general way and can be applied to more complex cases. Furthermore, they are

Page 6: Introduction To Dynamics

VI Preface

transparent and clear enough to be application-friendly. We start with these con-siderations at the very beginning. Afterwards, we treat a few selected problems oflinear and nonlinear dynamics, and finally, we present a more phenomenologicallyoriented chapter on problems of vibration formation. Obviously such an introduc-tion to dynamics can only present a selection of topics, which nevertheless shouldgive a useful basis for stepping into more advanced problems of dynamics. The se-lection in this book represents the result of a regularly revised course, which hasbeen and still is given for masters students at the Technische Universität München.

We owe many conceptional items together with some ideas of conceptual de-tails to Kurt Magnus, who founded the chair B of Mechanics at the MechanicalEngineering Department of the Technische Universität München with a focus ondynamics in 1966. Friedrich Pfeiffer, his successor, continued the course’s evolu-tion by including his significant industrial experience considering mainly problemsconnected with theory and practice. Heinz Ulbrich adapted the course to include hisview of mechanics, and finally, we have input from the younger generation repre-sented by Thorsten Schindler who teaches dynamics to students. We thank DanielRixen, the present chair, for the possibility to elaborate this version of the book.

Generations of assistants and students have worked for this course, the one orother presenting the course him/herself thus helping to improve it. Our warmestregards to them, especially for continual discussions of contents and formulations.Thanks also to the former Teubner Verlag and to the Springer Publishing House forthe excellent problem-free cooperation.

Munich, February 2015 Friedrich Pfeiffer, Thorsten Schindler

Page 7: Introduction To Dynamics

Contents

1 Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.3 Basic Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3.1 Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.3.2 Euler’s Cut Principle and Forces . . . . . . . . . . . . . . . . . . . . . . 41.3.3 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.3.4 Virtual Displacements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.4 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101.4.1 Coordinate Systems and Coordinates . . . . . . . . . . . . . . . . . . 101.4.2 Transformation of Coordinates . . . . . . . . . . . . . . . . . . . . . . . 151.4.3 Relative Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.5 Momentum and Moment of Momentum . . . . . . . . . . . . . . . . . . . . . . 221.5.1 General Axioms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221.5.2 Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231.5.3 Moment of Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

1.6 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251.7 Principles of d’Alembert and Jourdain . . . . . . . . . . . . . . . . . . . . . . . 27

1.7.1 Significance of Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . 271.7.2 Principle of d’Alembert . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281.7.3 Principle of Jourdain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

1.8 Newton-Euler Equations for Constrained Systems . . . . . . . . . . . . . 331.8.1 Single Rigid Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331.8.2 System with Multiple Rigid Bodies . . . . . . . . . . . . . . . . . . . 351.8.3 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

1.9 Lagrange’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401.9.1 Lagrange’s Equations of the First Kind . . . . . . . . . . . . . . . . 401.9.2 Lagrange’s Equations of the Second Kind . . . . . . . . . . . . . . 44

1.10 Hamilton’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551.10.1 Hamilton’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551.10.2 Hamilton’s Canonical Equations . . . . . . . . . . . . . . . . . . . . . . 57

Page 8: Introduction To Dynamics

VIII Contents

1.11 Practical Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

2 Linear Discrete Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692.1 Linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692.2 Classification of Linear Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 712.3 Solution Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

2.3.1 Linear Second-Order Systems . . . . . . . . . . . . . . . . . . . . . . . . 772.3.2 Linear First-Order Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 83

2.4 Stability of Linear Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982.4.1 Criteria Based on the Characteristic Polynomial . . . . . . . . . 982.4.2 Stability of Mechanical Systems . . . . . . . . . . . . . . . . . . . . . . 101

3 Linear Continuous Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1033.1 Models of Continuous Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . 1033.2 Simple Examples of Continuous Vibrations . . . . . . . . . . . . . . . . . . . 104

3.2.1 Beam as a Bending Vibrator . . . . . . . . . . . . . . . . . . . . . . . . . 1043.2.2 Beam as a Bending Vibrator with an End Mass . . . . . . . . . 1093.2.3 Beam as a Torsional Vibrator with an End Mass . . . . . . . . . 1123.2.4 Transverse Vibrations of a String . . . . . . . . . . . . . . . . . . . . . 115

3.3 Approximation of Continuous Vibration Systems . . . . . . . . . . . . . . 1173.3.1 Function Systems and Completeness . . . . . . . . . . . . . . . . . . 1173.3.2 Rayleigh-Ritz Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1193.3.3 Bubnov-Galerkin Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 1233.3.4 Boundary Conditions for the Rayleigh-Ritz and

Bubnov-Galerkin Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 1273.3.5 Choice of Trial Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 1293.3.6 Bending Vibrations of a Beam with Longitudinal Load . . . 129

3.4 Vibrations of Elastic Multibody Systems . . . . . . . . . . . . . . . . . . . . . 135

4 Methods for Nonlinear Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1374.1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1374.2 Phase Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1384.3 A 1-DOF Nonlinear Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

4.3.1 Piecewise Exact Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1434.3.2 Method of Weighted Residuals . . . . . . . . . . . . . . . . . . . . . . . 1444.3.3 Harmonic Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1464.3.4 Method of Least Squares . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1474.3.5 Practical Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

4.4 Stability of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1514.4.1 General Stability Definitions . . . . . . . . . . . . . . . . . . . . . . . . . 1524.4.2 Linear Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1554.4.3 Stability of Nonlinear Systems . . . . . . . . . . . . . . . . . . . . . . . 156

Page 9: Introduction To Dynamics

Contents IX

5 Vibration Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1655.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1655.2 Free Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1675.3 Forced Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1705.4 Self-excited Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

5.4.1 Hydraulic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1775.4.2 Drinking Bird . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1795.4.3 Woodpecker Toy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1805.4.4 Friction Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1815.4.5 Kármán Vortex Street . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1855.4.6 Flutter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1875.4.7 Pendulum Clock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

5.5 Parametrically Excited Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . 1915.5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1915.5.2 Motion and Stability of Parametrically Excited Vibrations 1925.5.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211

Page 10: Introduction To Dynamics

Chapter 1Basics

1.1 Introduction

“Mechanics is the science of motion; we define as its task: to describe completelyand in the simplest possible manner such motions as occur in nature.” With respectto engineering we should complete this statement by “as occur in nature and intechnology.” This more than a hundred-year-old statement was made by KIRCH-HOFF [34] and has lost neither its meaningfulness nor its assertion. Technical me-chanics as a science must also be as simple as possible but conversely descriptivelycomplete. If we consider motion as any kind of translation and rotation, even if onlyminimal as in the case of deformations, and also include the state of no motion(i.e., the state of rest), then motion describes mechanics as a whole. It comprisestwo fundamental aspects, that of geometry and kinematics describing positions andorientations, velocities and accelerations, and that of kinetics, describing the causeof motion. Regarding all possible interactions between material bodies or betweenmaterial bodies and their environment, we consider those possibilities, which pro-duce accelerations (or deformations) of these bodies. We call the driving magnitudeof such interactions forces. Thus, the kinematics of bodies and their interaction withforces, their statics and kinetics, define mechanics.

From a physical point of view, classical mechanics might be a self-contained areaof science, the laws of which can be derived completely in a deductive way. Classicalmechanics can be reduced to some fundamental axioms if we know the laws of forceinteractions. Axioms cannot be proven, but their statements have to be in accordancewith practical experience, without exception. Such a deductive approach is onlypossible for idealized models, which themselves represent a reduced picture of thereal world [16]; for many technical applications, it is not possible.

1.2 Modeling

With respect to technical mechanics, the aspect of modeling becomes one of themost important issues of mapping real world problems. Technical mechanics is an

© Springer-Verlag Berlin Heidelberg 2015 1F. Pfeiffer and T. Schindler, Introduction to Dynamics,DOI: 10.1007/978-3-662-46721-3_1

Page 11: Introduction To Dynamics

2 1 Basics

engineering science, which considers motion or deformations of technical systems.They generate loads on machines, mechanisms and structures, which must be knownfor their design. Mechanical modeling includes the replacement of a real machine,of real machine components or of real structures by certain basic elements. Consid-ering mechanics, this concerns for example masses, springs, dampers and frictionalelements, which according to the topology of a structure must be interconnected ina physically correct way, usually leading to certain types of constraints. This pro-cess requires a deep insight into the operational problems of a machine and a soundknowledge of practice on the one and of mechanical theories on the other side. Thequality of modeling decides on time and costs during a product development pro-cess, at least to a certain amount. Good models not only lead to quicker solutions,but also to better transparency of the problem under consideration and with it toaccelerated achievements for a technical problem.

What is a good model, or better, what is a good mechanical model? A mechanicalmodel will be good, if the mathematical model based on it gives us information closeto the reality, which is of special interest to us. We have to anticipate, that reality isknown, that it might be measurable or at least that it might be precisely describable.Therefore a good model should help us to come to a deeper understanding of thetechnical problems involved and of the design ideas behind them. To produce onlynumbers and charts will be not enough, we would like to produce insight. Creatingmodels has to keep that in mind.

How can we achieve a mechanical model? Usually we can assume, that everymachine, machine component or structure offers some important operational func-tions, which are easy to describe and to model. With regard to mechanical systems,these are for example some idealized motion sequences or vibrations, some effectsfrom kinematics and kinetics. We start with that. Looking a bit deeper into a struc-ture, we might realize, that machines cannot be built in an ideal way, that we areconfronted with disturbances, with "dirty effects", which in many cases cannot bemodeled straightforwardly and the mechanics of which is often not understood. Ex-actly at that point, the typical work of an engineer starts, which possesses more anintuitive-empirical character than a scientific one, for example the question, whatcan be neglected. A good mechanical model is always a minimalist model, notsmaller than necessary, but also not larger than adequate to the problem involved.Finding intuitively neglections, we may consider the geometric and kinematic situ-ation, the order of magnitude of forces and torques or of physical work and energy.Establishing a good model always needs an iteration process, which leads us withevery step to a better solution. In his famous lecture on "Clouds and Clocks" from1965 [54], Karl Popper told, that iterations are not only characteristic features ofevery intellectual work, but that they lead also from step to step to a deeper insightto the problem and to new questions finally achieving a really innovative solution,which at the beginning of such a process could not be perceived:

• mechanical modeling (theoretically and/or experimentally),• examination with respect to plausibility, comparisons with reality,• adaptation and improvements of models.

Page 12: Introduction To Dynamics

1.3 Basic Concepts 3

Establishing a mechanical model, we have to watch some important aspects con-cerning mathematical and numerical modeling, that means the whole sequence ofsteps to a final solution of our mechanical model ideas:

• discretization: Can we compose our model only by rigid body elements or evenby point masses or do we have to use elements with a continuum-mechanicalcharacter? How shall we model such nonrigid bodies?

• The character of expected motion: Does some basic motion exist or do we have atype of reference motion? Is there some state of rest? Is it possible to describe themotion as one with a (usually nonlinear) reference motion and small deviationsfrom it? Can we linearize, completely or in parts?

• coordinates: How many degrees of freedom exist for our model? Can we finda set of coordinates, which corresponds directly to these degrees of freedom? Ifnot, what sets of coordinates offer a formulation of constraints in a most simpleway?

• numerics: What solution methods fit best to our problem, analytically (if any) ornumerically? Can we put the mathematical formulation in a form, which corre-sponds in an optimal way to our solution possibilities? Is it possible to discoverwithin our mathematical model and formulation already some qualitative or evenquantitative results?

A perfect mechanical model, even a complicated one, will always be the simplestone possible, according to the well-known statement, that technology will be perfectif you cannot leave out anything. Especially for very complex systems we recom-mend to always start with a drastically simplified model for a better overview of theproblems involved. Then, in a second step, establishing a large model will be eas-ier. A good comprehension of the problems will always result in a better and fasterdevelopment process.

1.3 Basic Concepts

1.3.1 Mass

We consider dynamics in the sense as discussed in Section 1.1. That means we donot refer to relativistic aspects whatsoever. The only deviation from the classicalmass concept consists in the effects generated by rocket systems with their time-dependent masses. Focusing our future considerations mainly to technical artifacts,we usually know all relevant mass distributions and define:

• Masses are always positive, also in the time-dependent case, m > 0.• Masses are

– either constant with m = 0,– or not constant with m �= 0, where (m = dm

dt ).

• Masses can be added and divided into parts.

Page 13: Introduction To Dynamics

4 1 Basics

Another more physically oriented definition of mass is given by Synge [66]. Hestates, that a mass is “a quantity of matter in a body, a measure of the reluctanceof a body to change its velocity and a measure of the capacity of a body to attractanother gravitationally”.

Modeling masses depends on the problem under consideration. We might haverigid or elastic masses and in dynamics also interactions with fluid masses. Theoret-ically, we always get an interdependence of the selected mass model and the resultswe can achieve with such a model. However for many practical cases, the experi-ence of modeling tells us how to choose mass models. Nevertheless it makes senseto keep in mind these connections. In the following, we mainly consider systemswith constant masses.

1.3.2 Euler’s Cut Principle and Forces

In mechanics, we are interested in the interaction of bodies with forces or torques.If we therefore separate two bodies by isolating them we must at the same timearrange forces along the cutting line, which in the original configuration keeps thetwo bodies together. Thus by establishing free body diagrams, we transform internalforces to external ones acting on both sides of the cutting line with the same mag-nitude but opposite sign. This ingenious cut principle, first established by EULER,was characterized by Szabo [67] in a very appropriate way: “EULER teaches uswith the imagination of an artist to look in thought into the matter, where no eyeand no experiment can enter. With this he has laid a foundation for the only genuinemechanics, namely continuum mechanics.” The cut principle gives us the opportu-nity to establish the equations of motion for any part of a system, if we choose thecutting lines correctly and add to the applied forces and torques also the reactionforces and torques as isolated by these cuts. We need in addition a sign convention,which we may choose arbitrarily, but we must stay with it.

The cut principle allows us, to separate masses and mass systems from their envi-ronment. To illustrate the difference of internal and external forces depending on thecutting line selection, we use a simple example (Fig. 1.1). Considering the cuttingline 1 around the three masses, we see that all forces within that line are internalforces possessing no influence on the system 1. Selecting line 2, we come out withtwo external forces F12 and F32 and with two internal forces F13 and F31. Finally, theline selection 3 generates only external forces, namely F21 and F31.

The mechanical sciences are interested in the interaction of masses with forces.Dynamics as a part of mechanics is especially interested in those forces, which gen-erate motion. Therefore, we define the concept of active and passive forces. Activeforces can be moved in their direction of action, and from there they produce workand power. Passive forces cannot be moved with respect to their point of action. Ac-tive forces generate motion, passive forces prevent motion, they are the consequenceof constraints. All other definitions of forces are subsets of this concept. Internal orexternal forces, applied or constraint forces, volume or surface forces, they all maybe active or passive, depending on the specific system under consideration.

Page 14: Introduction To Dynamics

1.3 Basic Concepts 5

1

3

2

m2

m3

m1

F12

F32

F13

F31

F21

F23

Fig. 1.1. EULER’s cut principle and internal/external forces [16].

1.3.3 Constraints

If force interactions represent the very heart of mechanics, then constraints guaran-tee that the blood circulates through the vascular system. Constraints always possessa kinematic character. They are the mechanical controllers that tell systems whereto go and where not to go. In mechanical engineering, there is no machine or mech-anism that is not constrained. Constraints realize, at least kinematically, operationalrequirements and, applied correctly, guarantee the function of a mechanical system.Constraints can be bilateral or unilateral. In the first case, they represent ideal con-nections between two adjacent bodies, or between one body of the system and itsenvironment, and reduce the degree of freedom of the system. In the second case, aconnection may be open or closed, it may stick or slide, depending on the dynamicsof the system under consideration. Some typical examples are depicted in Fig. 1.2(pendulum), Fig. 1.3 (sledge), and in Fig. 1.4 (wheel).

Nearly all mechanical systems of practical relevancy are governed by a certainnumber of constraints, which depend on position, orientation, velocity, and time.Other forms of constraint do not exist, because it is clear from practical consid-erations that constraints describe only kinematic connection types. Constraints onan acceleration level are mathematically differentiated constraints and not in theiroriginal "physical" form. They are important in the theory of multibody systems.

On the basis of position coordinates z ∈ IRδ , velocity coordinates z ∈ IRδ ,time t, and constraint functions Φ ∈ IRm with m < δ , we give some structure tothese constraints [28, 68]. If the constraints depend explicitly on time, we callthem rheonomic. A constraint that does not depend on time is named scleronomic.

Page 15: Introduction To Dynamics

6 1 Basics

Constraints that depend only on position or orientation but not on velocity areholonomic constraints:

• holonomic-scleronomic constraints

Φ (z) = 0, (1.1)

• holonomic-rheonomic constraints

Φ (z, t) = 0. (1.2)

Every constraint might be differentiated with respect to time. Mathematically, theseequations are called hidden constraints. They possess no physical meaning but havethe character of invariants with the following linear structure:

0 = Φ (z, z, t) = W(z, t)T z+ w(z, t) , (1.3)

0 = Φ (z, z, z, t) = W(z, t)T z+ w(z, z, t) . (1.4)

Using such hidden constraints we know that they can be integrated to position level.Then, the following conditions according to SCHWARZ [58] have to be satisfied:

∂WTik

∂z j=

∂ 2Φ i

∂zk∂z j=

∂ 2Φ i

∂z j∂zk=

∂WTi j

∂zkand

∂WTik

∂ t=

∂ 2Φ i

∂zk∂ t=

∂ 2Φ i

∂ t∂zk=

∂ wi

∂zk.

(1.5)

Conversely, for the case z∈ IRδ , these conditions can also be used to decide whetheran integration to position level is possible [58]. However, in applications, we oftenfind constraints that are exclusively linear in some velocities with the property, thatthey cannot be integrated to position level. These are nonholonomic constraints.They can be expressed in the form:

• nonholonomic-scleronomic constraints

Φ (z, z) = W(z)T z+ w(z) = 0, (1.6)

• nonholonomic-rheonomic constraints

Φ (z, z, t) = W(z, t)T z+ w(z, t) = 0. (1.7)

Constraints of real physical significance, which include nonlinear relations in thevelocities, are not known.

Example 1.1 (holonomic-scleronomic constraints of the pendulum). The particlemass of the string pendulum in Fig. 1.2 moves on a circular trajectory with theradius R. This trajectory is described by its Cartesian position z := r with respect toa reference O. It moves in a gravity field with gravitational acceleration g. Cuttingthe string (Section 1.1), for example, leads to two forces acting on the pendulummass; the gravity force mg and a constraint force Fz, which forces the mass into

Page 16: Introduction To Dynamics

1.3 Basic Concepts 7

Ox

Rm

mg

y

r

Fz

Fig. 1.2. Holonomic constraints of the string pendulum.

the circular trajectory and, being a passive force, does not contribute to motion. Therelevant equations are:

mr = mg+Fz,

Φ(r) = rT r−R2 = 0.

In the following sections we present some more complete explanations.

Example 1.2 (holonomic-scleronomic constraints for the sledge). To a large ex-tent the motion of a sledge is governed by ground contours, which are thus respon-sible for the resulting constraint Φ(r) = 0 (Fig. 1.3). The position of the sledge isdescribed in Cartesian coordinates with z := r. The interaction with the ground istwofold. For the tangential direction, we have to consider frictional effects, and forthe normal direction, we must consider the normal constraint force as a passive forceinteracting with the ground and keeping the sledge on the ground contour. Assum-ing sliding of the sledge governed by COULOMB’s law of friction yields the slidingfriction force WT (r)μλN . The magnitude WT (r) represents the relevant normalizedtangential direction, μ ≥ 0 is the coefficient of sliding friction, and λN is a force pa-rameter resulting from the constraint in the relevant normalized normal directionWN(r). We get

mr = mg+WT (r)μλN +WN(r)λN ,

Φ(r) = 0.

We see that because of friction the parameter λN is involved in the free motion of thesledge in an active way. Therefore, (μλN) is not a constraint force in the classicalsense. It participates in the energy budget of the system and is an active force. For

Page 17: Introduction To Dynamics

8 1 Basics

Φ(r) = 0

x

mgr

WT μλN

WNλNy

Fig. 1.3. Holonomic constraints of the sledge.

μ = 0, we are left only with the constraint force λN forcing the sledge onto itstrajectory (Section 1.7).

Example 1.3 (Nonholonomic-scleronomic constraints of a wheel). We considera disc with radius R rolling on a plain (Fig. 1.4). In this case, we describe the discusing generalized coordinates z := (x,y,α,β )T and not Cartesian ones. Consideringthe rolling condition we get

ds = R dβ ,

and from this

dx = ds cosα = R cosα dβ ,dy = ds sinα = R sinα dβ .

αx

β

y

y

dssinα

dscosα

ds

z

R

Fig. 1.4. Nonholonomic constraints of a rolling disc.

Page 18: Introduction To Dynamics

1.3 Basic Concepts 9

Division by dt results in the following nonholonomic-scleronomic constraints:

Φ (z, z) =(

1 0 0 −R cosα0 1 0 −R sinα

)︸ ︷︷ ︸

WT

⎛⎜⎜⎝

xyαβ

⎞⎟⎟⎠

︸ ︷︷ ︸z

= 0.

For example,

Rsinα =∂WT

14

∂α�= ∂WT

13

∂β= 0.

The integration condition is not satisfied and the constraint is not integrable to po-sition level. If α ≡ α0 is constant, we can integrate the constraints resulting in a setof holonomic-scleronomic equations

(x− x0)−R cosα0 (β −β0) = 0,

(y− y0)−R sinα0 (β −β0) = 0

with integration constants x0, y0, and β0. The restriction α ≡α0 can be interpreted asrolling on a straight trajectory, where rolling forward and backward always resultsin the same starting position.

Constraints form constraint surfaces and generate constraint forces, which realize acertain motion structure as drawn up by the designer of the system. Examining theexamples and with some generalization, we conclude that these constraint forces arealways perpendicular to the constraint surfaces. This property will be the basis forthe principle of d’ALEMBERT, which is dealt with later (Section 1.7). Kinetically,motion can only take place on these constraint surfaces, and never perpendicular tothem.

1.3.4 Virtual Displacements

The concept of virtual displacement can to a certain extent be compared to the con-cept of variation in the calculus of variations [14]. The basic idea can be visualizedby considering a snap-shot of the generalized configuration z of a dynamic systemat a fixed time t. Then, a virtual displacement of a system is an arbitrary, imaginary,and small change δz compatible with the existing constraints [10, 59]. Using thisthought process, we have to bear in mind that virtual velocity changes need not nec-essarily be small, but compatibility with constraints must always be satisfied. Wecome back to this point later. Conversely, a real displacement of the system underconsideration takes places within the time interval dt, where all states, forces, andconstraints may be changing dependent on time.

Page 19: Introduction To Dynamics

10 1 Basics

Example 1.4 (Spherical pendulum). The point mass of a mathematical pendulummoves on the surface of a sphere with radius R (Fig. 1.5).

m

ϑ

ψ

r

y

z

x

δr

Fig. 1.5. Virtual displacement of a spherical pendulum, (‖ r ‖= R).

We use the Cartesian position z := r = (x,y,z)T for a description of the pendulum

mass by watching also the constraint φ(r) = rT r−R2 = x2 + y2 + z2 −R2 != 0. This

means that a small virtual displacement of the mass necessarily takes place "withinthis spherical surface." Up to terms of first order, we have

0!= φ(r+ δr) .

= φ(r)︸︷︷︸=0

+∂φ∂r

δr︸ ︷︷ ︸=:δφ

= 2rTδr = 2(xδx+ yδy+ zδ z).

This is a fundamental statement telling us with the term rTδr = 0 that a virtualdisplacement δr compatible with the constraints always has to be perpendicular tothe normal space of the constraint surface, that is δr ⊥ r. In the example of thespherical pendulum, this "constraint surface" is the sphere with radius R.

1.4 Kinematics

1.4.1 Coordinate Systems and Coordinates

We describe the motion of bodies by coordinate systems (Fig. 1.6). A Cartesiancoordinate system is a set of orthogonal unit vectors ex, ey, ez, which form a basis for

Page 20: Introduction To Dynamics

1.4 Kinematics 11

O

Iez

Ibody K1

body K2

O′1

K1ey

K1ez

Iex

Iey

K2ex

K2ey

K2ez O′

2K1

ex

Fig. 1.6. Translation and rotation of coordinates, I for "inertial".

all vectors in an inertial space. This basis of the vector space includes some origin O,which we may think of as being connected to some body (for example, some pointon the earth’s surface). The definition of an origin allows us to measure geometricmagnitudes and from this also dynamic processes. Depending on the state of motionof a coordinate system (O,ex,ey,ez), we call it inertial or noninertial (body-fixed).

The coordinate system (O,ex,ey,ez) possesses the property "inertial", if the basisvectors (ex,ey,ez) do not change with time, which means, that such a coordinatesystem might only move with constant velocity with respect to the postulated non-moving space. This is a question of definition. For technical dynamics, it is usuallysufficient to connect the earth or some building with an inertial coordinate system,for problems of space dynamics the sun might be a more suitable system. Often thecoordinate system can be assumed to be even space-fixed. We use the index I.

If we connect a coordinate system with a body to define a “body-fixed” coordi-nate system, we can choose to select any point for the origin. The center of massor another convenient point may be suitable for our evaluations, for example a jointin the case of robots. We use the index K. We may describe several coordinate sys-tems for one body. A basic property of a rigid body is the constant distance betweentwo material points. Rigid bodies have six degrees of freedom, three of which aretranslational and three rotational. Therefore, the three positions (x,y,z) and the threeorientations (α,β ,γ), given for example as Cardan angles, define the position andorientation of a rigid body with respect to any coordinate system, inertial or body-fixed. The position and the orientation of a body K1 can be described with respectto the inertial system I or with respect to the body-fixed system of body K2 by thesix magnitudes (x,y,z) and (α,β ,γ), where (x,y,z) are the coordinates of the centerof mass of the body K1, written in the bases I or K2, and where the Cardan angles(α,β ,γ) give the orientation between the coordinate system of K1 and those of I orK2, written correspondingly in the I- or K2-bases (see Fig. 1.6).

Returning to Fig. 1.6, we see that the motion of a body may be described bya relative translation of the origin O′ and by a relative rotation of the body-fixedcoordinate system K with respect to the reference coordinate system R = I (inertial

Page 21: Introduction To Dynamics

12 1 Basics

system) with origin O. The coordinates of O′ defined in R are Rx, Ry, Rz. The relativerotation of the body-fixed unit vectors Kex, Key, Kez are formally described by thecolumn vectors of the rotation matrix ARK ∈ IR3,3. Rotation matrices are orthogonaland possess the properties (the indices ’KR’ are defined after (1.23)):

AKR = A−1RK = AT

RK , (1.8)

detARK = 1. (1.9)

The rotation is described by nine "angular magnitudes," which depend on each other.Equation (1.8) defines six constraints according to Section 1.3.3, and (1.9) definesan additional sign condition. Altogether we are left with three translational degreesof freedom and also with three (3 = 9− 6) rotational degrees of freedom. Examplesof such minimal angle parametrizations are the EULER angles (Fig. 1.7)

ARK = ATz (ψ)AT

x (ϑ )ATz (ϕ)

=

⎛⎝cosϕ cosψ−cosϑ sinψ sinϕ −sinϕ cosψ−cosϑ sinψ cosϕ sinϑ sinψ

cosϕ sinψ+cosϑ cosψ sinϕ −sinϕ sinψ+cosϑ cosψ cosϕ −sinϑ cosψsinϑ sinϕ sinϑ cosϕ cosϑ

⎞⎠ ,

(1.10)

with the elementary rotations

nodal line

Ry

ϕψ

ϑ

ϑ

Rx

Rz

Kz

Kx

Ky

Fig. 1.7. EULER angles with azimuth- or precession angle ψ , elevation- or nutation angle ϑ ,and rotation angle ϕ .

Page 22: Introduction To Dynamics

1.4 Kinematics 13

Az(ψ) =

⎛⎝ cosψ sinψ 0−sinψ cosψ 0

0 0 1

⎞⎠ , (1.11)

Ax(θ ) =

⎛⎝1 0 0

0 cosθ sinθ0 −sinθ cosθ

⎞⎠ , (1.12)

Az(ϕ) =

⎛⎝ cosϕ sinϕ 0−sinϕ cosϕ 0

0 0 1

⎞⎠ (1.13)

and the Cardan angles (Fig. 1.8)

γ

Kx

Ry

Ky

α

α

β

β

γ

RzKz

Rx

α

β

γ

Fig. 1.8. Cardan angles.

ARK = ATx (α)A

Ty (β )A

Tz (γ)

=

⎛⎝ cosβ cosγ −cosβ sinγ sinβ

cosα sin γ+ sinα sinβ cosγ cosα cosγ− sinα sinβ sinγ −sinα cosβsinα sinγ− cosα sinβ cosγ sinα cosγ+ cosα sinβ sinγ cosα cosβ

⎞⎠

(1.14)

with the elementary rotations

Page 23: Introduction To Dynamics

14 1 Basics

Ax(α) =

⎛⎝1 0 0

0 cosα sinα0 −sinα cosα

⎞⎠ , (1.15)

Ay(β ) =

⎛⎝cosβ 0 −sinβ

0 1 0sinβ 0 cosβ

⎞⎠ , (1.16)

Az(γ) =

⎛⎝ cosγ sinγ 0−sinγ cosγ 0

0 0 1

⎞⎠ . (1.17)

We may also apply (1.8) and (1.9) as they use a kind of natural parametrization [30].We collect the descriptive parameters with respect to some coordinate system in avector of generalized coordinates. Using for example Cardan angles, we come outwith

z := (Rx ,Ry ,Rz ,α,β ,γ)T ∈ IR6. (1.18)

Considering a rigid body, we know, that any distance between two points of the rigidbody is constant enabling a description of any relative motion on it by one body-fixed coordinate system together with appropriate relative coordinates. For morecomplex systems of masses, for example multibody systems, we must also introducemore complex systems of descriptive coordinates. The position and orientation ofmany coordinate systems K1, . . . ,Kn are then defined in a configuration space ofdimension (δn):

z := (zT1 , . . . ,z

Tn )

T ∈ IRδn. (1.19)

If we are not able to find a set of minimal coordinates, we have to additionallyconsider a corresponding set of constraints of the form (Section 1.3.3)

φ i(z) = 0 for i ∈ {1, . . . ,m} and m < δn. (1.20)

In many cases of practical relevancy, we may find a complete set q of minimal co-ordinates with the number f := δn−m of degrees of freedom. However, this willbe not possible in every case, and we should also bear in mind, that there are noprecise rules for the detection of minimal coordinates. The corresponding mathe-matical formulae do not necessarily replace a physical search. Expressing z by theminimal coordinates q we obtain

φ i(z(q)) = 0 ∀q ∈ IR f (1.21)

for i ∈ {1, . . . ,m}. Summarizing all these coordinates we write

• r Cartesian position of some point,• z generalized coordinates,• q minimal coordinates.

Page 24: Introduction To Dynamics

1.4 Kinematics 15

The derivations r, z, q with respect to time are called velocity coordinates.

1.4.2 Transformation of Coordinates

The kinematic relations always form a basis for generating the equations of mo-tion. For large dynamic systems, we get various sets of coordinates, which requiretransformations from one to the other and from some inertial basis to some body-fixed coordinate systems. Typically, the fundamental kinetic laws apply with respectto an inertial, space-fixed basis, whereas bodies are better described in a body-fixedframe. As a consequence, we need transformations between all existing coordinates.Performing the corresponding processes, we should keep in mind, that a vectorsticks to a vector in spite of the fact that it will be described by different coordi-nates in different coordinate systems.

P

O

O′

Ix Iy

Iz

rO′

rO′P

rP

I

K

Kx

Kz

Ky

Fig. 1.9. Displacement and rotation of coordinates.

According to Fig. 1.9, we represent a relative displacement of the body-fixedpoint P by the two vectors from the inertial system to the body-fixed one and fromthe body-fixed one to point P. It is

KrP = KrO′ + KrO′P . (1.22)

In the following text, we omit the indices O or I, if they refer to some inertial coor-dinates.

According to Section 1.4.1, the rotation of coordinate systems is described bya rotation matrix, which also defines the transformation matrix. It transforms thevector KrP of the K-system into a vector IrP of the I-system [10, 40]:

IrP = AIK KrP . (1.23)

The double index IK means the following: matrix A transforms vector rP, given inthe K-system, into the I-system. All similar double indices should be interpreted inthe same way.

Page 25: Introduction To Dynamics

16 1 Basics

1.4.3 Relative Kinematics

From the discussion above it is clear, that we need all absolute and relative veloc-ities and accelerations of the bodies under consideration and of their coordinates.Therefore, the main goal in establishing relative kinematics (Fig. 1.10) consists inevaluating these magnitudes.

This is formal work provided we have established before a precise geometric de-scription of the whole system, including a convenient choice of coordinates and abest possible set of constraints based on a good choice of interconnections. NEW-TON’s "mass times acceleration is equal to force" is simple, but the accelerationscan be awfully complex. Let us start with (1.22), (1.23), and Fig. 1.10. From therewe write

O

rP

rO′

P

O′

ω

vO′

rO′P

K

rO′Ptrajectory of P on K

Iz

IxIy

Kz

Kx

Ky

Fig. 1.10. Relative kinematics of a body.

IrP = IrO′ +AIK KrO′P . (1.24)

Deriving this formally results in

I rP = I rO′ + AIK KrO′P +AIK K rO′P , (1.25)

I rP = I rO′ + AIK KrO′P + 2AIK K rO′P +AIK K rO′P . (1.26)

The vector IvP = I rP represents the absolute velocity of point P in the I-system andthe vector K rO′P its relative velocity as seen from O′ in the K-system. The trans-lational velocity of point O′ is an absolute velocity IvO′ = I rO′ represented withrespect to the I-system.

In a similar sense, the vector IaP = I rP is the absolute acceleration of point Pin the I-system and K rO′P its relative acceleration as seen from O′ in the K-system.Also IaO′ = I rO′ is the absolute acceleration of point O′ in the I-system. Going tothe body-fixed frame, we have to multiply these relations by AKI and get

Page 26: Introduction To Dynamics

1.4 Kinematics 17

KvP =KvO′ +ATIKAIK KrO′P + K rO′P , (1.27)

KaP =KaO′ +ATIKAIK KrO′P + 2AT

IKAIK K rO′P + K rO′P . (1.28)

Formal derivation of the identity matrix E = ATIKAIK results in

0 = E = ATIKAIK + A

TIKAIK (1.29)

demonstrating the skew-symmetric properties of

Kω := ATIKAIK =−(AT

IKAIK)T

=−KωT . (1.30)

It means that Kω can be represented by three independent components only. Com-bining this in a vector Kω yields

Kω KrO′P = Kω× KrO′P and Kω =

⎛⎝ 0 −Kω3 Kω2

Kω3 0 −Kω1

−Kω2 Kω1 0

⎞⎠ . (1.31)

Then, we get from (1.27)

KvP = KvO′ + Kω× KrO′P + K rO′P . (1.32)

We come back to Fig. 1.10 with the moving body K and understand this figure as amoving base with a man walking on it, for example some kind of merry-go-round.The man is represented by point P, moving along some trajectory. The absolutevelocity KvP of point P has two parts, first a relative velocity of P with respect to O′and second an applied velocity resulting from the body’s motion with respect to theinertial system, which writes

KvO′ + Kω× KrO′P . (1.33)

The rotational term Kω×KrO′P is generated by the angular velocity Kω of the bodyK written in a body-fixed frame with the distance KrO′P of the point P from O′. Bynot considering the translational velocity of point O′ in (1.32), we get as a result theabsolute rate of change of KrO′P with time:

KvP = K rO′P + Kω× KrO′P . (1.34)

This equation is one of the fundamental relations of relative kinematics. It appliesto every vector in and on a moving system like the body K, not only to the specialvector KrO′P. It is called Euler’s theorem or the Coriolis equation. In words:

The absolute rate of change of a vector with respect to time is equal to thesum of its applied rate of change and of its relative rate of change [41].

Page 27: Introduction To Dynamics

18 1 Basics

The determination of the angular velocities may be performed formally using therotation matrices and their derivations, but there is a nice and more elegant way toderive these angular velocities by considering the rotation sequences as depicted inFigs. 1.7 and 1.8.

Starting with the EULER angles, we notice from Fig. 1.7 the following sequenceof rotations: In a first step we rotate with the angle ψ about the z-axis and cometo an end-position in the form of a nodal line. From there we tilt with ϑ about thisnodal line, and then finally we rotate with ϕ about the final z-axis. To achieve arepresentation in inertial coordinates, we must transform back the second and thethird rotations into a global frame. We get

Iω = ezψ+ATz (ψ)exϑ +AT

z (ψ)ATx (ϑ)ezϕ

=

⎛⎝ sinψ sinϑ cosψ 0−cosψ sinϑ sinψ 0

cosϑ 0 1

⎞⎠⎛⎝ϕ

ϑψ

⎞⎠ .

In a similar way, we do that for the Cardan angles. We rotate first with α about thex-axis, second with β about the just-generated y-axis, and third with γ about the newz-axis. Then, we come out with

Iω = exα+ATx (α)eyβ +AT

x (α)ATy (β )ezγ

=

⎛⎝1 0 sinβ

0 cosα −sinα cosβ0 sinα cosα cosβ

⎞⎠⎛⎝α

βγ

⎞⎠ .

Some relationship for angular velocities dependent on a set of rotation parametersϕ , for example ϕ = (ϕ ,ϑ ,ψ)T or ϕ = (α,β ,γ)T , can always be established in theform

Iω = Y−1I

˙ϕ ,

but the inverse of this matrix Y−1I does not exist for all possible magnitudes of the

relevant angles. For our two sets above, we have

YI =1

sinϑ

⎛⎝ sinψ −cosψ 0

cosψ sinϑ sinψ sinϑ 0−sinψ cosϑ cosψ cosϑ sinϑ

⎞⎠ (EULER angles),

YI =1

cosβ

⎛⎝cosβ sinβ sinα −sinβ cosα

0 cosβ cosα cosβ sinα0 −sinα cosα

⎞⎠ (Cardan angles).

These expressions possess singularities, which are typical for angular systems con-taining a minimal set of parameters. The singularity with respect to the Cardanangles is also called the "gimbal lock". The singularities for the EULER and Car-dan angles are at ϑ = 0 and β = π

2 , respectively, which are in both cases special

Page 28: Introduction To Dynamics

1.4 Kinematics 19

angles of the second elementary rotation. This can be generalized. If we design thesecond elementary rotation in such a way, that the first and third elementary rota-tion take place about the same spatial axis, then the complete rotation will come outwith redundancy including a singularity. The singularity can be avoided by choos-ing another set of minimal parameters. It can be completely avoided by applyingnot a minimal set but a nonminimal set of rotation parameters, for example EULER

parameters or quaternions [30, 5, 63].Proceeding from relative velocities to relative accelerations, we consider (1.28)

and the point P moving on K in Fig. 1.10. By formal derivation, we get

K ˙ω =ddt

(AT

IKAIK)= A

TIK AIKAT

IK︸ ︷︷ ︸E

AIK +ATIKAIK =−Kω Kω+AT

IKAIK (1.35)

and in combination with (1.28) the result

KaP = KaO′ + Kω× KrO′P + Kω× (Kω× KrO′P)+ 2 Kω× K rO′P + K rO′P . (1.36)

We may derive this equation also by applying EULER’s theorem (1.32) concerningthe absolute velocity

KaP = Kω× KvP + K vP

= Kω× (KvO′ + Kω× KrO′P + K rO′P)+ddt(KvO′ + Kω× KrO′P + K rO′P)

(1.37)

together with some organization of the terms.The absolute acceleration possesses three components:

• applied acceleration

KaO′ + Kω× KrO′P + Kω× (Kω× KrO′P) (1.38)

with the absolute acceleration of the moving body (reference) KaO′ , the rotationalacceleration Kω× KrO′P, and the centripedal acceleration Kω× (Kω× KrO′P),

• Coriolis acceleration

2 Kω× K rO′P , (1.39)

• relative acceleration

K rO′P . (1.40)

Example 1.5 (Rotating disc with a radially guided ball). A disc K turns withconstant and positive angular velocity (Kω =

(0 0 ω

)T,and ω > 0) about a fixed

center O′ = O. A ball P moves with constant and positive velocity (v > 0) in atube fixed on the disc, which means KrO′P =

(vt 0 0

)T. We look for the absolute

Page 29: Introduction To Dynamics

20 1 Basics

Iy

Ix

KyKx

K rO′P

ωt

Kω×KrO′P

O′KrO′P

P

Fig. 1.11. Rotating disc with a radially guided ball.

velocity and the absolute acceleration of the ball, starting with the time t = 0. Ac-cording to the relations above, we get for the absolute velocity

KvP = Kω× KrO′P + K rO′P =(v ωvt 0

)T

where the two terms are the relative velocity v of the ball in radial direction and itsapplied velocity ωvt in circumferential direction. The absolute acceleration writes

KaP = Kω× (Kω× KrO′P)+ 2 Kω× K rO′P =(−ω2vt 2ωv 0

)T.

We recognize the centripedal acceleration −ω2vt in the negative radial directionand the Coriolis acceleration 2ωv in circumferential direction, well known fromthe devil wheel of modern fairs. One part of the Coriolis acceleration comes fromthe derivation K vP =

(0 ωv 0

)T, a second part comes from the angular motion

Kω×KvP =(−ω2vt ωv 0

)T, which gives in addition the centripedal acceleration

in negative radial direction.We perform an experiment of thought. Sitting for example on this ball not know-

ing the rotation of the disc, then the ball is exposed only to its relative accelerationwith respect to the disc. Considering now NEWTON’s law, we must express the ab-solute relative acceleration, "felt" by the ball, by all its components as discussedbefore, namely

KFa = m KaP = m Kω× (Kω× KrO′P)+m2 Kω× K rO′P +m K rO′P .

Therefore, the ball "feels" the centrifugal force in radial direction and the Coriolisforce in negative circumferential direction:

Page 30: Introduction To Dynamics

1.4 Kinematics 21

m K rO′P = KFa +[−m Kω× (Kω× KrO′P)]︸ ︷︷ ︸centrifugal force

+[−m2 Kω× K rO′P]︸ ︷︷ ︸Coriolis force

.

These forces have to be added to the momentum equation, if this relation is to beevaluated in a moving system. The Coriolis force will produce a load on the righttube side (negative circumferential direction).

Example 1.6 (Disc on a car). We consider a rocking disc on a car (Fig. 1.12). Therocking function ϕ(t) and the position xW (t) of the car are given. We want to eval-uate the absolute velocity IvA and the absolute acceleration IaA of the contact pointA of the disc. Performing this task, we proceed with the following steps:

• The center point M of the disc can be treated quite easily to give:

IrM =(xW +L+Rϕ R 0

)T.

From this, we get its absolute velocity as

IvM = I rM =(xW +Rϕ 0 0

)T.

• With respect to the point A, we apply (1.32). The angular velocity of the disc is

Iω =(0 0 −ϕ

)T. The relative velocity of A will be zero in a body-fixed frame

and as a consequence also in a space-fixed frame after a transformation. Theradius vector is IrMA =

(0 −R 0

)T, and finally we get

IvA =(xW +Rϕ 0 0

)T+(0 0 −ϕ

)T × (0 −R 0)T

=(xW 0 0

)T.

The result corresponds exactly to the roll condition.

Ix

MRϕ

L Rϕ A

xW

O

xM

Iy

xW

Fig. 1.12. Disc on a car (with the linearization R tan(ϕ)≈ Rϕ).

Page 31: Introduction To Dynamics

22 1 Basics

• The absolute acceleration of the center point M writes

IaM = I rM =(xW +Rϕ

).

• The relative kinematics of point A follows from the relation (1.36). The relativeacceleration of A will also be zero, A being body-fixed. We get

IaA =(xW +Rϕ 0 0

)T+(0 0 −ϕ

)T × (0 −R 0)T

+(0 0 −ϕ

)T ×((

0 0 −ϕ)T × (0 −R 0

)T)

=(xW +Rϕ 0 0

)T − (Rϕ 0 0)T

+(0 0 −ϕ

)T × (−Rϕ 0 0)T

=(xW Rϕ2 0

)T.

1.5 Momentum and Moment of Momentum

1.5.1 General Axioms

According to Section 1.1, mechanics is interested in those interactions that generateaccelerations of bodies or systems of bodies. Therefore, we formulate the very basicaxiom that generally external forces on bodies are in equilibrium with the mass orinertia forces [16],

∫K(rdm− dFa) = 0, (1.41)

and that all internal forces vanish,∫

KdFi = 0. (1.42)

Considering absolute changes with time, this statement is true for an inertial systemof coordinates (Fig. 1.13), where a radius vector r points to a mass element dmloaded by external forces dFa. Calculating the torque with respect to point O andkeeping in mind the equilibrium statement above, this torque has to be balanced bythe moment of momentum in the form:

∫K

r× (rdm− dFa) = 0. (1.43)

This relation obviously can only be true, if at the same time the torque of all internalforces vanishes, which writes

∫K

r× dFi = 0. (1.44)

This requirement turns out to be much more complex than the simple conditionthat internal forces have to vanish. It anticipates the symmetry of Cauchy’s stress

Page 32: Introduction To Dynamics

1.5 Momentum and Moment of Momentum 23

O

x

z

y

r

dm

dFa

K

Fig. 1.13. Forces and torques on a mass element.

tensor, which is equivalent to the well-known axiom of Boltzmann. Consideringthe laws of continuum mechanics, the symmetry of Cauchy’s stress tensor followsfrom a moment of momentum balance for a volume element of the continuum [6].The four postulates as presented above may be considered as axioms resulting fromcentury-old experience. They are sufficient to derive all relations of theoretical andapplied mechanics.

1.5.2 Momentum

The equations (1.41), (1.42), and (1.44) also include the three NEWTONian axioms.

AXIOM 1: A body at rest remains at rest and a body in motion moves in astraight line with unchanging velocity, unless some external force acts on it.

To illustrate this basic law, which we find already in the statements ofGALILEI [67], we use the notation introduced by EULER for the momentum andmoment of momentum laws. Referring to this axiom we have no external, thus noactive forces, which means

∫K dFa = 0 and therefore

∫K rdm = 0 resulting in

p =

∫K

rdm = const., (1.45)

which represents the law of conservation of momentum. Considering the mass cen-ter of a body, we get

Page 33: Introduction To Dynamics

24 1 Basics

p = rSm with rSm =∫

Krdm. (1.46)

AXIOM 2:The rate of change of the momentum of a body is proportional tothe resultant external force that acts on the body.

For the mass element of Fig. 1.13, we obtain from (1.41)

rdm− dFa = 0, (1.47)

which represents also the momentum budget for a point mass. The time derivativeof the momentum is mass times acceleration if we are dealing with a constant mass,as in the above equation. For masses that are not constant the time derivative of themass must also be considered. In terms of the definitions, we write

dpdt

= F, with p =

∫K

rdm, and F =

∫K

dFa. (1.48)

If we consider again the center of mass of the body, we come out with

m(dvS

dt) = FS. (1.49)

The velocity vS is defined with respect to an inertial system. It is an absolute veloc-ity. The force vector FS is the vector sum of all forces that act on the body. Generally,this vector sum does not pass through the center of mass resulting in an additionaltorque, which has to be regarded in the moment of momentum equation.

AXIOM 3: Action and reaction are equal and opposite.

At the time of NEWTON, this finding was new. However, it is very obvious fromexperience. Wherever any force acts on a body or on the environment, we get asa reaction the same force with opposite sign. My feet transfer my weight to theground, as a reaction the ground is loaded with my weight force in the oppositedirection. There is no mechanical interaction without this basic property.

1.5.3 Moment of Momentum

Leonhard EULER was the first to recognize the moment of momentum equation asan original and independent basic law of dynamics [67]. In agreement with (1.43)and (1.44), we define as moment of momentum for a body K

LO :=∫

Kr× (rdm) (1.50)

Page 34: Introduction To Dynamics

1.6 Energy 25

and in addition the torque of the external forces

MO :=∫

Kr× dFa. (1.51)

From this, we formulate the following

moment of momentum equation:

dLO

dt= MO. (1.52)

We have to keep in mind that the moment of momentum as well as the torquedepend always on some reference point. As we know from the discussion above, thedefinition of the moment of momentum equation (1.52) presupposes Boltzmann’saxiom or, equivalently, the symmetry of Cauchy’s stress tensor [6]. Deriving themoment of momentum with respect to time requires EULER’s differentiation rule(Section 1.4.3) [41].

1.6 Energy

A mass element dm moving by dr from a point 1 to a point 2 in a force field (Fig.1.14) produces physical work. This work might be expressed either by the externalforces or by the equivalent acceleration (1.41). It is

dW =

∫ r2

r1

dFTa dr = dm

∫ r2

r1

rT dr. (1.53)

r1

r22

1

dr

dm

dFa

Fig. 1.14. Mass element in a field of forces.

Page 35: Introduction To Dynamics

26 1 Basics

The acceleration term can be manipulated in the following form

dm∫ r2

r1

rT dr = dm∫ r2

r1

drT

dtdr = dm

∫ r2

r1

rT dr (1.54)

resulting after an integration in a dependency on the kinetic energy

dm∫ r2

r1

rT dr =12

dm[r2

2 − r21

]=: dT2 − dT1. (1.55)

From this, we conclude, that the work done by the external forces dFa is equal tothe difference between the corresponding kinetic energies.

Energy equation

dW = dT2 − dT1. (1.56)

We now focus our considerations on fields of forces, where along a closed tra-jectory Γ no work is done. Examples are gravitational fields, the electric fields ofCOULOMB, all central force fields, and also the forces of springs. This means that

∮Γ

dFTa dr = 0. (1.57)

Systems with such features are called conservative systems. They are free fromenergy losses and also from energy sources. Applying STOKES’ theorem [77] to(1.57), we get

∮Γ

dFTa dr =

∫A∇× dFadA. (1.58)

The area dA is surrounded by the closed trajectory Γ , where ∇× dFa is the rota-tion of the force field through this area. With dA being arbitrary, we arrive at theirrotationality of the force field in the form

∇× dFa = 0, (1.59)

which is equivalent to (1.57). The rotation of a flow field corresponds to twice itsangular velocity.

Both relations (1.57) and (1.59) are necessary conditions for the existence of apotential; for practical applications they are also sufficient conditions [58]:

dFTa =: −∇dV. (1.60)

The magnitude dV is called the potential energy of the mass element under consid-eration, sometimes also its potential function. From the above equations, we obtain

Page 36: Introduction To Dynamics

1.7 Principles of d’Alembert and Jourdain 27

∫ r2

r1

dFTa dr =−

∫ r2

r1

(∂dV∂r

)T

dr =−∫ dV2

dV1

d(dV) =−(dV2 − dV1) . (1.61)

Therefore, the work dW is identical with the negative difference of the potential dV :

dW =−(dV2 − dV1) = dT2 − dT1. (1.62)

Considering these relations and rearranging a bit we see, that for conservative sys-tems the total energy at point 1 must be the same as for point 2:

dE1 := dT1 + dV1 = dT2 + dV2 =: dE2. (1.63)

For a total mass m and not only a mass element dm, we get the

energy equation for conservative systems:

T +V = const. . (1.64)

For conservative systems, the energy budget does not change. However, for ex-ample having friction in the system, the total energy would decrease and the lineintegral (1.57) would not vanish. Force fields depending on time and/or velocity arenot conservative. The discussion of energy results more or less from the discussionof momentum and moment of momentum, including Boltzmann’s axiom. From this,the energy statements are not an additional axiom. However they are, fortunately, afirst integral of motion, at least for conservative systems, and thus an invariant mag-nitude of motion. This can be very useful with respect to further considerations ofdynamics.

The relations for momentum and moment of momentum are sufficient for analyz-ing the motion of any mechanical system. However, in many cases their applicationis costly and sometimes difficult. For most problems of practical relevancy, we needto deal with constraints, which makes another approach necessary. We come backto this immediately. Anyway, in connection with these principles, the merit of IsaacNEWTON (1642-1727) consists in the formulation of the three basic axioms of mo-mentum, and the merit of EULER (1707-1783) in the introduction of the cut principleand a concise formulation of the moment of momentum axiom as an original andindependent law.

1.7 Principles of d’Alembert and Jourdain

1.7.1 Significance of Constraints

In this section we introduce the following concept of forces [50]. The pair "im-pressed" and "constraint (reaction)" forces describes in the first case given or ap-plied forces, very often interpretable as physical laws (gravitation, magnetic fields,

Page 37: Introduction To Dynamics

28 1 Basics

springs and dampers etc.), and in the second case forces produced by constraints.The pair "active" and "passive" forces describes forces that contribute or do not con-tribute to motion. Reaction forces are always passive forces due to their constrainingcharacter. Impressed forces may be active or passive [51]. As already discussed inSection 1.3, the definition of internal and external forces depends on the structure ofthe free body diagram, that is how we have designed our cuts.

The existence of constraints implies two difficulties. The first one concerns theindependence of coordinates, which are constrained. Therefore, the original coordi-nate set, for example in some three-dimensional workspace, does not represent thepossible number of degrees of freedom. Some of the equations of motion dependon each other. The second difficulty is connected with the forces due to constraints.These constraint forces are not given a priori, they must be evaluated by the solutionprocess. Moreover, as the constraint forces do not contribute to the motion of thesystem, they are reaction forces holding the system together, where we should bearin mind that passive forces and motion means passive forces and relative motion.From the technical viewpoint, we need them as forces in bearings, guides, joints,and the like; they determine system design. The inverse is also true: every systemdesign defines constraints.

The derivation of the equations of motion in respect of constraints should takeplace with the requirements that

• the constraints can be taken into account by some automatic procedure,• it is as simple and transparent as possible,• the resulting differential equations can easily be solved.

These requirements define something like the squaring of a circle, but they alsodefine a direction to go in. The most effective tools for considering dynamic sys-tems with constraints are the principles of mechanics, like those of d’ALEMBERT,JOURDAIN, GAUSS, and HAMILTON [10, 16, 28, 59, 68], to give a few and mostimportant examples. We start with the principles of d’ALEMBERT and JOURDAIN.

1.7.2 Principle of d’Alembert

Summing up (1.41) and (1.42) by considering a body K, partitioning additionally theforces into impressed and constraint forces (instead of external and internal forces),we get

∫K

rdm =

∫K

dF =

∫K(dFe + dFz) . (1.65)

Applying a virtual displacement δr to the mass element dm (Fig. 1.15), we comeout with

(rdm− dFe)T δr = (dFz)T δr = δW z. (1.66)

Page 38: Introduction To Dynamics

1.7 Principles of d’Alembert and Jourdain 29

Φ(r) = 0

xy

zK

dm

δ rr

Fig. 1.15. Virtual displacement of a mass element.

This is the virtual work as a result of the virtual displacement δr. It is generated ei-ther by the impressed forces dFe and the mass forces rdm or by the constraint forcesdFz. Without loss of generality, we assume holonomic-scleronomic constraints inthe following:

Φ (r) = 0 with r ∈ IR3, Φ ∈ IRm, m < 3. (1.67)

The constraint equations do not allow a free choice of the virtual displacements,which always have to be compatible with these constraints. We write (Section 1.3.4)

0 !=Φ(r+ δr) .

=Φ(r)︸ ︷︷ ︸=0

+∂Φ∂r

δr =: δΦ . (1.68)

This relation means geometrically, that m constraints (1.67) in the space IR3 spanin total m constraint surfaces {Φν(r) = 0}m

ν=1, the surface normals nν of which are

proportional to(

∂Φν∂r

)T, see Fig. 1.16:

(∂Φν∂r

)T

⊥ δr. (1.69)

Any virtual displacement can only take place on the surfaces of constraints, not per-pendicular to them. Conversely, the constraint forces as generated by the constraintsΦ (r) = 0 can only be perpendicular to these surfaces, which means in the normalsurface direction (Section 1.3.3), because the situation where any body leaves theconstraint surfaces must be avoided. It is a very hard condition, that free motion mustand only can take place within these surfaces. The term (dFz)T δr= 0 represents thesubstantial meaning of d’ALEMBERT’s principle, which writes using LAGRANGE’sversion [28, 68]:

Page 39: Introduction To Dynamics

30 1 Basics

r1

r3

r2

Φν (r) = 0δr

nν ∼(

∂Φν∂r

)T

Fig. 1.16. Constraint surfaces.

constraint forces do no work,∫

K(dFz)T δr = 0. (1.70)

With respect to generating motion they are "lost forces", which compel the dy-namic system into the surfaces of constraints. The motion itself will only be pro-duced by the impressed forces.

Principle of d’ALEMBERT:∫

K(rdm− dFe)T δr = 0. (1.71)

It should be noted that constraint forces may be shifted in the tangential planeof the constraint surfaces without violating conditions (1.69) and (1.70), becausethese conditions only say that the constraint forces dFz must be perpendicular to δr.This property is important for problems including friction, where for COULOMB’sfriction law the absolute value of sliding friction is proportional to the absolute valueof the constraint force in normal direction of a contact.

With respect to the static case (r = 0), D’ALEMBERT’s principle (1.71) reducesto the principle of virtual work [41]:

δW =∫

K(dF)T δr = 0. (1.72)

Page 40: Introduction To Dynamics

1.7 Principles of d’Alembert and Jourdain 31

According to D’ALEMBERT’s principle, we may split up the original relation∫

K(rdm− dFe) =

∫K

dFz (1.73)

into two terms, one being true for the directions perpendicular to the constraintsurfaces Φ (r) = 0 (⊥), and the other being true for all directions tangential to theconstraint surfaces (‖):

∫K(rdm− dFe)⊥+

∫K(rdm− dFe)‖ =

∫K(dFz)⊥+

∫K(dFz)‖ . (1.74)

As the motion definitely takes place only on the surfaces Φ (r) = 0, we require thatthe normal and tangential parts become zero independently:

∫K(rdm− dFe)‖ =

∫K(dFz)‖ = 0, (1.75)

∫K(rdm− dFe − dFz)⊥ = 0, (1.76)

which means:

• The impressed tangential forces alone contribute to any acceleration of masses.• The normal components of the impressed forces are in static equilibrium with

the constraint forces and the mass forces (rdm)⊥.

We close this section with some historical remarks concerning the evolution of theprinciple of D’ALEMBERT [67]:

1. Daniel BERNOULLI (1700 - 1782)Forces not contributing to accelerations are "lost forces":

rdm− dFe = dFz.

2. Jean-Baptiste le Rond d’ALEMBERT (1717 - 1783)The sum of all "lost forces" must be in an equilibrium state.

3. Joseph-Louis LAGRANGE (1736 - 1813)"Lost forces" do no work. From this we have the so-called LAGRANGE’s version:

(rdm− dFe)T δr = (dFz)T δr = 0.

1.7.3 Principle of Jourdain

The principle of JOURDAIN follows qualitatively from the same arguments that weused for the principle of D’ALEMBERT in the form of LAGRANGE. Constraint equa-tions define a certain type of motion, they give an instruction as to how motionshould develop. A revolute joint for example prescribes at its location an angularmotion with one degree of freedom only. The constraint forces resulting from sucha constraint are always perpendicular to the surface as defined by the corresponding

Page 41: Introduction To Dynamics

32 1 Basics

kinematic magnitudes of the constraint. They cannot be displaced in their positivedirection, and constraint torques cannot be rotated. Thus, the example of the rev-olute joint requires that the corresponding constraint force is perpendicular to thejoint axis.

From these properties we conclude that such constraint forces not only do nowork according to D’ALEMBERT (1.70), but that they also cannot produce power.This is the substantial statement of JOURDAIN’s principle. The scalar product of thevector dFz of constraint forces and the vector δ r of nearly any virtual velocity mustvanish.

Philip JOURDAIN (1879 - 1919):Constraint forces generate no power.

∫K(dFz)T δ r = 0. (1.77)

The principle of JOURDAIN is obviously very advantageous with respect to non-holonomic constraints, because we can use them directly. Holonomic constraints(1.7) must be differentiated formally to give

Φ (r, r, t) = W (r, t)T r+ w(r, t) = 0 with r ∈ IR3, Φ ∈ IRm, m < 3.(1.78)

D’ALEMBERT’s principle is limited to holonomic constraints and small virtual dis-placements. JOURDAIN’s principle is much more general including all possible con-straints, if we allow formal differentiation. We consider changes concerning thevirtual velocity, δ r, δr = 0, δ t = 0 [50], which results in

0 !=Φ(r, r+ δ r, t) =Φ(r, r, t)︸ ︷︷ ︸

=0

+∂Φ∂ r

δ r =: δΦ . (1.79)

This again confirms the same properties as before; the virtual velocities and theconstraint normal must be perpendicular to each other:

W ⊥ δ r with W =∂Φ∂ r

. (1.80)

We derived (1.78) by formal differentiation of some holonomic constraint resultingin a linear relationship with respect to velocity. This is of much more general sig-nificance, because additionally all physically meaningful nonholonomic constraintsalways possess this linear form. Up to now no real exception has been discerned.However, this means that the variation of the velocity ∂ r in JOURDAIN’s principle(1.77) need not be small, necessarily. In summation, we can write

Page 42: Introduction To Dynamics

1.8 Newton-Euler Equations for Constrained Systems 33

∫K(rdm− dFe)T δ r = 0. (1.81)

1.8 Newton-Euler Equations for Constrained Systems

The really brilliant principles of D’ALEMBERT and JOURDAIN offer an elegant andtransparent possibility to combine the equations of motion of a single body or of asystem of bodies with all constraints concerning such a mass system. We considerthe most convenient methods in the following.

1.8.1 Single Rigid Body

The relationship (1.81) includes the absolute velocity r and the absolute accelerationr of an arbitrary point of a rigid body K. For expressing the absolute magnitudes byrelative ones we define this arbitrary point as point P. We obtain from (1.32) and(1.36) together with the rigid body assumptions K rO′P = 0 and K rO′P = 0

r = vO′ + ωrO′P, (1.82)

r = aO′ +(

˙ω+ ωω)

rO′P. (1.83)

Provided we find a set of minimal coordinates q, q, we express r, r by the func-tions r(q, t) and r(q, q, t). These minimal coordinates satisfy all constraints (1.78)automatically. The time derivation r can be written in the form

r =(

∂r∂q

)q+

∂r∂ t

. (1.84)

From this, we come to the very important relation(

∂ r∂ q

)=

(∂r∂q

). (1.85)

According to the agreement above, we consider only δ q-variations and get

δ r =(

∂ r∂ q

)δ q =

(∂r∂q

)δ q. (1.86)

Regarding (1.82), the relation ωrO′P = rTO′Pω gives

δ r =[(

∂vO′

∂ q

)+ rT

O′P

(∂ω∂ q

)]δ q. (1.87)

Composing this relation together with (1.83) and (1.81) we derive the followingintegral

Page 43: Introduction To Dynamics

34 1 Basics

∫Kδ qT

[(∂vO′

∂ q

)+ rT

O′P

(∂ω∂ q

)]T {[aO′ +

(˙ω+ ωω

)rO′P

]dm− dFe}= 0.

(1.88)

Now, we choose the virtual velocities δ q in an arbitrary way, because the minimalcoordinates satisfy the constraints automatically. Applying the fundamental lemmaof variational calculus to the equation above, which requires the integrand to be

zero, and arranging the terms according to(

∂vO′∂ q

)Tand

(∂ω∂ q

)T, we finally get

0 =

(∂vO′

∂ q

)T [∫K

(aO′ +

(˙ω+ ωω

)rO′P

)dm−

∫K

dFe]

+

(∂ω∂ q

)T [∫K

rO′PaO′dm+

∫K

rO′P ˙ωrO′Pdm+

∫K

rO′PωωrO′Pdm−∫

KrO′PdFe

].

(1.89)

These are the momentum and the moment of momentum equations projected intothe free directions of motion, which means, into the tangential directions of theconstraint surfaces. We return to the concept of impressed forces (Section 1.5)

Fe =

∫K

dFe, (1.90)

MeO′ =

∫K

rO′P dFe (1.91)

and apply the definitions of mass center (1.46)

rO′S =1m

∫K

rO′P dm (1.92)

and inertia tensor

ΘO′ =−∫

KrO′PrO′P dm, (1.93)

which can be referred to the mass center by the rule of HUYGENS-STEINER

Θ S =−∫

KrSPrSP dm =−

∫K(rSO′ + rO′P)(rSO′ + rO′P)dm

=ΘO′ −mr2O′S −

∫K

rSO′ rO′P dm−∫

KrO′PrSO′ dm

=ΘO′ +mr2O′S. (1.94)

We finally obtain

Page 44: Introduction To Dynamics

1.8 Newton-Euler Equations for Constrained Systems 35

0 = JTO′[maO′ +m

(˙ω+ ωω

)rO′S −Fe]

+ JTR

⎡⎢⎣ mrO′SaO′︸ ︷︷ ︸

orbital influence

+ ΘO′ ω︸ ︷︷ ︸relative influence

+ ωΘO′ω︸ ︷︷ ︸centrifugal influence

−MeO′

⎤⎥⎦ (1.95)

with the JACOBIANs

JO′ =

(∂vO′

∂ q

), (1.96)

JR =

(∂ω∂ q

). (1.97)

For the manipulation of the centrifugal influences we used the JACOBI identity

i× (j×k) =−j× (k× i)−k× (i× j) (1.98)

with i = rO′P, j = ω , k = ω× rO′P.

1.8.2 System with Multiple Rigid Bodies

The equation (1.95) can easily be extended to a multibody system (MBS) with manyrigid bodies by summing up all n bodies of the multibody system. All magnitudesare indexed by i, and (1.95) becomes

n

∑i=1

JTO′

i

[miaO′

i+mi

(˙ω i + ω iω i

)rO′

iSi−Fe

i

]

+n

∑i=1

JTRi

[mirO′

iSiaO′

i+ΘO′

i,iω i + ω iΘO′

i,iω i −Me

O′i ,i

]= 0. (1.99)

1.8.3 Remarks

1. By choosing the mass center of bodies as reference points, which in many caseswill be possible, the momentum and moment of momentum equations simplifyconsiderably:

0 = JTS [maS −Fe]+ JT

R [ΘSω+ ωΘ Sω−MeS] . (1.100)

2. What is the principal meaning of the relations (1.95)? The row vectors of theJACOBIANs JO′ , JR represent the gradients with respect to q of the constraintsurfaces given by vO′ (q, q, t) and ω (q, q, t). The multiplication of the JACOBIAN

Page 45: Introduction To Dynamics

36 1 Basics

of translation JO′ with the momentum equation and that of the JACOBIAN ofrotation JR with the moment of momentum equation describes the projection tothe constraint surfaces as discussed in Section 1.7.2 (Fig. 1.16). Therefore, therelation (1.95) cuts out all those forces and torques, which are "lost forces" withrespect to the motion. Only those forces and torques are taken into considerationthat contribute to acceleration of the body into the free directions allowed by theconstraints.

3. For every row of JTO′ ,JT

R , we have a scalar product with the corresponding mo-mentum or moment of momentum equation in (1.95), and of course also in (1.99).Each of these scalar products may be evaluated in different coordinate systemswithout changing the result, provided that the two vector components of sucha scalar product are defined in the same coordinate system. This simplifies theevaluation considerably, because we could evaluate the momentum equation in aspace-fixed frame, and the moment of momentum equation in a body-fixed framewith a constant inertia tensor (1.93), for example.

4. For the free motion of a single rigid body without constraints, we may chooseq = r, indicating that the coupling of translation and rotation vanishes. The cor-responding equations are (Section 1.5).

maO′ +m(

˙ω+ ωω)

rO′S = Fe, (1.101)

mrO′SaO′ +ΘO′ω+ ωΘO′ω = MeO′ (1.102)

5. The second equation above represents the moment of momentum equation inthe same form as that of the second part of (1.95), namely given with respectto a moving body-fixed point. Conversely, (1.52) refers to an inertial system. Tounderstand the relationship between both representations, we go back to (1.50)and take into account additionally the relations (1.22) and (1.82). We obtain

LO :=∫

Kr× (rdm) = LO′ +m [rO′ × (vO′ +ω× rO′S)+ rO′S × vO′ ] . (1.103)

Considering EULER’s theorem, the momentum equation, and some standardrules for the vector product, we modify the second term to give the absoluteinertial change

(mvO′)× (vO′ +ω× rO′S)+ rO′ × (maO′ +m( ˙ω+ ωω)rO′S)

+(ω× rO′S)× (mvO′)+ (mrO′S)× aO′

= rO′ ×Fe +(mrO′S)× aO′ . (1.104)

Taking the torque equation into account

MO = MO′ + rO′ ×Fe, (1.105)

Page 46: Introduction To Dynamics

1.8 Newton-Euler Equations for Constrained Systems 37

we realize that the moment of momentum with respect to an inertial frame (1.52)follows from the moment of momentum with respect to a body-fixed frame (1.95)with moving reference O′, or vice versa.

6. Starting with the absolute changes (m KaO′ +m(

K ˙ω+ Kω Kω)

KrO′S) of the mo-mentum and the absolute changes (KΘO′ Kω + Kω KΘO′ Kω) of the moment ofmomentum represented in a body-fixed frame with reference O′, we may repro-duce momentum and moment of momentum by EULER’s theorem to give

Kp = m KvO′ +m Kω KrO′S , (1.106)

KLO′ = KΘO′ Kω . (1.107)

For an arbitrary reference frame, we evaluate momentum and moment of mo-mentum dependent on the relevant angular velocities by

RωR × Rp+ Rp , (1.108)

RωR × RLO′ + RLO′ . (1.109)

7. Assuming that O′ = S is the mass center with (rO′S = 0) or that alternativelyO′ is an inertial body-fixed point (aO′ = 0), the moment of momentum equationbecomes

ΘO′ ω+ω×ΘO′ω = MeO′ , (1.110)

which corresponds to the vector form of the dynamic EULER equation, if weassume a principal coordinate system [41, 23]. It is an important tool of gyrody-namics.

8. The JACOBIAN of rotation has been defined by (1.97)

JR =

(∂ω∂ q

). (1.111)

The JACOBIAN of translation writes (1.85):

JT =

(∂ r∂ q

)=

(∂r∂q

). (1.112)

According to Section 1.4, we know

ω = Y−1(ϕ) ˙ϕ . (1.113)

With ϕ = ϕ(q, t), it is

˙ϕ =

(∂ ϕ∂q

)q+

∂ ϕ∂ t

(1.114)

and analogously also here

Page 47: Introduction To Dynamics

38 1 Basics

(∂ ˙ϕ∂ q

)=

(∂ ϕ∂q

). (1.115)

With the parametrization ϕ = ϕ(q, t), it is ˙ϕ = ˙ϕ(q, q, t) and ω = ω(q, q, t). Itfollows

JR = Y−1(ϕ(q, t))(∂ ˙ϕ∂ q

). (1.116)

The relation

JR =

(∂ ˙ϕ∂ q

)(1.117)

can be used for plane rotations only.

Example 1.7 (Car with a rotating disc). To illustrate the moment of momen-tum equation, we consider an automobile with remote control and a high-speedgyro [29]. Figure 1.17 depicts the operation. The car drives along a left-hand bend,which has the effect of a constraint torque MS(t). Applying the moment of momen-tum equation in a very simple form

ΔLS(t) = MS(t)Δ t,

we state the following:

• On the left side of Fig. 1.17, the moment of momentum vector LS(t) of the high-speed gyro points towards the left side. It will elevated by ΔLS(t). Consequently,the inside wheels lift off the ground.

• On the right side of Fig. 1.17, the moment of momentum vector LS(t) of thehigh-speed gyro points towards the right side. It will elevated by ΔLS(t). Conse-quently, the outside wheels lift off the ground.

ΔLS(t)LS(t)

ΔLS(t)

MS(t) MS(t)

LS(t)

Fig. 1.17. Car with remote control – moment of momentum.

Example 1.8 (Edge-runner mill). Edge-runner mills have been used for thousandsof years, but they are still a very important machine component for example in chem-ical engineering plants. Two wheels are connected by a rigid axis and are driven by

Page 48: Introduction To Dynamics

1.8 Newton-Euler Equations for Constrained Systems 39

ωx

ωz

ω

F r

R

Pz

x

Fig. 1.18. Edge-runner mill.

a vertical axis (Fig. 1.18). It makes sense for this case to introduce an intermedi-ate coordinate system, which is neither body-fixed nor inertial, and which rotateswith the angular vertical velocity component ωx but not with the horizontal angularvelocity component ωz. Nevertheless, it is no inertial system and therefore (1.95)applies. We write

LP +ω×LP = MP, with ω = (ωx,0,0)T , LP = (Aωx,0,Cωz)

T ,

where A and C are the moments of inertia for the x- and z-directions, respectively.For a stationary rotation, we have ωx = 0, ωz = 0 and therefore also LP = 0. Thegyroscopic torque follows with

MP = ω×LP = (0,−Cωxωz,0)T = (Mx,My,Mz)T .

This result may be used to calculate the pressure acting on the ground by the rotatingwheels. Taking into account the rolling condition (rωz +Rωx = 0), we obtain

F =|My|

R= (

Cr)ω2

x .

Example 1.9 (Airplane in a loop). An airplane with one powertrain only is drivenby a propeller-engine unit. It has a moment of momentum L (Fig. 1.19). If such anairplane is starting, landing, or performing a loop, this moment of momentum vectoris rotated around the y-axis generating a small moment of momentum vector ΔL inthe z-direction. This change of moment of momentum gives a yaw to the airplanearound the z-axis, which of course should not happen. Therefore, this additionalmoment of momentum is usually counterbalanced by a trimming tab.

The relevant vectors are the following:

ωP =(ωx,0,0)T propeller and engine ,

ωS =(0,ωy,0)T start, landing, loops ,

L =(Lx,0,0)T moment of momentum of propeller and engine ,

ΔL =(0,0,ΔLz)T moment of momentum change, yaw around z-axis .

Page 49: Introduction To Dynamics

40 1 Basics

x

yz

ωS

L(t)

L(t +Δ t)

ΔLωP(t)

Fig. 1.19. Starting aircraft.

1.9 Lagrange’s Equations

In the Sections 1.7 and 1.8, we presented some direct methods for interconnectedmechanical systems, where interconnection is described by constraints of a certainkinematic-algebraic form. The dynamics of such systems has been evaluated byprojecting the motion into the not-constrained directions. In special cases where theequations of motion can be formulated by minimal coordinates, we satisfy automat-ically all constraints and then we apply any solution procedure. Otherwise we muststay with the projection equations together with the constraints, that is differentialalgebraic equations (DAE), which require special treatment. In the following, weconsider these procedures allowing a determination of motion in the presence ofconstraints [5, 63, 11, 51].

1.9.1 Lagrange’s Equations of the First Kind

We consider a system of n interconnected rigid bodies and describe the velocity ofa single rigid body i by (1.82)

ri = (vO′ + ωrO′P)i =(vO′ + rT

O′Pω)

i . (1.118)

Then, we choose a set of velocity coordinates

zTi =

(vT

O′ ωT)

i ∈ IR6 (1.119)

and determine the virtual velocity vector

δ ri =(I rT

O′P)

i

(δvO′δω

)i. (1.120)

The accelerations follow from the relations (1.83). The virtual power of the com-plete interconnected system is given by the arguments connected with JOURDAIN’sprinciple (Section 1.7)

Page 50: Introduction To Dynamics

1.9 Lagrange’s Equations 41

n

∑i=1

∫Ki

δ rTi (ridmi − dFe

i − dFzi ) = 0 (1.121)

and further on by (1.83) and (1.120)

n

∑i=1

∫Ki

(δvO′δω

)T

i

(I

rO′P

)i

{[aO′ +

(˙ω+ ωω

)rO′P

]dm− dFe − dFz}

i = 0. (1.122)

Considering also (1.95), we obtain

n

∑i=1

(δvO′δω

)T

i

(maO′ +mrT

O′Sω+mωωrO′S −Fe −Fz

mrO′SaO′ +ΘO′ ω+ ωΘO′ω−MeO′ −Mz

O′

)i= 0 (1.123)

or with better organization

n

∑i=1

δ(

vO′ω

)︸ ︷︷ ︸

z

T

i

⎧⎪⎪⎪⎨⎪⎪⎪⎩(

mI mrTO′S

mrO′S ΘO′

)︸ ︷︷ ︸

M

(aO′ω

)︸ ︷︷ ︸

z

+

(mωωrO′SωΘO′ω

)︸ ︷︷ ︸

fg

−(

Fe

MeO′

)︸ ︷︷ ︸

fe

−(

Fz

MzO′

)︸ ︷︷ ︸

fz

⎫⎪⎪⎪⎬⎪⎪⎪⎭

i

=0.

(1.124)

All abbreviations in (1.124) are defined in an IR6-space. With

z :=

⎛⎜⎝

z1...

zn

⎞⎟⎠ , fg :=

⎛⎜⎝

fg1...

fgn

⎞⎟⎠ , fe :=

⎛⎜⎝

fe1...

fen

⎞⎟⎠ , fz :=

⎛⎜⎝

fz1...

fzn

⎞⎟⎠ (1.125)

we arrive at the corresponding vectors in the IR6n-space, and with

M := diag(Mi) ∈ IR6n,6n (1.126)

we define the mass matrix of the total system. From this, we write

δ zT (Mz+ fg − fe − fz) = 0. (1.127)

The virtual velocities δ z cannot be chosen arbitrarily, but they must satisfy the con-straints (1.78), namely

Φ (z, z, t) = W (z, t)T z+ w(z, t) = 0 with z ∈ IR6n, Φ ∈ IRm, m < 6n.(1.128)

On the one hand with (1.79), we have

δΦ = W(z, t)T δ z = 0 (1.129)

Page 51: Introduction To Dynamics

42 1 Basics

including the JACOBIAN of the generalized force directions W ∈ IR6n,m. Equation(1.129) confirms the fact that the columns of W(z, t) are perpendicular to δ z.

On the other hand, the principle of JOURDAIN (1.77) writes

0 =n

∑i=1

∫Ki

δ rTi dFz

i =n

∑i=1

∫Ki

(δvO′δω

)T

i

(I

rO′P

)idFz

i = δ zT fz, (1.130)

which says that the vector of the generalized constraint forces is perpendicular alsoto the virtual velocity δ z. This allows us to represent the constraint forces as a linearcombination of the columns of W giving

fz =−W(z, t)λ with λ ∈ IRm. (1.131)

We combine the two relations (1.131) and (1.127) thus taking into considerationthe constraints. Consequently, the virtual velocities may take any values compatiblewith the constraints. Furthermore, applying the fundamental lemma of variationalcalculus we finally get (6n+m) linear equations for the unknown quantities z ∈ IR6n

and λ ∈ IRm

(M W

WT 0

)(zλ

)+

(fg − fe

w

)=

(00

). (1.132)

Instead of the original constraint, we use the corresponding (hidden) constrainton the acceleration level (Section 1.3.3) for the above equation. Equation (1.132)represents a saddle point relation, because the solution of

(M W

WT 0

)(zλ

)=

(00

)(1.133)

is a saddle point of the quadratic form

(zT λ T )( M W

WT 0

)(zλ

). (1.134)

Solving the first set of equations in (1.132) for z

z =−M−1(fg − fe +Wλ ) (1.135)

and including that into the second set of equations in (1.132) results in an expressionfor the LAGRANGE-multiplier

Page 52: Introduction To Dynamics

1.9 Lagrange’s Equations 43

λ =−(WT M−1W)−1 [WT M−1(fg − fe)− w]. (1.136)

Inserting (1.136) in (1.135) gives a reduced set of differential equations for z,which satisfies all constraints

Example 1.10 (Null space matrix for a pendulum). We once again consider thependulum of Fig. 1.2 with z = r and introduce the minimal coordinate q = ϑ(Fig. 1.20). The Cartesian position r can be parameterized by the minimal coor-dinate q:

r = r(q) =(Rsinϑ −Rcosϑ 0

)T.

Using the coordinates r, we have

mr = mg+Fz,

Φ(r(q)) = r(q)T r(q)−R2 = 0,

where the holonomic-scleronomic constraints are satisfied automatically [21, 28,41]. The JACOBIAN with respect to the mass center

JS =∂r∂q

=(Rcosϑ Rsinϑ 0

)T

gives the actual free motion direction of the point mass. We get

Fz

Ox

Rm

mg

y

rϑ JS

Fig. 1.20. Null space matrix of a pendulum.

Page 53: Introduction To Dynamics

44 1 Basics

0 =∂Φ∂q

=∂Φ∂r

∂r∂q

= 2rT JS.

The JACOBIAN of translation represents the null space matrix due to the specialselection of z = r, and it projects the applied forces (gravitation) into the free direc-tion, namely along the circle around the suspension point of the pendulum. Only thatcomponent can influence actively the energy budget of the pendulum. The equationsof motion for q are

mJTS r = mJT

S g+ JTS Fz︸︷︷︸=0

.

Example 1.10 of a pendulum considers minimal and nonminimal parametrizations,q and z respectively, and introduces a null space matrix ∂z

∂q , the rows of which areperpendicular to the columns of the JACOBIAN of the generalized force directionsW ((1.129) and (1.130)). This more formal property means that the rows of thenull space matrix must have the direction of the tangent vectors to the constraintsurfaces, because the generalized constraint forces can only be part of the normalspace of these surfaces.

Coming back to (1.132) and searching for some parametrization z(q) by a setof coordinates q, we are able to satisfy the constraints automatically. Projecting therelations (1.132) into the free directions of motion by multiplying in a further stepwith the transposed null space matrix ( ∂z

∂q )T from the left-hand side, we arrive at

the NEWTON-EULER equations formulated in minimal coordinates (Section 1.8).Their solution represents a trajectory in time expressed by the minimal coordinatesq. If in addition the generalized constraint forces are needed, for design purposesfor example, they might be evaluated by the relation z(q) applied to (1.136), whichsometimes is called inverse kinetics.

1.9.2 Lagrange’s Equations of the Second Kind

1.9.2.1 Derivation

We restrict our derivation to holonomic constraints, more general constraints fol-low however similar ideas as presented here [28, 50]. Starting with the principle ofd’ALEMBERT (1.71) and assuming a multibody system, we write

n

∑i=1

∫Ki

(rdm− dFe)Ti δri = 0. (1.137)

To carry out some manipulation with this expression, we consider the kinetic energyof all single bodies

T =n

∑i=1

Ti =n

∑i=1

12

∫Ki

rTi ridmi. (1.138)

Page 54: Introduction To Dynamics

1.9 Lagrange’s Equations 45

For the following calculations, we consider a single rigid body and combine laterthe results for a multibody system. We modify the acceleration in (1.137) to thefollowing form

∫Ki

rTi dmiδri =

ddt

∫Ki

rTi dmiδri −

∫Ki

rTi dmiδ ri

=ddt

∫Ki

∂∂ ri

(12

rTi ridmi

)δri − δ

∫Ki

(12

rTi ridmi

). (1.139)

This reshaping uses the exchangeability of integration, differentiation, and variationdue to the linear properties of the equations and due to constant areas of integra-tion [20, 28]. With the minimal coordinates ri = ri (q, t) (ri ∈ IR3, q ∈ IR f ), (1.85),and variation with respect to δq, we get (δ t = 0):

δri =

(∂ri

∂q

)δq =

(∂ ri

∂ q

)δq. (1.140)

Together with (1.138), the relation (1.139) gives

∫Ki

rTi dmiδri =

ddt

∫Ki

∂∂ ri

(12

rTi ridmi

)(∂ ri

∂ q

)δq− δTi

=ddt

∫Ki

∂∂ q

(12

rTi ridmi

)δq− δTi

=ddt

(∂Ti

∂ qδq)− δTi. (1.141)

We vary the kinetic energy Ti = Ti(q, q, t) without considering time (δ t = 0, seeSection 1.3.4)

δTi =∂Ti

∂qδq+

∂Ti

∂ qδ q (1.142)

and get

∫Ki

rTi dmiδri =

[ddt

(∂Ti

∂ q

)− ∂Ti

∂q

]δq+

∂Ti

∂ q

[ddt

(δq)− δ q]. (1.143)

Again exchanging differentiation and variation, the last term in (1.143) vanishes( d

dt (δq)− δ q = 0) [28, 50]. The virtual work of the applied forces dFei is

δW ei =

∫Ki

(dFei )

T δri =

∫Ki

(dFei )

T(∂ri

∂q

)δq = QT

i δq (1.144)

where

Page 55: Introduction To Dynamics

46 1 Basics

Qi =∫

Ki

(∂ri

∂q

)T

dFei =

∫Ki

(∂ ri

∂ q

)T

dFei (1.145)

are the generalized forces of the body Ki. Combining (1.137), (1.143), and (1.145),we come out with

n

∑i=1

[ddt

(∂Ti

∂ q

)−(∂Ti

∂q

)−QT

i

]δq = 0. (1.146)

With q being minimal coordinates and therefore δq being arbitrary, we achieveLAGRANGE’s equations of motion of the second kind by the fundamental lemma ofvariational calculus

n

∑i=1

[ddt

(∂Ti

∂ q

)−(∂Ti

∂q

)−QT

i

]= 0. (1.147)

The generalized forces Qi can be subdivided into conservative and nonconservativeforces, where the first ones, QK , can be derived from a potential V (Section 1.6):

QKi =−(∂Vi

∂q

)T

. (1.148)

Most mechanical systems include generalized conservative forces QK as well asgeneralized nonconservative forces QNK . Introducing

V =n

∑i=1

Vi, (1.149)

QNK =n

∑i=1

∫Ki

(∂ri

∂q

)T

dFeNKi

, (1.150)

we get f LAGRANGE’s equations of the second kind in the following compact form:

ddt

(∂T∂ q

)−(∂T∂q

)+

(∂V∂q

)= QT

NK . (1.151)

The special case of ddt

(∂T∂ qs

)= 0 (cyclic momenta) will be treated in Section

4.4.3 and in the following examples.

1.9.2.2 Remarks on Evaluation

Lagrange’s equations (1.151) have been derived with respect to arbitrary rigid me-chanical bodies. Their derivation by the virtual work principle and their formulation

Page 56: Introduction To Dynamics

1.9 Lagrange’s Equations 47

in terms of energy even allows us to apply them to physical systems of a completelydifferent nature [25]. We restrict our considerations to rigid bodies.

The kinetic energy of a single body writes by combination of (1.138), (1.82), and(1.93):

Ti =12

∫Ki

[vO′ + ωrO′P]Ti [vO′ + ωrO′P]i dmi

=

{12

mvTO′vO′ +mvT

O′ωrO′S +12

∫Ki

[−rO′Pω ]T [−rO′Pω ]dm

}i

=

{12

mvTO′vO′ +mvT

O′ωrO′S +12ωTΘO′ω

}i. (1.152)

This formula has three parts: a term concerning translation, a coupling term concern-ing both translation and rotation, and a rotational term. The coupling term vanishesfor the mass center as the reference O′

i = Si and thus rO′iSi

= 0. Then, we get

Ti =

{12

mvTS vS +

12ωTΘ Sω

}i. (1.153)

The potential energy very often consists of a general spring potential [15] and apotential for gravity giving

Vf =12

(rF1F2 − r0

F1F2

)TC(rF1F2 − r0

F1F2

), (1.154)

Vg =−mgT rOS. (1.155)

The vector rF1F2 is the distance of the spring end points, r0F1F2

is the distance forthe springs without load, C a positive-definite matrix of the spring constants, g thegravity acceleration, and rOS the vector from an inertial point to the mass centers,of course evaluated in the same coordinate frame. For multibody systems anotherrepresentation is frequently applied, namely

Vf =12

c(‖ rF1F2 ‖ −l0)2 , (1.156)

Vf =12

c(α−α0)2 (1.157)

suitable for translational springs with the spring constant c and the load-free lengthl0 and for rotational springs with fixed axis, a turning angle α , and a load-free an-gular displacement of α0.

Nonconservative forces are related to their point of application using (1.85).

QNK =n

∑i=1

∫Ki

(∂ri

∂q

)T

dFeNKi

=n

∑i=1

∫Ki

(∂ ri

∂ q

)T

dFeNKi

(1.158)

Page 57: Introduction To Dynamics

48 1 Basics

For a rigid body we apply (1.82) and use as a reference point O′ regarding in additionthe definitions of Section 1.8. It is

QNK =n

∑i=1

∫Ki

[(∂vO′

∂ q

)T

+

(∂ω∂ q

)T

rO′P

]

i

dFeNKi

=n

∑i=1

(JT

O′iFe

NKi+ JT

RiMe

NKi ,O′). (1.159)

This procedure introduces a torque

MeNKi ,O′ =

∫Ki

[rO′P]idFeNKi

(1.160)

meaning that applied free torques can also be projected by the JACOBIANs of rota-tion into the free directions of motion. For forces, one uses directly, if possible, theJACOBIANs of translation with respect to the points of force applications.

Lagrange’s equations of the second kind offer an analytical route to the equationsof motion, provided, we are able to find a set of minimal coordinates. For manypractical cases this will be possible, and then we proceed stepwise in the followingform:

• Look for a convenient set of minimal coordinates q ∈ IR f with respect to theproblem,

• evaluate the kinetic energies Ti and the potential energies Vi,• calculate the generalized nonconservative forces and torques QNK ,• evaluate Lagrange’s equations of the second kind.

1.9.2.3 Examples

Example 1.11 (KEPLER’s laws of planetary trajectories). The famous three lawsof Johannes KEPLER write:

1. The orbit of every planet is an ellipse with the sun at one of the two foci(ellipse law).

2. A line joining a planet and the sun sweeps out equal areas during equalintervals of time (area law).

3. The square of the orbital period of a planet is directly proportional to thecube of the semimajor axis of its orbit.

We prove the second law using LAGRANGE’s equations of the second kind(Fig. 1.21). We choose minimal coordinates q=(r,ϕ)T and describe the vector fromthe point M to the planet mass m and the angular velocity of the moving coordinatesystem by

Page 58: Introduction To Dynamics

1.9 Lagrange’s Equations 49

Kx

r

v

m

M

ϕ

Ix

Kz

Ky

Iy

Fig. 1.21. KEPLER’s laws.

Kr =(r 0 0

)T,

Kω =(0 0 ϕ

)T.

The absolute velocity of m then writes

Kv =(r 0 0

)T+(0 0 ϕ

)T × (r 0 0)T

=(r 0 0

)T+(0 rϕ 0

)T=

⎛⎝1 0

0 r0 0

⎞⎠ q,

giving the kinetic energy as

T =12

m KvTKv =

12

qT(

m 00 mr2

)q.

The gravity potential of the central force field is

V = γMmr2

with the gravitational constant γ . Nonconservative forces do not exist, thus La-grange’s equations result in

(m 00 mr2

)q+

(−2γ mMr3

0

)=

(00

).

The second row of this equation indicates a cyclic coordinate accompanied by acyclic (generalized) momentum, which is

Page 59: Introduction To Dynamics

50 1 Basics

p =∂T∂ ϕ

= mr2ϕ

in the direction of ϕ and with the properties ∂T∂ϕ = 0 and ∂V

∂ϕ = 0. From this, we getKEPLER’s second law

A =12

r2ϕ =p

2m.

Similar results can be obtained for a particle in any mechanical field of forces, forexample for linear-elastic bearings V = 1

2 cr2 with a spring constant c.

Example 1.12 (Spherical pendulum). We consider the spherical pendulum fromExample 1.4; we also want to identify the cyclic coordinate as in Example 1.11(Fig. 1.22). We choose the minimal coordinates q = (ψ ,ϑ)T . With respect to theK-system the vector from O to P writes

KrOP =(0 0 −R

)T.

The angular velocity of the K-system can be described by EULER’s sequence ofrotation:

Kω =(ϑ ψ sinϑ ψ cosϑ

)T.

m

v

ϑ

ψ

O

P

g

Kz

Ky

Kx

R

Iy

Iz

Ix

Fig. 1.22. Spherical pendulum.

Page 60: Introduction To Dynamics

1.9 Lagrange’s Equations 51

With this we get the absolute velocity of the mass m with respect to the K-system

Kv = Kω× KrOP =(−Rψ sinϑ Rϑ 0

)T

and finally the kinetic energy as

T =12

m KvTKv =

12

qT(

mR2 sin2 ϑ 00 mR2

)q.

Introducing Ig =(0 0 −g

)T, the gravitational potential writes

V =−mgRcosϑ .

Considering all this, LAGRANGE’s equations of the second kind come out with

(mR2 sin2 ϑ 0

0 mR2

)q+

(2mR2 sinϑ cosϑϑψ

0

)︸ ︷︷ ︸

ddt

(∂T∂ q

)

−(

0mR2 sinϑ cosϑψψ

)︸ ︷︷ ︸

∂T∂q

+

(0

mgRsinϑ

)︸ ︷︷ ︸

∂V∂q

=

(00

).

Because of ∂T∂ψ = ∂V

∂ψ = 0, the angle ψ is a cyclic coordinate, and consequently thecyclic momentum

(∂T∂ψ

)= (mR2 sin2 ϑ)ψ

in the direction of ψ will be constant. The projection of the string onto the horizontalplane sweeps out equal areas in equal intervals of time (Example 1.11), comparableto KEPLER’s second law.

Example 1.13 (Differential gear). We consider the differential gear of Fig. 1.23and investigate the behavior after a change in load. The differential gear consists ofa drive gear 1, a hypoid gear 2, two driven gears 3,4, and two differential gears 5,6(satellite gears) with the angular positions

zT =(ϕ1 ϕ2 ϕ3 ϕ4 ϕ5 ϕ6

)T.

The following kinematic conditions hold

Φ (z) =

⎛⎜⎜⎝

ϕ1 − a1ϕ2

ϕ2 − 12 (ϕ3 +ϕ4)

ϕ5 − a2(ϕ3 −ϕ4)ϕ6 + a2(ϕ3 −ϕ4)

⎞⎟⎟⎠= 0.

Page 61: Introduction To Dynamics

52 1 Basics

M1

M4M3

2

1

6

543

Fig. 1.23. Differential gear.

The magnitudes a1 and a2 are transmission ratios. The system has six bodies, fourconstraints, and therefore two degrees of freedom. The most convenient minimalcoordinates are

qT =(q1 q2

)T=(ϕ3 ϕ4

)T.

They allow an easy representation of the other four coordinates

z = z(q) =

⎛⎜⎜⎜⎜⎜⎜⎝

ϕ1

ϕ2

ϕ3

ϕ4

ϕ5

ϕ6

⎞⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎝

12 a1 (q1 + q2)

12 (q1 + q2)

q1

q2

a2 (q1 − q2)−a2 (q1 − q2)

⎞⎟⎟⎟⎟⎟⎟⎠.

From the constraints and the last relation, we get the directions of the generalizedforces in the form

W =

⎛⎜⎜⎝

1 −a1 0 0 0 00 1 − 1

2 − 12 0 0

0 0 −a2 +a2 1 00 0 +a2 −a2 0 1

⎞⎟⎟⎠

T

=

(∂Φ∂z

)T

.

The JACOBIAN of z writes

Page 62: Introduction To Dynamics

1.9 Lagrange’s Equations 53

∂z∂q

=

⎛⎜⎜⎜⎜⎜⎜⎝

a12

a12

1/2 1/21 00 1a2 −a2

−a2 a2

⎞⎟⎟⎟⎟⎟⎟⎠.

With this, we project the nonconservative applied forces

fe =(M1 0 −M3 −M4 0 0

)T

into the directions of the two minimal coordinates

QNK =

(∂z∂q

)T

fe =

( a12 M1 −M3a12 M1 −M4

).

Using again ( ∂z∂q ) brings the kinetic energy into the form

T =12

zT Mz =12

qT

[(∂z∂q

)T

M(

∂z∂q

)]

︸ ︷︷ ︸M∗=

(M∗

11 M∗12

M∗12 M∗

22

)

q

with M = diag(Ji). The transverse moments of inertia of the two bodies 5,6 areincluded in the moment of inertia of body 2. The mass matrix projected into thedirection of the two minimal coordinates is given by

M∗11 = J1

(a1

2

)2+

14

J2 + J3 + a22 (J5 + J6) ,

M∗12 = J1

(a1

2

)2+

14

J2 − a22 (J5 + J6) ,

M∗22 = J1

(a1

2

)2+

14

J2 + J4 + a22 (J5 + J6) .

Hence,

T =12

M∗11ϕ

23 +M∗

12ϕ3ϕ4 +12

M∗22ϕ

24 .

If J3 = J4, then M∗11 = M∗

22. Generally it is M∗11 > M∗

12.

Page 63: Introduction To Dynamics

54 1 Basics

Evaluating LAGRANGE’s equations of the second kind with M∗11 = M∗

22 results in

(∂T∂ ϕ3

)= M∗

11ϕ3 +M∗12ϕ4,

(∂T∂ϕ3

)= 0,

(∂T∂ ϕ4

)= M∗

11ϕ4 +M∗12ϕ3,

(∂T∂ϕ4

)= 0

and from this the equations of motion in minimal coordinates:

M∗11ϕ3 +M∗

12ϕ4 =12

M1a1 −M3,

M∗11ϕ4 +M∗

12ϕ3 =12

M1a1 −M4.

Usually one has to take into account the torques Mi = Mi (ϕi) from some com-plicated relations that can be treated only numerically. For our considerations, wesimplify that and discuss the following situations.

1. Stationary motion:

ϕ3 = ϕ4 = 0

being possible only for M3 = M4 =12 M1a1.

2. Acceleration of the driven bodies (resolution for ϕ3 and ϕ4):

ϕ3 =12 M1a1 (M∗

11 −M∗12)−M∗

11M3 +M∗12M4

M∗211 −M∗2

12

,

ϕ4 =12 M1a1 (M∗

11 −M∗12)−M∗

11M4 +M∗12M3

M∗211 −M∗2

12

.

3. Acceleration of the driving body:

ϕ1 =a1

2(ϕ3 + ϕ4) =

a1

2

(M1a1 −M3 −M4

M∗11 +M∗

12

).

4. Sudden load change, for example ΔM3:

M3 = M30 +ΔM3

results in ϕi = ϕi0 +Δϕi. Starting with a stationary motion (ϕ3 = ϕ4 = 0) afterthe load change, we get

Page 64: Introduction To Dynamics

1.10 Hamilton’s Equations 55

Δϕ3 =− M∗11ΔM3

M∗211 −M∗2

12

< 0 deceleration,

Δϕ4 =+M∗

12ΔM3

M∗211 −M∗2

12

> 0 acceleration,

Δϕ1 =−a1

2

(ΔM3

M∗211 −M∗2

12

)< 0.

As M∗11 > M∗

12 it is | Δϕ3 |>| Δϕ4 |.

1.10 Hamilton’s Equations

HAMILTON’s principle and connected with it, HAMILTON’s canonical equations,were formerly of greater significance in the fields of purely theoretical mechan-ics but they are now moving into the fields of engineering applications due to theincreasing significance of nonlinear dynamics in many fields of technology [71].Therefore, it makes sense to discuss this approach. In the following section, we useminimal coordinates.

1.10.1 Hamilton’s Principle

We start with relations (1.137), (1.141), and (1.144) from Section 1.9:

n

∑i=1

∫Ki

(rdm− dFe)Ti δri = 0, (1.161)

∫Ki

rTi dmiδri =

ddt

(∂Ti

∂ qδq)− δTi, (1.162)

∫Ki

(dFei )

T δri = δW ei . (1.163)

Additionally, we consider the total kinetic energy, the total virtual work of the ap-plied forces

T =n

∑i=1

Ti, (1.164)

δW e =n

∑i=1

δW ei (1.165)

and the generalized momentum as the derivation of the kinetic energy with respectto the generalized minimal velocities (Example 1.11):

pT =

(∂T∂ q

). (1.166)

Page 65: Introduction To Dynamics

56 1 Basics

These steps result in the central equation of LAGRANGE

("Zentralgleichung") [50]

ddt

(pT δq

)− δT − δWe = 0. (1.167)

Integrating this relation from t1 to t2 gives

∫ t2

t1δ (T +We)dt =

(pT δq

)∣∣t2t1. (1.168)

Assuming for the virtual displacements δq(t1) = 0 and δq(t2) = 0, we come to theform

∫ t2

t1(δT + δW e)dt = 0. (1.169)

For conservative systems, the applied forces might be derived from potentials (Sec-tion 1.6):

(dFe)Ti =−∂ (dVi)

∂ri, (1.170)

and therefore

δW ei =

∫Ki

(dFei )

T δri =−δVi. (1.171)

With

δV =n

∑i=1

δVi, (1.172)

it follows∫ t2

t1δ (T −V )dt =:

∫ t2

t1δL = 0 (1.173)

with the LAGRANGE function L = T −V . The variations δL must be compati-ble with the constraints. Assuming holonomic constraints, as we have done forD’ALEMBERT’s principle and which allow the exchangeability of integration andvariation, we finally get

Page 66: Introduction To Dynamics

1.10 Hamilton’s Equations 57

HAMILTON’s principle as a variational problem [28]:

δ∫ t2

t1Ldt = 0 or

∫ t2

t1Ldt → stationary. (1.174)

HAMILTON’s principle is an integral principle which alters trajectory elements,whereas D’ALEMBERT’s principle represents a differential principle that comparesneighbouring states. The varied magnitudes of the integral principles possess thedimension of an action (= energy · time). Thus, the integral principles are alsocalled principles of least action [22]. To solve (1.174), we consider a bundle offunctions

qε(t) = q0(t)+ εηq(t) (1.175)

with the parameter ε and take into account the compatibility condition

ηq(t1) = ηq(t2) = 0. (1.176)

The relation (1.174) means

0 =d

(∫ t2

t1Ldt

)∣∣∣∣ε=0

=

∫ t2

t1

(∂L∂q

ηq(t)+∂L∂ q

ηq(t)

)dt

=

∫ t2

t1

(∂L∂q

ηq(t)−ddt

(∂L∂ q

)ηq(t)

)dt +

[∂L∂ q

ηq(t)

]t2

t1︸ ︷︷ ︸=0

. (1.177)

With ηq being arbitrary, we come out with the EULER-LAGRANGE equations

ddt

(∂L∂ q

)− ∂L

∂q= 0T , (1.178)

which are LAGRANGE’s equations of the second kind with L = T −V (1.151).

1.10.2 Hamilton’s Canonical Equations

To derive HAMILTON’s canonical equations, we represent the mechanical systemby minimal coordinates q and by generalized momentum coordinates p instead ofthe usually considered velocity coordinates q. We start with LAGRANGE’s equa-tions of the second kind and restrict our considerations to conservative, holonomic-scleronomic systems. More general presentations may be found in [28].

Page 67: Introduction To Dynamics

58 1 Basics

LAGRANGE’s equations of the second kind (1.178) together with (1.166) resultin

∂L∂q

= pT . (1.179)

The LAGRANGE function depends on q and q giving

δL =

(∂L∂q

)δq+

(∂L∂ q

)δ q = pTδq+pTδ q, (1.180)

on the one hand. On the other hand, the virtual change of the product(pT q

)results

in

δ(pT q

)= pTδ q+ qT δp. (1.181)

At this point, we introduce the HAMILTON function

H (q,p) = pT q−L, (1.182)

determine formally its variation

δH =

(∂H∂q

)δq+

(∂H∂p

)δp (1.183)

and compare this relation with the difference between (1.181) and (1.180)

δ(pT q−L

)= qTδp− pT δq, (1.184)

then finally coming to the canonical equations of motion of

William HAMILTON (1805-1865):

qT =∂H∂p

, pT =−∂H∂q

, (1.185)

which are 2 f differential equations of first order, the so-called HAMILTONIAN

system.The HAMILTON function H possesses a clear interpretation. We obtain from

(1.182) with (1.166) and

T =12

qT Mq

the following result

Page 68: Introduction To Dynamics

1.11 Practical Considerations 59

H = pT q−L = qT Mq− (T −V) = 2T − (T −V) = T +V.

The HAMILTON function simply is the sum of kinetic and potential energies, whichmeans the total energy of a mechanical system.

1.11 Practical Considerations

Whatever method we use in mechanics, we always end up with a set of nonlin-ear differential equations of motion of first or second order, which are linear withrespect to the accelerations but nonlinear with respect to velocities, positions, andorientations. These equations represent the mathematical model of the problem andare an intermediate activity of our sequence (real-world problem - mechanical mod-eling - mathematical modeling - numerical modeling - simulation). It should bekept in mind that mathematics can give only these results and information from abasis, which we have defined as both assumptions and constraints of the mechanicalmodel. Therefore, establishing the mechanical model requires extreme care, empir-ical knowledge, and instinctive feeling combined with a good understanding of themechanical problems, at least from a qualitative point of view. With respect to thisstep, large expenditure can be generated but also omitted, if done intelligently.

The degrees of freedom (DOF) as expressed by minimal coordinates determinethe size of the mathematical previously also of the mechanical model. Additionalsimplifications might be feasible by linearization, by using invariants of motion likeenergy integrals of conservative systems, or by modifying the equations of motion,for example by transforming the differential equations. In any case, we should tryto find a set of minimal coordinates and if this is not possible, we have to add therelevant constraints, but again trying to find a minimalistic formulation.

One could say that this represents an old-fashioned procedure in the face of mod-ern commercial computer codes, but it does not for two reasons. First, technologicalprogress is not possible without understanding the underlying problems, and theprocess described above helps significantly in increasing our understanding. Sec-ond also for computer codes, the users have to establish a mechanical model, andit is advisable that this model is carefully established on the basis of the thoughtsdiscussed above. The quality of the results depends on such considerations. Com-mercial codes usually do not use a minimal formulation but a structure that leads tofast and efficient numerical algorithms [5, 63]. Interpretation of the results, however,depends on a thorough understanding of the mechanical model and of the real-worldmachine.

What we have mentioned in no way depends on the choice of the mechanicallaws that were applied, NEWTON-EULER, LAGRANGE, or HAMILTON. However,the choice of the mechanical fundament for the derivation of the equations of motionconsiderably influences the expenditure in establishing these equations. This has tobe considered very carefully; we discuss it in the following.

Page 69: Introduction To Dynamics

60 1 Basics

For the derivation of the equations of motion, we must provide a kinematicfoundation, by the way one of the most frequent sources of errors and mistakes(Section 1.4). In the first step, we choose coordinate systems; this is not to be un-derestimated, because a good choice helps to reduce effort, a bad choice increaseseffort. In the second step, we determine positions, orientations, velocities, and ac-celerations, on the basis of these coordinate systems. In a third step, we try to findminimal coordinates, and, if necessary, we establish the constraints. Velocities andaccelerations are usually needed in an absolute form and with respect to inertialand body-fixed coordinate systems. This depends on the problem under consider-ation. Systems with small degrees of freedom might conveniently be representedwith respect to inertial coordinates, which simplifies all derivations with respect totime. The study of kinematics does not depend on the methods of kinetics, whichthemselves offer a broad variety of possibilities.

Momentum and moment of momentum equations according to NEWTON-EULER

without application of any other mechanical principle can be preferentially used forsmall systems and in combination with EULER’s cut principle. The result is a setof relations for each free body diagram including all reaction forces and torques.The elimination of these reaction forces without D’ALEMBERT’s or JOURDAIN’sprinciple might be cumbersome for large systems. Therefore, such a direct methodmakes sense only for smaller systems of clear kinematic structure.

Combining the momentum and moment of momentum equations with JOUR-DAIN’s principle (Section 1.8) results in a very efficient method, which emergedafter a very long period of discussion within the multibody systems community.The constraint forces can be eliminated by JOURDAINs principle (1.77), and theequations of motion finally include only the applied forces. The JACOBIANs oftranslation and rotation project the motion into the free directions as defined bythe constraints. After solving these equations, we additionally can go back to cer-tain free body diagrams for an evaluation of reaction forces (inverse kinetics).More flexibility with respect to constraint forces or contact forces, for example,is offered by LAGRANGE’s equations of the first kind. In summary, we state thatNEWTON-EULER together with D’ALEMBERT or JOURDAIN on the one hand, andLAGRANGE on the other hand, are presently the best procedures for treating largedynamic systems [51].

The application of LAGRANGE’s equations of the second kind or of HAMIL-TON’s principle requires evaluation of the kinetic and potential energies expressedby generalized coordinates (Section 1.9). Knowing the energies, the equations ofmotion follow by corresponding differentiations. This process is the most impor-tant argument against an automatized application of the analytical methods as givenby LAGRANGE and HAMILTON, because computing time for differentiation issignificantly greater than for the determination of the JACOBIANs for the NEW-TON-EULER method. Nevertheless, analytical methods are very convenient for me-chanical systems with a small number of DOF, especially in those cases where atreatment "by hand" is achievable.

Page 70: Introduction To Dynamics

1.11 Practical Considerations 61

The following table provides a survey of the methods presented in this chapter.

• momentum- (NEWTON) and moment of momentum (EULER) (Chap-ter 1.5):

dpdt

= F,dLO

dt= MO.

• principle of d’ALEMBERT (Section 1.7):∫

K(rdm− dFe)T δr = 0.

• principle of JOURDAIN (Section 1.7):∫

K(rdm− dFe)T δ r = 0.

• NEWTON-EULER equations (Section 1.8):

(∂vO′

∂ q

)T [mvO′ +m

(˙ω+ ωω

)rO′S −Fe]

+

(∂ω∂ q

)T [mrO′SvO′ +ΘO′ ω+ ωΘO′ω−Me

O′]= 0.

• LAGRANGE’s equations of the first kind (Section 1.9):(

M WWT 0

)(zλ

)+

(fg − fe

w

)=

(00

).

• LAGRANGE’s equations of the second kind (Section 1.9):

[ddt

(∂T∂ q

)− ∂T

∂q+

∂V∂q

]T

= QNK .

• HAMILTON’s canonical equations (Section 1.10):

qT =∂H∂p

, pT =−∂H∂q

.

Example 1.14 (Robot with three revolute joints). We evaluate the equations ofmotion of the robot of Fig. 1.24 applying two methods, first LAGRANGE’s equa-tions of the second kind and second the NEWTON-EULER equations in minimalcoordinates. According to the three revolute joints, the robot has three degrees offreedom, one rotation around a vertical axis and two rotations around horizontal

Page 71: Introduction To Dynamics

62 1 Basics

axes. The two links have the lengths l1, l2 and the distances s1, s2 from their centersof mass. The joints are loaded by torque motors with the torques M1, M2, and M3.The relevant acceleration of gravity is Ig =

(0 0 −g

)T .

For minimal coordinates, we choose the angles q :=(q1 q2 q3

)T . The body-fixed coordinates K1 and K2 are located at the centers of mass S1, S2 of the links.The necessary coordinate transformations are thus given by

AIK1 =

⎛⎝cosq1 cosq2 −sinq1 cosq1 sinq2

sin q1 cosq2 cosq1 sinq1 sinq2

−sinq2 0 cosq2

⎞⎠

=

⎛⎝cosq1 −sinq1 0

sin q1 cosq1 00 0 1

⎞⎠⎛⎝ cosq2 0 sinq2

0 1 0−sinq2 0 cosq2

⎞⎠ ,

AK1K2 =

⎛⎝ cosq3 0 sinq3

0 1 0−sinq3 0 cosq3

⎞⎠ .

Iz

O

q2

M1 M2

M3

s1

S1

q3

s2

S2

Ix Iy

K1x

K1y

K1 z

l1

K2x

K2y

K2 z

l2

q1

Fig. 1.24. Robot with three revolute joints.

Page 72: Introduction To Dynamics

1.11 Practical Considerations 63

The absolute velocities and accelerations of the two links are (in body-coordinates)

K1ω1 =

⎛⎝−q1 sinq2

q2

q1 cosq2

⎞⎠=

⎛⎝−q1 sinq2

0q1 cosq2

⎞⎠+

⎛⎝ 0

q2

0

⎞⎠ ,

K2ω2 =

⎛⎝−q1 sin(q3 + q2)

q3 + q2

q1 cos(q3 + q2)

⎞⎠= AT

K1K2 K1ω1 +

⎛⎝ 0

q3

0

⎞⎠ ,

K1vS1 = s1

⎛⎝ 0

q1 cosq2

−q2

⎞⎠= K1

ω1 ×⎛⎝s1

00

⎞⎠ ,

K2vS2 =

⎛⎝ q2l1 sinq3

q1 (l1 cosq2 + s2 cos(q3 + q2))−q2l cosq3 − (q2 + q3)s2

⎞⎠

= K2ω2 ×

⎡⎣⎛⎝l1 cosq3

0l1 sin q3

⎞⎠+

⎛⎝s2

00

⎞⎠⎤⎦+

ddt

⎛⎝l1 cosq3

0l1 sin q3

⎞⎠ .

With respect to the moments of inertia, we choose principal axes resulting in

KiΘ Si,i =

⎛⎝Ai 0 0

0 Bi 00 0 Ci

⎞⎠ .

We have to take into account nonconservative forces and torques for both methods,for LAGRANGE’s equations of the second kind and for NEWTON-EULER equations.With the JACOBIANs

(∂ K1ω1

∂ q

)T

=

⎛⎝−sinq2 0 cosq2

0 −1 00 0 0

⎞⎠ ,

(∂ K2

ω2

∂ q

)T

=

⎛⎝−sin(q3 + q2) 0 cos(q3 + q2)

0 −1 00 1 0

⎞⎠

and the torques

K1M1 =

⎛⎝−M1 sinq2

M3 +M2

M1 cosq2

⎞⎠ , K2

M2 =

⎛⎝ 0

M3

0

⎞⎠ ,

we obtain

Page 73: Introduction To Dynamics

64 1 Basics

QNK =

(∂ K1ω1

∂ q

)T

K1M1 +

(∂ K2ω2

∂ q

)T

K2M2 =

⎛⎝M1

M2

M3

⎞⎠ .

1. LAGRANGE’s equations of the second kindTo derive the equations of motion, we need the energies. The kinetic energy Tresults from

T =2

∑i=1

Ti =12

2

∑i=1

(mi Ki

vTSi Ki

vSi + KiωT

i KiΘ Si,i Ki

ωTi

)

=12

m1s21

(q2

1 cos2 q2 + q22

)+

12

(A1q2

1 sin2 q2 +B1q22 +C1q2

1 cos2 q2)

+12

m2

{q2

1 [l1 cosq2 + s2 cos(q3 + q2)]2 + q2

2

[l21 + 2l1s2 cosq3

]

+2q2q3l1s2 cosq3 +(q3 + q2)2 s2

2

}

+12

{A2q2

1 sin2 (q3 + q2)+B2 (q3 + q2)2 +C2q2

1 cos2 (q3 + q2)},

and the potential energy is

V =−2

∑i=1

mi IgT

IrSi =−m1gs1 sinq2 −m2g [l1 sinq2 + s2 sin(q3 + q2)] .

Then, the equations of motion follow from

ddt

(∂T∂ q

)− ∂T

∂q+

∂V∂q

= QTNK

with

∂T∂ q1

= m1s21q1 cos2 q2 +A1q1 sin2 q2 +C1q1 cos2 q2

−m2q2 [l1 cosq2 + s2 cos(q3 + q2)]2

+A2q1 sin2 (q3 + q2)+C2q1 cos2 (q3 + q2) ,

∂T∂ q2

= m1s21q2 −B1q2 −m2

(l21 + 2l1s2 cosq3

)q2

−m2l1s2 cosq3q3 − (q3 + q2) s22m2 −B2 (q3 + q2) ,

∂T∂ q3

= m2l1s2 cosq3q2 +m2s22 (q3 + q2)+B2 (q3 + q2) ,

Page 74: Introduction To Dynamics

1.11 Practical Considerations 65

ddt

(∂T∂ q1

)=[(

m1s21 +C1

)cos2 q2 +A1 sin2 q2 +A2 sin2 (q3 + q2)

+C2 cos2 (q3 + q2)+m2 (l1 cosq2 + s2 cos(q3 + q2))2]

q1

−{−(A1 −C1 −m1s21 −m2l2

1

)sin2q2 + 2m2l1s2 sin(q3 + 2q2)

+(C2 −A2 +m2s2

2

)sin2(q3 + q2)

}q1q2{(

A2 −C2 −m2s22

)sin2(q3 + q2)

−2m2l1s2 cosq2 sin(q3 + q2)} q1q3,

ddt

(∂T∂ q2

)=−[m1s2

1 +B1 +m2(l21 + 2l1s2 cosq3

)+m2s2

2 +B2]

q2

− [m2l1s2 cosq3 +m2s22 +B2

]q3 + 2m2l1s2 sinq3q2q3

+m2l1s2 sinq3q23,

ddt

(∂T∂ q3

)=[m2l1s2 cosq3 +m2s2

2 +B2]

q2 +[B2 +m2s2

2

]q3

−m2l1s2 sinq3q2q3,

∂T∂q1

= 0,

∂T∂q2

=

[−1

2A1 sin2q2 +

12

C1 sin 2q2 +12

m1s21 sin2q2

+12(C2 −A2) sin2(q3 + q2)+

12

m2l21 sin2q2

+m2l1s2 sin(q3 + 2q2)+12

m2s22 sin2(q3 + q2)

]q2

1,

∂T∂q3

=

[12

(A2 −C2 −m2s2

2

)sin2(q3 + q2)

−m2l1s2 cosq2 sin(q3 + q2)] q21

−m2l1s2 sinq3q22 +m2l1s2 sinq3q2q3

and

∂V∂q1

= 0,

∂V∂q2

= m1gs1 cosq2 +m2g(l1 cosq2 + s2 cos(q3 + q2)) ,

∂V∂q3

=−m2gs2 cos(q3 + q2) .

Page 75: Introduction To Dynamics

66 1 Basics

With

M11 =(m1s2

1 +C1)

cos2 q2 +A1 sin2 q2 +A2 sin2 (q3 + q2)

+C2 cos2 (q3 + q2)+m2 [l1 cosq2 + s2 cos(q3 + q2)]2 ,

M22 = m1s21 +m2

(l21 + 2l1s2 cosq3 + s2

2

)+B1 +B2,

M33 = m2s22 +B2,

M23 = M32 =−[m2(l1s2 cosq3 + s2

2

)+B2

],

h1 =−[−(A1 −C1 −m1s21 −m2l2

1

)sin2q2 + 2m2l1s2 sin (q3 + 2q2)

+(m2s2

2 +C2 −A2)

sin2(q3 + q2)]

q1q2 +[(

A2 −C2 −m2s22

)sin2(q3 + q2)

−2m2l1s2 cosq2 sin(q3 + q2)] q1q3,

h2 =

[−(m1s2

1 −A1 +C1 +m2l21

) 12

sin2q2 +12

(A2 −C2 −m2s2

2

)sin2(q3 + q2)

−m2l1s2 sin(q3 + q2)] q21 +m2l1s2 sinq3q2

3 + 2m2l1s2 sin q3q2q3

+m1gs1 cosq2 +m2g(l1 cosq2 + s2 cos(q3 + q2)) ,

h3 =[m2(l1s2 cosq2 + s2

2 cos(q3 + q2))

sin(q3 + q2)

+12(C2−A2)sin2(q3 + q2)

]q2

1 +m2l1s2 sinq3q22−m2gs2 cos(q3 + q2) ,

we come to a more compact form⎛⎝M11 0 0

0 M22 M23

0 M23 M33

⎞⎠ q+

⎛⎝h1

h2

h3

⎞⎠=

⎛⎝M1

M2

M3

⎞⎠ .

2. NEWTON-EULER equationsUsing the cut principle, we apply the momentum and moment of momentumequations to every body of the robot and get:

2

∑i=1

⎧⎨⎩(

∂ KivSi

∂ q

)T (mi Ki

vSi − KiFe

i

)

+

(∂ Kiω i

∂ q

)T (KiΘSi,i Kiω i + Kiω i Ki

Θ Si,i Kiω− KiMi

)}= 0.

The rotational and translational accelerations are

Page 76: Introduction To Dynamics

1.11 Practical Considerations 67

K1vS1 = s1

⎛⎝ l − q2

1 cos2 q2 − q22

q1 cosq2 −2q1q2 sinq2−q2 − q2

1 sinq2 cosq2

⎞⎠ ,

K2vS2 =

⎛⎝ −q2l1 sinq3 − q2

1 [l1 cosq2 + s2 cos (q3 +q2)]cos (q3 +q2)q1 [l1 cosq2 + s2 cos (q3 +q2)]−2q1q2l1 sinq2

−q2 (l1 cosq3 + s2)− q3s2 − q21 [l1 cosq2 + s2 cos(q3 +q2)]sin(q3 +q2)

⎞⎠

+

⎛⎝ −q2

2 (l1 cosq3 + s2)− q23s2 −2q2q3s2

−2q1q2s2 sin(q3 +q2)−2q1q3s2 sin(q3 +q2)−q2

2l1 sinq3

⎞⎠ ,

K1ω1 =

⎛⎝−q1 sinq2 − q1q2 cosq2

q2q1 cosq2 − q1q2 sinq2

⎞⎠ ,

K2ω2 =

⎛⎝−q1 sin(q3 +q2)− q1 (q3 + q2)cos (q3 +q2)

q3 + q2q1 cos(q3 +q2)− q1 (q3 + q2)sin(q3 +q2)

⎞⎠ .

From this, we come to the JACOBIANs of rotation and translation

(∂ K1

vS1

∂ q

)T

=

⎛⎝0 s1 cosq2 0

0 0 s1

0 0 0

⎞⎠ ,

(∂ K2

vS2

∂ q

)T

=

⎛⎝ 0 l1 cosq2 + s2 (q3 + q2) 0−l1 sinq3 0 l1 cosq3 + s2

0 0 −s2

⎞⎠ .

The applied forces write

K1Fe

1 =−m1g

⎛⎝−sinq2

0cosq2

⎞⎠ ,

K2Fe

2 =−m2g

⎛⎝−sin(q3 + q2)

0cos(q3 + q2)

⎞⎠ .

Combining all this by the use of the NEWTON-EULER equations, we finally comeout with the same results as compared with LAGRANGE’s equations of the secondkind.

Page 77: Introduction To Dynamics

Chapter 2Linear Discrete Models

2.1 Linearization

In the presentation of the methods in Chapter 1, we have already assumed concreteconceptual models, that is rigid bodies with homogeneous, constant mass that arearbitrarily connected by constraints. The motion of such multibody systems withf degrees of freedom is described by ordinary and usually nonlinear second-orderdifferential equations with (minimal) form

M(q, t) q = h (q, q,Q, t) . (2.1)

Here q (t) is the vector of minimal coordinates and M an always symmetric positive-definite mass matrix. The vector h contains gyroscopic and dissipative forces, aswell as all applied forces and moments. The vector Q contains applied forces andmoments, which we would like to parametrize. In many practically relevant cases,the quantities M and h do not depend explicitly on time.

According to the previous model considerations, we assume, first, that a refer-ence motion or a stationary state for the system’s motion can be found, and, second,that small oscillations occur about this reference. Then, the vector of minimal co-ordinates q(t) can be split into a reference vector q0(t) and a perturbation vectorηq(t) [42]. Similarly, we split the vector Q:

q(t) = q0(t)+ηq(t) , (2.2)

Q(t) = Q0(t)+ηQ(t) . (2.3)

We insert these expressions into the original equation of motion (2.1) and expandthe latter as a TAYLOR series at q0(t), q0(t), q0(t) and Q0(t) with

© Springer-Verlag Berlin Heidelberg 2015 69F. Pfeiffer and T. Schindler, Introduction to Dynamics,DOI: 10.1007/978-3-662-46721-3_2

Page 78: Introduction To Dynamics

70 2 Linear Discrete Models

M((

q0 +ηq), t)= M(q0, t)+

f

∑i=1

(∂M∂qi

)0ηqi + hot , (2.4)

h((

q0 +ηq),(q0 + ηq

),(Q0 +ηQ

), t)= h(q0, q0,Q0, t)

+

(∂h∂q

)0ηq +

(∂h∂ q

)0ηq +

(∂h∂Q

)0ηQ + hot . (2.5)

The abbreviation hot stands for higher order terms. According to the rules of vectorand tensor analysis [49]

• the derivative of a scalar with respect to a vector results in a row vector,• the derivative of a vector with respect to a vector results in a second-order tensor

(JACOBIAN matrix),• the derivative of a second-order tensor (mass matrix) with respect to a vector

results in a third-order tensor.

Assuming that also ηq and ηq are small, we obtain two equations of motion, onefor the reference motion and one for the linearized perturbation:

• reference motion

M(q0, t) q0 = h(q0, q0,Q0, t) , (2.6)

• perturbation (linearized)

M(q0, t) ηq +P(q0, q0,Q0, t) ηq +R(q0, q0, q0,Q0, t)ηq = f (2.7)

with the right-hand side

f =(

∂h∂Q

)0ηQ , (2.8)

the matrix for velocity-depending forces

P(q0, q0,Q0, t) =−(∂h∂ q

)0, (2.9)

as well as the matrix for position-depending forces

R(q0, q0, q0,Q0, t) =

(∂M∂q

)0

q0 −(∂h∂q

)0. (2.10)

Page 79: Introduction To Dynamics

2.2 Classification of Linear Systems 71

2.2 Classification of Linear Systems

The above splitting into reference and perturbation usually results in a nonlinearsystem of differential equations for the reference motion (2.6), and a linear matrix-vector system for the perturbation (2.7). In most practical cases, the reference mo-tion is known or it may be determined from the nonlinear equations. Then, the per-turbation about such a reference motion will be of interest. We consider it in thefollowing. It is worthwhile, in a first step, to take a closer look at the structure ofthe homogeneous linearized equations of motion (2.7) (f = 0). The matrices P andR can always be decomposed into a symmetric and a skew-symmetric part:

D =12

(P+PT )= DT , (2.11)

G =12

(P−PT )=−GT , (2.12)

K =12

(R+RT )= KT , (2.13)

N =12

(R−RT )=−NT . (2.14)

Thus, we obtain from (2.7):

Mηq +(D+G) ηq +(K+N)ηq = f . (2.15)

The single matrices have the following properties:

M mass matrix (symmetric)The term Mηq represents the inertia forces. It follows from the kinetic energy ofthe perturbation T = 1

2 ηTq Mηq.

D damping matrix (symmetric)The term Dηq represents damping forces that are proportional to the velocity. Itfollows from RAYLEIGH’s damping power 1

2 ηTq Dηq.

G gyroscopic matrix (skew-symmetric)The expression Gηq contains gyroscopic forces. The power ηT

q Gηq = 0 van-ishes due to the skew-symmetry of G. Gyroscopic forces do not change the en-ergy balance of the system [46].

K stiffness matrix (symmetric)The expression Kηq contains the conservative perturbation forces (position-depending forces). It can be derived from the potential of the perturbationV = 1

2ηTq Kηq.

N circulatory matrix (skew-symmetric)The circulatory matrix defines nonconservative forces Nηq, which occur for ex-ample in turbines and bearings [72]. On the one hand, the power ηqNηq doesnot vanish, on the other hand ηqNηq is not symmetric. The integral

Page 80: Introduction To Dynamics

72 2 Linear Discrete Models

∫ ηq2

ηq1

ηTq Ndηq =

∫ t2

t1ηT

q Nηqdt

depends on the path of integration and therefore on the choice of the minimalcoordinates (Section 1.6).

If the damping forces (D = 0) and the nonconservative position-depending forces(N = 0) do not exist, we get a conservative system, for which the total energy isconstant and thus the energy conservation law is valid (Section 1.6). The classifica-tion of discrete linear systems concerning these physically interpretable matrices notonly has considerable advantages for the mathematical and numerical treatment, butalso gives direct qualitative insight about the system’s behavior. This is importantand helpful in particular for stability predictions [45].

For the further treatment of systems of differential equations, it is often conve-nient to transform f linear second-order differential equations to 2 f linear first-orderdifferential equations in state space form. By means of the substitution

x =

(ηqηq

), (2.16)

we obtain

x = A(t)x+b(t) (2.17)

with

A =

(0 E

−M−1R −M−1P

), b =

(0

M−1f

). (2.18)

For this formulation, standard solution methods exist, which we will study in thefollowing. The vector x is called the state vector. Accordingly, we speak of the statespace with the coordinates x, whereas the coordinates ηq represent the configurationspace (Section 1.4).

Example 2.1 (Spherical pendulum). We consider the spherical pendulum inFig. 2.1. In Example 1.12, the equations of motion, those of the cyclic coordinatesincluded, have been derived with the minimal coordinates q = (ψ ,ϑ)T :

mR2 sin2 ϑψ+ 2mR2 sinϑ cosϑϑψ = 0 ,

mR2ϑ −mR2 sinϑ cosϑψ2 +mgRsinϑ = 0 ,

sin2 ϑψ =C .

Page 81: Introduction To Dynamics

2.2 Classification of Linear Systems 73

m

R

ϑ

Ψ

Fig. 2.1. Spherical pendulum.

• First, we consider the reduction to the planar motion (ψ ≡ 0)

ϑ +gR

sinϑ = 0 .

The condition ϑ0 = 0 defines the stationary state of the system and is obtainedfrom q = q = 0. We define the stationary state as the reference motion and weare interested in the plane motion of the pendulum for small deflections

ϑ = ϑ0︸︷︷︸=0

+ηϑ .

With sinϑ =ϑ− 16ϑ

3+hot, the perturbation satisfies the linear differential equa-tion

ϑ +gRϑ = 0 .

• Another reference motion is the cone-shaped path (ϑ ≡ ϑ0 �= 0). Depending onthe choice as to which of the original differential equations is used, one obtainsthe relations

ψ2 =C2

sin4 ϑ0=

gRcosϑ0

.

With the moment equilibrium about P (Fig. 2.2)

mgRsinϑ0 −marRcosϑ0 = 0 ,

we get

gsinϑ0 − ψ2Rsinϑ0 cosϑ0 = 0

Page 82: Introduction To Dynamics

74 2 Linear Discrete Models

and the physical interpretation of the reference motion:

ψ2 =g

Rcosϑ0.

We are interested in a perturbation in the direction of

P

R

ϑ

ψ

vψ = Rsinϑ0ψ

r = Rsinϑ0

g

ar =v2ψr = ψ2Rsinϑ0

Fig. 2.2. Moment equilibrium for the spherical pendulum.

ϑ = ϑ0 +ηϑ

and derive the corresponding differential equation for ηϑ . We use ψ = Csin2 ϑ in

the second original differential equation:

ϑ −C2(

cosϑsin3 ϑ

)+

gR

sinϑ = 0 .

The differential equation is of the type

ϑ + f (ϑ) = 0 .

With

sinϑ .= sinϑ0 +ηϑ cosϑ0 ,

cosϑ .= cosϑ0 −ηϑ sinϑ0 ,

we obtain

Page 83: Introduction To Dynamics

2.2 Classification of Linear Systems 75

0.= ϑ0︸︷︷︸

=0

+ηϑ + f (ϑ0)︸ ︷︷ ︸=0

+

(∂ f∂ϑ

)0ηϑ

= ηϑ +

{− C2

sin6 ϑ[−sin4 ϑ − 3sin2 ϑ cos2 ϑ

]+( g

R

)cosϑ

}0ηϑ

= ηϑ +

(g

Rcosϑ0

)(sin2 ϑ0 + 3cos2 ϑ0 + cos2 ϑ0

)ηϑ

= ηϑ +

[(g

Rcosϑ0

)(1+ 3cos2 ϑ0

)]ηϑ .

Therefore, it is

ω =

√( gR

)( 1cosϑ0

+ 3cosϑ0

)

2√

( gR )

π2

ω

ϑ0

Fig. 2.3. Eigen angular frequency of the perturbation for the cone-shaped reference.

the eigen angular frequency of the perturbation (Fig. 2.3). We consider somespecial cases for ϑ0:

1. ϑ0 � 1 : ω ≈ 2√

gR and ψ ≈

√gR

Because of the perturbation, the originally circular orbit for ψ is deformed toan ellipse (Fig. 2.4a).

2. ϑ0 → π2 : ω = ψ → ∞

The equatorial curve is somewhat inclined (Fig. 2.4b).

Page 84: Introduction To Dynamics

76 2 Linear Discrete Models

The case ϑ0 > π2 is not possible, otherwise ψ2 = g

Rcosϑ0would become

negative.

3. 0 < ϑ0 <π2

As for ϑ0 � 1, one obtains ellipse-shaped curves, which repeat periodically(Fig. 2.4c) with 2ψ0 > ω > ψ0. If ϑ0 is not too large, one can interpret thisas an ellipse with a rotating principal axis. For two complete oscillations withperiod T , the marching angle ψ∗ satisfies

ψ∗ =∫ 2T

0ψdt − 2π ≈ ψ02T − 2π .

With

2T =

(4πω

)=

4πψ0

√1+ 3cos2 ϑ0

,

we get

ψ∗ ≈ 2π

(2√

1+ 3cos2 ϑ0

− 1

).

ϑϑ0

ηϑ

(a) ϑ0 � 1: curve of theperturbed cone-shaped path(seen from above).

(b) ϑ0 → π2 : curve of the

perturbed cone-shapedpath.

Ψ ∗

(c) 0 < ϑ0 < π2 : curve

of the perturbed cone-shaped path.

Fig. 2.4. Limiting cases for the curve of the perturbed cone-shaped path.

2.3 Solution Methods

With the selection and definition of a model as well as the mathematical descriptionof this model in the form of differential equations, we have a mechanical and math-ematical model for the system. The equations of motion are the basis for all furtherinvestigations. For the evaluation and interpretation of a dynamic system, we areinterested in the following aspects and problems:

• Motion (time response, frequency response): motion, motion patterns, frequency,damping, stability, amplitude, and phase response functions.

Page 85: Introduction To Dynamics

2.3 Solution Methods 77

• Control: If the system is controlled, questions of observability, controllability,control quality, control stability, and control optimization have to be answered.

• Perturbation: perturbation of the system, sensitivity of parameters, deterministicand stochastic perturbation.

• Optimization: Optimization of the dynamic system as a whole (process + con-troller) with respect to certain performance criteria, design strategies for the op-timization of parameters and structures regarding sensitivities or other criteria.

The necessary methods involve a wide range of mathematics, system and controltheory. They have multidisciplinary character. In the context of this paper, only themost important methods can be discussed.

2.3.1 Linear Second-Order Systems

In this section, we consider the original second-order differential equation and re-strict ourselves to the typical cases, simple eigenvalues and periodic excitation. Themore general case is discussed in Section 2.3.2.

We first restrict ourselves to conservative mechanical systems without a right-hand side:

Mηq +Kηq = 0 , ηq ∈ IR f , {M,K} ∈ IR f , f . (2.19)

The matrices M and K are assumed to be constant. For the solution vector, wechoose the approach

ηq = ηqeλ t . (2.20)

We obtain a homogeneous system of equations for the vector ηq:

(λ 2M+K

)ηq = 0 . (2.21)

The homogeneous system has a nontrivial solution if and only if the determinantvanishes

P(λ ) := det(λ 2M+K

)= 0 , (2.22)

which defines a characteristic equation for its zeros, the so-called eigenvalues {λi}i,using the characteristic polynomial P(λ ) [65]. Analogously, the eigenvalues of ahomogeneous linear second-order differential equation with damping can be deter-mined:

P(λ ) := det(λ 2M+λ (D+G)+ (K+N)

)= 0 . (2.23)

One always obtains as many complex conjugate eigenvalue pairs as degrees of free-dom. In the case of (2.19), the eigenvalues are even purely imaginary [65, 46]:

λi =± jωi . (2.24)

Page 86: Introduction To Dynamics

78 2 Linear Discrete Models

The eigenvectors{ηqi

}i

are derived from the eigenvalues up to a factor

(λ 2

i M+λi (D+G)+ (K+N))ηqi

= 0 , (2.25)

where we assume that we can find f linearly independent complex conjugate eigen-vector pairs with this procedure. Complex conjugate eigenvalues always belong tocomplex conjugate eigenvectors [65]:

λi = δi + jωi ⇒ ηqi= α i + jβ i , (2.26)

λi+ f = δi − jωi ⇒ ηqi+ f= α i − jβ i . (2.27)

The mapping λi to λi+ f is a question of definition. In case of (2.19) with δi = 0,we obtain real eigenvectors ηqi

= α i (no phase shift). For conservative vibrationsystems, these eigenvectors represent the mode shapes; in general, one obtains onlyfundamental vibrations. The solution of the homogeneous system is obtained as alinear combination of all fundamental vibrations:

ηq(t) =f

∑i=1

eδit(

ciηqie jωit + ci+ f ηqi+ f

e− jωit). (2.28)

Without damping, all eigenvectors are real and occur twice; ci+ f is complex conju-gate to ci. This gives the real representation

ηq(t) =f

∑i=1

ηqi(ai cos(ωit)+ bi sin(ωit)) (2.29)

with ai = 2ℜ(ci) and bi =−2ℑ(ci).With the help of the modal matrix

V =(ηq1

, ηq2, . . . , ηq f

)(2.30)

from (2.29), we get the matrix-vector relationship

ηq = V

⎡⎢⎢⎢⎢⎣

⎛⎜⎜⎜⎜⎝

cos(ω1t) 0 · · · 0

0. . .

. . ....

.... . .

. . . 00 · · · 0 cos

(ω f t

)

⎞⎟⎟⎟⎟⎠

⎛⎜⎝

a1...

a f

⎞⎟⎠

+

⎛⎜⎜⎜⎜⎝

sin(ω1t) 0 · · · 0

0. . .

. . ....

.... . .

. . . 00 · · · 0 sin

(ω f t

)

⎞⎟⎟⎟⎟⎠

⎛⎜⎝

b1...

b f

⎞⎟⎠

⎤⎥⎥⎥⎥⎦ . (2.31)

We write

Page 87: Introduction To Dynamics

2.3 Solution Methods 79

ηq = V [cos(Ω t)a+ sin(Ω t)b] (2.32)

with

cos(Ω t) = diag{cos(ωit)} , sin (Ω t) = diag{sin(ωit)} , (2.33)

a =(a1, · · · ,a f

)T, b =

(b1, · · · ,b f

)T. (2.34)

With the initial conditions ηq = ηq0and ηq = ηq0

at t = 0, the unknown vectors aand b can be defined:

ηq0= Va , ηq0

= VΩb , with Ω = diag{ωi} . (2.35)

Because of the invertibility of V, it follows

a = V−1ηq0, (2.36)

b =Ω−1V−1ηq0(2.37)

and therefore without damping:

ηq(t) = Vcos(Ω t)V−1ηq0+Vsin(Ω t)Ω−1V−1ηq0

. (2.38)

The initial conditions determine the components of the vectors a and b and thus theeigen angular frequencies. The modal matrix V acts as a distribution among the in-dividual degrees of freedom using the eigenvectors, which are arranged column-by-column correspondingly. The practical significance of the eigenvectors ηq1

, . . . , ηq fis often underestimated. The distribution among the individual degrees of freedomgives an idea which components oscillate and which do not and which componentsvibrate against each other, especially for forced oscillations. This is essential forlarge mechanical systems and allows conclusions for possible design improvements.

We use the solution (2.38) in the differential equation (2.19). For simplicity, letηq0

= 0. With

ηq =−VΩ 2 cos(Ω t)V−1ηq0(2.39)

and

M−1Kηq = M−1KVcos(Ω t)V−1ηq0, (2.40)

it follows

0 =−VΩ 2 cos(Ω t)V−1ηq0+M−1KVcos(Ω t)V−1ηq0

=[−VΩ 2 +M−1KV

]cos(Ω t)V−1ηq0

(2.41)

As ηq0is arbitrary, it is

Ω 2 = V−1M−1KV . (2.42)

Page 88: Introduction To Dynamics

80 2 Linear Discrete Models

Conversely, the equations of motion

ηq +M−1Kηq = 0 (2.43)

can be transformed with the modal transformation

ηq = Vξ or ξ = V−1ηq (2.44)

to

0 = Vξ +M−1KVξ = ξ +V−1M−1KVξ

= ξ +Ω 2ξ . (2.45)

The modal transformation causes a decoupling of f equations, such that a singlescalar equation, which is not coupled with other modes of vibration, results for eachoscillation frequency. We call ξ modal coordinates as opposed to the natural coor-dinates ηq and speak of fundamental vibrations. The natural coordinates are repre-sented as a linear combination of the eigenvectors using the modal transformation;the modal coordinates are the corresponding weights or fractions:

ηq =∑iηqi

ξi . (2.46)

As a solution of (2.45), it follows

ξ (t) = cos(Ω t)ξ 0 + sin(Ω t)Ω−1ξ 0 . (2.47)

Comparing this with the solution for ηq(t) gives:

ξ 0 = V−1ηq0, (2.48)

ξ 0 = V−1ηq0. (2.49)

For the derivation of the solution method, we have assumed that there are enoughlinearly independent eigenvectors in the system. This is not always correct, but itcan be shown for conservative systems (2.45).

We consider the conservative system

Mηq +Kηq = 0 (2.50)

with symmetric and positive definite mass matrix M and symmetric stiffness matrixK. Then, it should first be noted that M−1K is diagonalizable. In each case, there isa matrix V, such that V−1M−1KV can be decomposed into JORDAN blocks (Sec-tion 2.3.2.2). Without restriction, we can consider a single JORDAN block. Thencolumn-by-column,

Page 89: Introduction To Dynamics

2.3 Solution Methods 81

M−1Kηq1= ληq1

, (2.51)

M−1Kηq2= δ1ηq1

+ληq2, (2.52)

M−1Kηq3= δ2ηq2

+ληq3, (2.53)

...

With (2.51), it follows

ηTq2

Kηq1= ληT

q2Mηq1

, (2.54)

and with (2.52), it is

ηTq1

Kηq2= δ1ηT

q1Mηq1

+ληTq1

Mηq2. (2.55)

Finally, we obtain

δ1 =ηT

q1Kηq2

−ληTq1

Mηq2

ηTq1

Mηq1

= 0 . (2.56)

This procedure can be continued iteratively; so M−1K is diagonalizable and V con-tains the eigenvectors.

Because of the diagonalizability of M−1K, the equations of motion (2.50) be-come decomposed into one-mass oscillators. What about M and K? The example

M = K =

(1 11 2

)(2.57)

yields the double eigenvalue 1 for M−1K; the eigenvectors are linearly independentbut arbitrary. In particular, we can achieve that VT MV and VT KV are not diagonal.For simplicity, we assume that all eigenvalues of M−1K are different. Then, VT MVand VT KV are diagonal matrices, because apparently every eigenvalue-eigenvectorpair

(λ , ηq

)of M−1K solves the equation

(λM−K) ηq = 0 . (2.58)

Thus, for two such pairs(λn, ηqn

)and

(λm, ηqm

), it is

ηTqn

Mηqmλm = ηT

qnKηqm

, (2.59)

ηTqm

Mηqnλn = ηT

qmKηqn

. (2.60)

This yields

ηTqn

Mηqm(λm −λn) = 0 (2.61)

Page 90: Introduction To Dynamics

82 2 Linear Discrete Models

and finally our proposition. The work of the inertia and elastic forces of mode n withrespect to the displacement of mode m vanishes. With ηq = Vξ , this means

VT MV︸ ︷︷ ︸DM

ξ +VT KV︸ ︷︷ ︸DK

ξ = 0T (2.62)

and therefore

V−1M−1KV = D−1M DK . (2.63)

With damping from the complex fundamental vibrations

{ηqi

e−δit e jωit}2 f

i=1, (2.64)

we obtain real fundamental vibrations{

e−δitri ∗ e j(ωitE+ϕi)}2 f

i=1(2.65)

with componentwise multiplication ∗. Here, r is the real amplitude and ϕ is thephase shift, which are contained in the complex eigenvectors.

We consider a system with external forces f:

Mηq +(D+G) ηq +(K+N)ηq = f . (2.66)

Its solution consists of the solution of the homogeneous system (2.28) and a partic-ular solution. We restrict ourselves to periodic excitations

f(t) =∞

∑k=−∞

Fke jkΩt , Fk ∈ Cn . (2.67)

Since solutions of linear systems may be superposed, we consider, without restric-tion, even a harmonic excitation

f(t) = Fke jkΩt , (2.68)

which we intend to treat with a special approach for the particular solution:

ηqpk= ηqpk

e jkΩt , ηqpk∈ C

n . (2.69)

Insertion into the differential equation yields

ηqpk=(−k2Ω 2M+ jkΩP+R

)−1

︸ ︷︷ ︸G( jΩ)

Fk . (2.70)

The complex frequency response function G( jΩ) contains amplitude and phaseinformation. It gives the stationary transmission of the vibration system to a unit

Page 91: Introduction To Dynamics

2.3 Solution Methods 83

excitation function depending on the excitation angular frequency. The frequencyresponse function is a frequency analysis tool, but can also be used to represent theparticular solution of (2.69) in the time domain. Since in practice, the solution ofthe homogeneous system (2.28) decays because of damping, the steady-state com-ponent dominates after some time. This explains the nomenclature and the question,why usually only the stationary solution is determined if one is interested in thesystem’s response to some excitation (Chapter 5).

General solution methods for multiple eigenvalues are discussed in the nextsection.

2.3.2 Linear First-Order Systems

A similar solution procedure as in the last section can be specified, if we start fromthe state space form at the end of Section 2.2:

x = Ax , x ∈ IR2 f , A ∈ IR2 f ,2 f . (2.71)

The matrix A is assumed to be constant. With the approach

x(t) = xeλ t , (2.72)

we obtain a homogeneous linear system of equation for the eigenvectors x:

(λE−A) x = 0 . (2.73)

The characteristic equation for the eigenvalues λ is given via the characteristic poly-nomial:

P(λ ) = det(λE−A) = 0 . (2.74)

For a real matrix A, this polynomial is a real polynomial of degree 2 f . The 2 f(complex) zeros λi are called eigenvalues.

2.3.2.1 Modal Behavior for a System without Multiple Eigenvalues

The eigenvector xi corresponding to the eigenvalue λi is determined by the homo-geneous system of equations

(λiE−A) xi = 0 (2.75)

up to an arbitrary constant factor. The solution can be represented as a linear com-bination of these eigenvectors for systems without multiple eigenvalues [76]:

x(t) =2 f

∑i=1

cixieλit (2.76)

Page 92: Introduction To Dynamics

84 2 Linear Discrete Models

with the initial condition

x0 = x(0) =2 f

∑i=1

cixi . (2.77)

The eigenvectors xi from (2.76) can be summarized to a modal matrix

X =(x1 x2 . . . x2 f

), (2.78)

the eigenvalues λi to a diagonal matrix

Λ = diag{λi} (2.79)

and the constants ci to the column vector

c =(c1, · · · ,c2 f

)T. (2.80)

This yields

x(t) =2 f

∑i=1

xieλit ci =

(x1, · · · , x2 f

)⎛⎜⎜⎜⎜⎝

eλ1t 0 · · · 0

0. . .

. . ....

.... . .

. . . 00 · · · 0 eλ2 f t

⎞⎟⎟⎟⎟⎠c = XeΛ tc . (2.81)

The above definition of eΛ t is an extension of the scalar calculation rules to matri-ces:

eΛ t = E+Λ t +12(Λ t)2 + hot = diag

{∞

∑k=0

(λit)k

k!

}= diag

{eλit

}. (2.82)

Further, it holds

x0 =2 f

∑i=1

xici = Xc or c = X−1x0 . (2.83)

This yields the solution

x(t) =(

XeΛ tX−1)

x0 . (2.84)

The expression in brackets is called the fundamental matrix:

Φ(t) = XeΛ tX−1 with eΛ t = X−1Φ(t)X . (2.85)

Φ(t) and eΛ t are similar matrices.

Page 93: Introduction To Dynamics

2.3 Solution Methods 85

Decoupled, the solution is particularly easy. This can be achieved by the modaltransformation:

x(t) = Xζ (t) , (2.86)

x(0) = Xζ (0) . (2.87)

Then, it is

ζ (t) = X−1AXζ (t) .

If we now substitute the solution (2.84) in 0 = x−Ax, we obtain

0 = XΛeΛ tX−1x0 −A(

XeΛ tX−1)

x0 = [XΛ −AX]eΛ tX−1x0 (2.88)

and thus the decoupled system

ζ (t) =Λζ (t) (2.89)

with the solution

ζ = eΛ tζ 0 and ζ 0 = X−1x0 . (2.90)

This is known as the normal form of the vibration system and ζ as normal coordi-nates.

Normal coordinates ζ i change over time independently of the other normal co-ordinates ζ k. Normal coordinates ζ are complex in the general case and difficult tointerpret. They result from a linear combination of position and velocity coordinates,and thus usually have no clear and plausible interpretation. However, because of the

decoupling(ζ =Λζ

), they offer mathematically significant benefits, particularly

in the context of further considerations like the development of system models forcontroller design.

Each normal coordinate is either zero, an increasing/damped periodic oscillation,or an aperiodic motion. The properties result from the eigenvalues λi, which formcomplex conjugate pairs for matrices with real coefficients:

λi = δi ± jωi with {δ ,ω} ∈ IR . (2.91)

With

e(δ+ jω)t = eδ t (cosωt + j sinωt) ,

the following basic characteristics result (Section 5.2):

1. δi = 0According to Fig, 2.5a, one obtains a stationary continuous vibration.

2. ωi = 0Fig. 2.5b shows increasing or decaying asymptotic solutions.

Page 94: Introduction To Dynamics

86 2 Linear Discrete Models

3. δi �= 0 and ωi �= 0One gets increasing and decaying oscillations (Fig. 2.5c and Fig. 2.5d).

ζ i

t

(a) Stationary continuous vibration.

ζ i

δi = 0δi < 0

δi > 0

t

(b) Asymptotic solution.

ζ i

t

(c) Increasing oscillation.

ζ i

t

(d) Decaying oscillation.

Fig. 2.5. Basic characteristics.

Example 2.2 (Double pendulum). We consider a double pendulum with two pointmasses m and two massless pendulum rods of length l. As minimal coordinates weuse the absolute angles ηq = (ϕ1,ϕ2)

T . We assume that these are small: ϕ1 � 1 andϕ2 � 1. The acceleration due to gravity with constant g acts in negative Iy direction.First, we derive the equations of motion using the LAGRANGE equations of thesecond kind. We linearize during this process in order to save computational effort.This procedure requires understanding of the system; linearization during and aftersetting up the equations of motion are not equivalent. Then, we calculate the modeshapes. We start with kinematic considerations and describe Cartesian positions andvelocities of the two masses using the minimal coordinates:

• Cartesian positions:

x1 = l sinϕ1 ,

y1 =−l cosϕ1 ,

x2 = l sinϕ1 + l sinϕ2 ,

y2 =−l cosϕ1 − l cosϕ2 .

Page 95: Introduction To Dynamics

2.3 Solution Methods 87

Iy

Ix

ϕ2

ϕ1

m

lϕ1

m

lϕ2

l

l

Fig. 2.6. Double pendulum.

• Cartesian velocities:

x1 = lϕ1 cosϕ1 ,

y1 = lϕ1 sinϕ1 ,

x2 = lϕ1 cosϕ1 + lϕ2 cosϕ2 ,

y2 = lϕ1 sinϕ1 + lϕ2 sinϕ2 .

The square of the velocities

v21 =

(x2

1 + y21

)= (lϕ1)

2 ,

v22 = x2

2 + y22 = (lϕ1)

2 +(lϕ2)2 + 2l2ϕ1ϕ2 (cosϕ1 cosϕ2 + sinϕ1 sinϕ2)︸ ︷︷ ︸

cos(ϕ2−ϕ1).=1

is required for the calculation of the kinetic energy.

After having calculated the energies, we arrive at the evaluation of the LAGRANGE

equations of the second kind.

Page 96: Introduction To Dynamics

88 2 Linear Discrete Models

• Energies (ϕ1 � 1 and ϕ2 � 1):

T =12

m(lϕ1)2 +

12

m[(lϕ1)

2 +(lϕ2)2 + 2l2ϕ1ϕ2

],

V = mgl (1− cosϕ1)+mgl [(1− cosϕ1)+ (1− cosϕ2)]

= mgl

[ϕ2

1 +12ϕ2

2

]+ hot .

The potential energy is developed up to quadratic terms resulting in linear equa-tions of motion after the evaluation of the LAGRANGE equations of the secondkind.

• LAGRANGE equations of the second kind (1.151):With (

∂T∂ ϕ1

)= ml2 [2ϕ1 + ϕ2] ,

(∂T∂ ϕ2

)= ml2 [ϕ2 + ϕ1] ,

(∂V∂ϕ1

)= 2mglϕ1 ,

(∂V∂ϕ2

)= mglϕ2 ,

we obtain

ml2(

2 11 1

)(ϕ1

ϕ2

)+mgl

(2 00 1

)(ϕ1

ϕ2

)=

(00

).

We summarize

Mηq +Kηq = 0

with

M =

(2 11 1

), K =

(gl

)(2 00 1

)

and

M−1 =

(1 −1−1 2

).

We condense the equation of motion taking advantage of ω =√

gl

Page 97: Introduction To Dynamics

2.3 Solution Methods 89

ηq+ω2(

2 −1−2 2

)︸ ︷︷ ︸

(M−1K)

ηq = 0

and start the standard solution procedure:

• Characteristic equation with approach ηq = ηqeλ t :

0 = det

((λ 2 00 λ 2

)+ω2

(2 −1−2 2

))=(2ω2 +λ 2)2 − 2ω4 .

• Eigenvalues of the characteristic equation:

λ1,2,3,4 =± jω√

2∓√

2 .

• Eigenvectors with a linear system of equations:((

2ω2 +λ 2i

) −ω2

−2ω2(2ω2 +λ 2

i

))(

ηq1

ηq2

)= 0 .

From the first equation, we get the ratio

ηqi2

ηqi1

=

(2ω2 +λ 2

i

ω2

)=±

√2

and finally

ηq1=

(1

+√

2

), ηq2

=

(1

−√2

).

The eigenvectors are real and double. We summarize them to the modal matrix:

V =

(1 1

+√

2 −√2

).

Consequently, there are two eigen angular frequencies

ω1 = ω√

2−√

2 = 0,765

√gl,

ω2 = ω√

2+√

2 = 1,848

√gl,

which we arrange: cos(Ω t) = diag{cosωit} and sin (Ω t) = diag{sinωit}. With thisinformation, we consider again the solution:

Page 98: Introduction To Dynamics

90 2 Linear Discrete Models

• Solution with ηq0= 0:

ηq =[Vcos(Ω t)V−1]ηq0

.

We obtain:

(ϕ1 (t)ϕ2 (t)

)=

12

⎛⎝

(ϕ10 +

ϕ20√2

)cos(ω1t)+

(ϕ10 − ϕ20√

2

)cos(ω2t)

√2(ϕ10 +

ϕ20√2

)cos(ω1t)−√

2(ϕ10 − ϕ20√

2

)cos(ω2t)

⎞⎠ .

• Modal transformation:With the modal transformation

ηq = Vξ ,

we have the opportunity for an interpretation. From

ξ 0 =12

(1 1√

21 − 1√

2

)(ϕ10

ϕ20

),

it follows

ξ (t) =12

⎛⎝(ϕ10 +

ϕ20√2

)cos(ω1t)(

ϕ10 − ϕ20√2

)cos(ω2t)

⎞⎠ .

• Interpretation:As a weighting function, modal coordinates describe the fraction of the modeshapes with respect to the overall motion. If we choose the coordinates of aneigenvector as the natural initial conditions, the other modal coordinate disap-pears. The two mode shapes can be realized easily in an experiment (Fig. 2.7).

The first mode (left) has the eigen angular frequency ω1; the second mode (right)oscillates with the eigen angular frequency ω2. The corresponding eigenfrequenciesare defined as

fk =ωk

2π.

The period of oscillation is obtained from

Tk =2πωk

.

Since the ratio

ηq22

ηq21

=−√

2

Page 99: Introduction To Dynamics

2.3 Solution Methods 91

VN

Fig. 2.7. Mode shapes of the double pendulum (VN = vibration node).

of the second mode is time independent, a so-called vibration node is formed. This isa characteristic feature of the mode shapes: the first mode does not have a vibrationnode, the second mode has one vibration node. In the illustration of Fig. 2.7 theargument only applies in a linearized sense for small deflections. A mode defines asynchronous response: amplitudes stay within a constant ratio.

For general damping according to (2.65), one obtains different amplitudes andphase shifts for the fundamental solution components. The relative behavior of thecomponents is no longer time independent. Therefore, we will not find a vibrationnode in general damped systems.

Example 2.2 shows that the eigenvectors of M−1K are linearly independent but notorthogonal in general. Symmetry of M−1K, however, is sufficient for the orthogo-nality of the eigenvectors. To see this, we combine two eigenvalue equations to theexpression

0 = ηTqn

M−1Kηqm−ηT

qmM−1Kηqn

= ηTqnηqm

(λm −λn) .

For the restriction to simple eigenvalues, it follows the proposition: for multipleeigenvalues, there is freedom of choice.

If we consider the effect of the modal transformation on a right-hand side f, thenVT f defines its fraction in the equations for the modal coordinates ξ . There is nocontribution to ξ n, if and only if ηqn

⊥ f. For excitation at point k with f = ek, thereis no contribution to ξ n, if and only if

0 = ηTqn

ek = ηTqnk

,

that is, if the discrete mode ηqnhas a vibration node at k.

Page 100: Introduction To Dynamics

92 2 Linear Discrete Models

2.3.2.2 Modal Behavior for a System with Multiple Eigenvalues

The system matrix A in (2.71) is diagonalizable with a similarity transformationonly if enough linearly independent eigenvectors can be found. This can be guar-anteed with pairwise distinct eigenvalues. If there are multiple eigenvalues λi as thezeros of the characteristic polynomial with the algebraic multiplicity vi , there existsa number di (geometric multiplicity) of linearly independent eigenvectors:

di = 2 f − rg(λiE−A) (2.92)

where rg denotes the rank of the matrix. It holds [65, 46]

1 ≤ di ≤ vi with i = 1, · · · ,m . (2.93)

The matrix A is not diagonalizable, but a fundamental matrix Φ(t) can be derived.First, there always exists a regular matrix X ∈ IR2 f ,2 f , such that

X−1AX =

⎛⎜⎜⎜⎝

J1

J2 0

0. . .

Jm

⎞⎟⎟⎟⎠=: J (2.94)

decomposes into so-called JORDAN blocks

Ji :=

⎛⎜⎜⎜⎜⎝

λi 1 0. . .

. . .

. . . 10 λi

⎞⎟⎟⎟⎟⎠ . (2.95)

These differ from a diagonal matrix by the upper secondary diagonal which is filledwith ones. The number of JORDAN blocks with the same eigenvalue λi is equal tothe number of corresponding linearly independent eigenvectors. For each eigenvalueλi, there exist exactly di JORDAN blocks with the same λi on the diagonal accordingto (2.92).

With the transformation ζ = X−1x, we obtain

ζ = Jζ . (2.96)

To derive the solution, we begin by considering the first JORDAN block of length l:

⎛⎜⎝ζ 1...ζ l

⎞⎟⎠=

⎛⎜⎜⎜⎜⎝

λ1 1 0. . .

. . .

. . . 10 λ1

⎞⎟⎟⎟⎟⎠

⎛⎜⎝ζ 1...ζ l

⎞⎟⎠ . (2.97)

Page 101: Introduction To Dynamics

2.3 Solution Methods 93

The solution is obtained as a backward recursion of integration steps starting withthe last equation. The initial step satisfies

ζ l = ζ l,0eλ1t . (2.98)

As

ζ l−1 = λ1ζ l−1 + ζ l ⇔d(ζ l−1e−λ1t

)dt

= ζ le−λ1t , (2.99)

the recursion step reads

ζ l−1 =(ζ l,0t + ζ l−1,0

)eλ1t . (2.100)

Finally, we obtain

⎛⎜⎝ζ 1...ζ l

⎞⎟⎠= K1(t)

⎛⎜⎝

ζ 1,0...

ζ l,0

⎞⎟⎠ (2.101)

with

K1(t) :=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎝

1 t t2

2tl−1

(l−1)!. . .

. . . t2

2. . . t

0 1

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎠

eλ1t . (2.102)

We repeat this with the other JORDAN blocks:

ζ (t) =

⎛⎜⎜⎜⎜⎜⎝

K1(t)0

. . .0

Km(t)

⎞⎟⎟⎟⎟⎟⎠

︸ ︷︷ ︸=:K

ζ 0 . (2.103)

In natural coordinates, we get

x(t) =(XK(t)X−1)︸ ︷︷ ︸

=:Φ(t)

x0 . (2.104)

A method for calculating the columns of the matrix X =(x1, · · · ,x2 f

)is obtained

directly from (2.94):

Page 102: Introduction To Dynamics

94 2 Linear Discrete Models

AX = XJ . (2.105)

If we start again with the first JORDAN block J1 of length l, it follows

Ax1 = λ1x1 ,

Ax2 = λ1x2 + x1 ,

... =...

Axl = λ1xl + xl−1 .

This iteration can be solved by forward recursion of systems of linear equations.The vector x1 is called eigenvector corresponding to the eigenvalue λ1; the vectorsx2, · · · ,xl are called generalized eigenvectors. We proceed similarly with all otherJORDAN blocks.

Example 2.3 (Wagon with a pendulum). A wagon (mass m1) with a pendulum(mass m2, length l) can move freely on horizontal rails (coordinate s). The pendulumcan rotate without friction (angle ϕ). The acceleration of gravity acts in the negative

Iy direction (Fig. 2.8).

m2

m1

s

ϕl

Iy

Ix

Fig. 2.8. Wagon with a pendulum.

We describe the motion with minimal coordinates q = (s,ϕ)T . The equations ofmotion are derived from energy expressions

T =12

m1s2 +12

m2(s2 + l2ϕ2 + 2l cosϕ sϕ

),

V =−m2gl cosϕ

with the LAGRANGE equations of the second kind:

Page 103: Introduction To Dynamics

2.3 Solution Methods 95

(m1 +m2 m2l cosϕm2l cosϕ m2l2

)(sϕ

)+

(−m2l sinϕϕ2

m2gl sinϕ

)=

(00

).

We are interested in small displacements and small velocities around the equilibriumposition q0 = 0. Then, we have

(m1 +m2 m2l

m2l m2l2

)(sϕ

)+

(0 00 m2gl

)(sϕ

)=

(00

)

with

(s,ϕ) = (s0,ϕ0)+(ηs,ηϕ

).

For m1 = m2 =: m, the matrix A writes

A =

⎛⎜⎜⎝

0 0 1 00 0 0 10 −lω 0 00 2ω 0 0

⎞⎟⎟⎠

with ω2 = gl . This yields the following eigenvalues:

λ1,2 =± j√

2ω ,

λ3,4 = 0 .

The eigenvalue λ3 = 0 has algebraic multiplicity v3 = 2. We have to consider itsgeometric multiplicity separately:

d3 = 2 f − rg(λ3E−A) = 1 .

The matrix A cannot be diagonalized. Thus, for the corresponding JORDAN normalform, it follows

J =

⎛⎝J1

J2

J3

⎞⎠

with

J1 = λ1 ∈ IR1,1 ,

J2 = λ2 ∈ IR1,1 ,

J3 =

(0 10 0

)∈ IR2,2 .

The matrix X, which transforms A to the JORDAN normal form according to (2.94),can be calculated with (2.105):

Page 104: Introduction To Dynamics

96 2 Linear Discrete Models

X =

⎛⎜⎜⎝

l l 1 1−2 −2 0 0

j√

2ω l − j√

2ω l 0 1− j2

√2ω j2

√2ω 0 0

⎞⎟⎟⎠ .

The first three columns x1, x2, and x3 of X are the eigenvectors corresponding to theeigenvalues λ1, λ2, and λ3. The fourth column is the generalized eigenvector for thedouble eigenvalue λ3 = λ4 = 0 (rigid body motion).

2.3.2.3 Forced Oscillations

We consider an externally excited system in the state space form of the equations(2.16) to (2.18):

x(t) = Ax(t)+b(t) . (2.106)

The matrix A is assumed to be constant. The solution of (2.106) consists of a homo-geneous part (2.84) for x=Ax and a particular part for the external excitation [46]:

x(t) =Φ(t)x0 + xp(t) . (2.107)

The particular solution xp can be found by variation of constants:

xp(t) =Φ(t)c(t) . (2.108)

From (2.106), we obtain the relation

Φ(t)c(t)+Φ(t)c(t) = xp(t) = Axp(t)+b(t) = AΦ(t)c(t)+b(t) . (2.109)

With the fundamental solution Φ(t) = AΦ(t), it follows:

Φ(t)c(t) = b(t) . (2.110)

Due to (2.85), it holds

c(t) =Φ(−t)b(t) , (2.111)

which we integrate:

c(t) = c0 +

∫ t

0Φ (−τ)b (τ)dτ . (2.112)

Substituting the values c0 = 0 for t = 0 and noting, that Φ(t)Φ (−τ) = Φ (t − τ)according to (2.85), yield the particular solution

Page 105: Introduction To Dynamics

2.3 Solution Methods 97

xp =

∫ t

0Φ (t − τ)b(τ)dτ

︸ ︷︷ ︸Duhamel integral

(2.113)

and the overall solution

x(t) =Φ(t)x0 +∫ t

0Φ (t − τ)b(τ)dτ . (2.114)

Example 2.4 (Periodic excitation). Many technically important excitations in ma-chine dynamics can be represented as periodic oscillations. As a representative, weconsider the right-hand side

b(t) = b0e jωt .

More complicated periodic excitation can be represented by FOURIER series; thesolution is given by the superposition principle [41]. From (2.114), we get:

x(t) =Φ(t)x0 +

∫ t

0Φ (t − τ)b0eiωτdτ .

Since according to (2.85), the fundamental matrix satisfies Φ(t) = XeΛ tX−1, onehas to evaluate some simple integrals with exponential functions in the equationabove.

A common and relatively simple way to describe forced oscillations is via theLAPLACE transformation [76]. The function

F(s) =L { f}(s) :=∫ ∞

0e−st f (t)dt (2.115)

is called LAPLACE transform of f (t). The value s is a complex variable. Applying(2.115) to (2.106), we obtain

x(s) = (sE−A)−1 b(s) . (2.116)

These two rules are basically used:

L {a1 f1 + a2 f2}(s) = a1L { f1}+ a2L { f2}(s) (linearity) , (2.117)

L{

f (n)}(s) = snL { f}(s)−

n−1

∑k=0

f (k)(0)sn−k−1 (differentiation pre-image) .

(2.118)

Since we are only interested in the particular solution, the initial value is omittedin the differentiation rule. Equation (2.116) assumes the existence of the inverse(sE−A)−1. Since formally the inverse of a matrix is formed from its adjoint andits determinant [65], the determinant det(sE−A) appears in the denominator of

Page 106: Introduction To Dynamics

98 2 Linear Discrete Models

(2.116). If the excitation frequencies contained in b(s) equal the eigenvalues of A,it holds det(sE−A) = 0 and we obtain resonances. The absolute value of x(s) de-fines the amplitude frequency response function; the phase of x(s) corresponds tothe phase frequency response function. The evaluation of (2.116) is often easier toperform in individual cases than that of (2.114), because one obtains directly theamplitude and phase frequency response functions which are typically of interest(Section 2.3.1).

2.4 Stability of Linear Systems

We have already seen that the eigenvalues of the system play a central role for thetime behavior. For the question of stability, we notice the following:

Theorems of LYAPUNOV [36, 45]The equilibrium position x(t)≡ 0 of the linear time-invariant system

x(t) = Ax(t) , x(t0) = x0 (2.119)

is

• asymptotically stable, if and only if all eigenvalues λi of A have negativereal parts, ℜ(λi)< 0 (decaying oscillations),

• stable, if and only if A has no eigenvalues λi with positive real parts andfor all eigenvalues with ℜ(λk) = 0, it holds: Rg(λkE−A) = 2 f − vk,

• unstable, if and only if at least one eigenvalue λi of A has a positive realpart, ℜ(λi) > 0, or a multiple eigenvalue with vanishing real part occurs,the algebraic multiplicity of which is larger than its geometric multiplicity.

2.4.1 Criteria Based on the Characteristic Polynomial

The eigenvalues λi of A follow from the characteristic equation

P(λ ) = det(λE−A) = a0λ 2 f + a1λ 2 f−1 + · · ·+ a2 f−1λ + a2 f = 0 . (2.120)

To assess the stability in terms of the eigenvalues, the solution of the characteristicequation is necessary. Of much greater interest is the question whether one candraw conclusions about the stability of the equilibrium position x ≡ 0 only by thestructure of the characteristic polynomial. We give some criteria on the basis of thecoefficients of the characteristic polynomial. Further details can be found in [38, 45,10].

Page 107: Introduction To Dynamics

2.4 Stability of Linear Systems 99

2.4.1.1 Stodola Criterion

Necessary for negative real parts of the eigenvalues λi is the condition

ak > 0 , k = 0,1, · · ·2 f . (2.121)

Since the characteristic equation is determined only up to a factor, especially −1,the condition a0 > 0 must be assumed without restriction.

With this criterion, we can exclude asymptotic stability. To prove asymptoticstability, the characteristic equation has to be further investigated.

2.4.1.2 Routh-Hurwitz Criterion

Necessary and sufficient for negative real parts of the eigenvalues λi is thecondition

Hk > 0 , k = 1, · · · ,2 f . (2.122)

The HURWITZ determinants Hk are leading principle minors of the HURWITZ ma-trix H, which is formed from the coefficients of the characteristic polynomial:

H =

⎛⎜⎜⎜⎜⎜⎜⎝

a1 a3 a5 a7 · · · 0a0 a2 a4 a6 · · · 00 a1 a3 a5 · · · 00 a0 a2 a4 · · · 0· · · · · · · · · · · · · · · 00 · · · · · · · · · · · · an

⎞⎟⎟⎟⎟⎟⎟⎠

. (2.123)

The matrix H ∈ IR2 f ,2 f is constructed as follows:

1. The first row is obtained by inserting the coefficients of the characteristic poly-nomial in the corresponding columns, where the index is increased by 2 from onecolumn to the next.

2. In every further row, one makes use of the coefficients of the characteristic poly-nomial, whose index is lowered by 1 in comparison with the corresponding col-umn of the preceding row. The row is completed according to the first rule.

3. As one leaves the domain of definition for the coefficients with this procedure,we have to set ak = 0 for k > 2 f and k < 0.

Page 108: Introduction To Dynamics

100 2 Linear Discrete Models

Since the characteristic equation is again determined only up to a factor, especially−1, we have to assume a0 > 0 without restriction.

Example 2.5 (ROUTH-HURWITZ criterion for 2 f = 4). We assume a0 > 0 andtake a look at

H =

⎛⎜⎜⎝

a1 a3 0 0a0 a2 a4 00 a1 a3 00 a0 a2 a4

⎞⎟⎟⎠ .

Then, we get the following leading principal minors:

H1 = a1 ,

H2 = a1a2 − a0a3 ,

H3 = a1a2a3 − a21a4 − a0a2

3 ,

H4 = a4H3 .

We draw the following conclusions from the ROUTH-HURWITZ criterion:

H4 > 0 ⇒ a4 > 0 ,

H3 =−a21a4 + a3H2 ⇒ a3 > 0 ,

H2 > 0 ⇒ a2 > 0 ,

H1 > 0 ⇒ a1 > 0 .

In particular, the STODOLA criterion follows:

a0,a1,a2,a3,a4 > 0 .

If we assume the STODOLA criterion to be satisfied, then it is sufficient to ana-lyze only every second HURWITZ determinant, that is either H1,H3,H5 . . . > 0 orH2,H4,H6 . . . > 0 (theorem of CREMER, see also LIÉNARD-CHIPART criterion).

2.4.1.3 Liénard-Chipart Criterion

The ROUTH-HURWITZ criterion includes the STODOLA criterion and can be com-bined with it as follows.

Necessary and sufficient for negative real parts of the eigenvalues λi is thecondition

a2 f > 0, H2 f−1 > 0, a2 f−2 > 0, H2 f−3 > 0, · · · .

Page 109: Introduction To Dynamics

2.4 Stability of Linear Systems 101

2.4.2 Stability of Mechanical Systems

For the stability analysis, we started from the theorems of LYAPUNOV (eigenvalueanalysis) in a first step. For this, the explicit calculation of the eigenvalues is nec-essary. Consequently, we ask ourselves whether we can make statements about thestability without calculating the eigenvalues. The consideration of the characteris-tic equation yields the necessary coefficient criterion (STODOLA criterion). A fur-ther investigation gives the necessary and sufficient conditions of HURWITZ (“alge-braic criterion”). Both criteria can be combined to the LIÉNARD-CHIPART criterion.For mechanical systems, all these criteria, however, do not make use of the specialstructure of the system matrix A, that is from the physical meaning of the matricesM,D,G,K,N. We discuss this briefly below [45].

2.4.2.1 Nongyroscopic Conservative Systems

The mechanical system

Mηq +Kηq = 0 (2.124)

is critically stable, that is stable but not asymptotically stable, if and only ifthe stiffness matrix K is positive definite:

K = KT > 0 . (2.125)

For K < 0 the system is unstable.

The scalar analogue for the one degree of freedom oscillator

mηq + kηq = 0 with λ1,2 =± j

√km

(2.126)

is defined by the relations

k > 0 → critically stable , k < 0 → unstable . (2.127)

A matrix K = KT is called positive definite if the associated quadratic form satisfiesxT Kx > 0 for x �= 0. This is the case if and only if all leading principal minors arelarger than zero or all eigenvalues are positive [65].

Since the potential energy of a conservative system can be written in the form

V =12ηT

q Kηq , (2.128)

the theorems of DIRICHLET and LAGRANGE follow directly from the matrix con-dition above.

Page 110: Introduction To Dynamics

102 2 Linear Discrete Models

Theorem of LAGRANGE (Mechanique Analytique, 1788):If the potential energy V is a positive definite quadratic function in the neigh-borhood of an equilibrium position ηq = 0, then all eigenvalues λ 2

i are nega-tive real.

According to (2.128), it is V (0) = 0 for ηq = 0. If V > 0 for ηq �= 0, then K ispositive definite. This means that V (0) is an absolute minimum of V (ηq).

Theorem of DIRICHLET (1846):If V has an absolute minimum in the equilibrium position ηq = 0, then it isstable.

If λ 2 < 0, then λ =± jω is imaginary (ω real). As

eλ t = e± jωt = cos(ωt)± j sin(ωt) , (2.129)

only critical stability can be concluded, so undamped vibrations about the equilib-rium position occur.

2.4.2.2 Gyroscopic Conservative Systems

The mechanical system

Mηq +Gηq +Kηq = 0 (2.130)

is critically stable for K > 0. For K < 0, the system is stable if [40]

det(G) �= 0 is sufficiently large . (2.131)

2.4.2.3 Damped Systems

The system

Mηq +(D+G) ηq +Kηq = 0 (2.132)

is asymptotically stable if D = DT > 0 and the stiffness matrix is positivedefinite, regardless of G.

Page 111: Introduction To Dynamics

Chapter 3Linear Continuous Models

3.1 Models of Continuous Oscillators

Discrete systems are composed of rigid bodies, the essential property of which isthat the distance between two points in the interior of such bodies remains constantwith time. Elastic bodies are continua, which can deform elastically. We assume thattheir masses are homogeneous and isotropic. Furthermore, we restrict ourselves tolinear-elastic bodies and thus to small deformations. The vibrations of these bod-ies are determined by their mass and stiffness distributions, similar to masses andsprings in the discrete case. Each vibrating elastic system is characterized by eigen-frequencies and mode shapes. For each eigenfrequency, there is a correspondingmode shape which the structure takes when it oscillates at this frequency. There areinfinite many eigenfrequencies and modes, usually in a systematic order. This prop-erty makes linear-elastic vibration systems relatively easy to understand, but it doesnot apply to all continuous systems, such as rotating fluids.

The calculation of eigenfrequencies and mode shapes is based on the equations ofelastodynamics and can also be done analytically in simpler cases depending on theconfiguration of the analyzed component. In more complicated cases, we apply ap-proximation methods. Continuum systems can be decomposed or discretized math-ematically or physically. The mathematical discretization methods are based on theequations of motion in the form of partial differential equations, which are solvedwith sophisticated tools of numerical analysis (numerical integration, finite elementmethod [76, 9, 64]). In a physical discretization, one decomposes the continuum intodiscrete elements, for example also into finite elements [74, 37, 78, 4], into elementchains for the transfer-matrix method, and into a multibody system [11, 51, 62].The discretizations often lead to similar systems of equations using different inter-pretations. Many of these discretization methods have the disadvantage that theircomplexity is very large and that the numerical results often lose physical trans-parency.

In practice, elastic vibration systems are frequently incorporated into discretesystems. Then, we have a vibration system with rigid and elastic components,which we call an elastic multibody system [11]. Since the motion should always be

© Springer-Verlag Berlin Heidelberg 2015 103F. Pfeiffer and T. Schindler, Introduction to Dynamics,DOI: 10.1007/978-3-662-46721-3_3

Page 112: Introduction To Dynamics

104 3 Linear Continuous Models

described with a minimum effort and thus a minimum number of degrees of free-dom, such a system is modelled sufficiently accurately with rigid body degrees offreedom for the rigid components and with elastic degrees of freedom for the elasticcomponents; the crucial factor is the selection of the elastic degrees of freedom.

As we have already seen, the possible modes of vibration of an elastic body aregiven by its eigenmodes. Any mode of vibration resulting from excitation can begenerated by superposition of eigenmodes. The contribution of each eigenmode tothe total elastic deformation depends on the motion of the overall system, in whichthe elastic part is integrated. Therefore, we introduce eigenmodes together with amultiplicative time-dependent modal coordinate as elastic degrees of freedom.

The question of the necessary number of such degrees of freedom can be an-swered only approximately. It depends essentially on the effect of structural damp-ing on the higher eigenfrequency amplitudes, on the type of excitation, and onthe other frequencies expected in the system (related to other components or con-trollers). For example, operating frequencies and speeds are both limitations. In thecase of elastic eigenfrequencies in such an excitation domain, they have to be con-sidered. The modeling described above allows in any case the reduction of the de-grees of freedom to the necessary minimum. How to obtain the mode shapes of theindividual elastic components is a secondary question. Depending on the compo-nent, they can be calculated analytically or numerically.

3.2 Simple Examples of Continuous Vibrations

In the following, we discuss some simple analytical examples.

3.2.1 Beam as a Bending Vibrator

We consider a beam as a bending vibrator as shown in Fig. 3.1 [1, 69]. The deflectionof the beam at point x and time t is denoted by w(x, t), and the inclination of the

L

ϕ

w

z

x

Q(x, t)

Q(x, t)+( ∂Q∂x )dx

w(x, t)

dxdx

Fig. 3.1. Bending beam.

Page 113: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 105

bending line by ϕ . Q is the transverse force, M the bending moment, EI the bendingstiffness of the beam, and ρA its mass density [41]. The following relations hold:

• Kinematics:

ϕ ≈(∂w∂x

). (3.1)

• Elastostatics:

M(x) =−EI(x)

(∂ 2w∂x2

), (3.2)

Q(x) =

(∂M∂x

),

∂Q∂x

=− ∂ 2

∂x2

[EI(x)

∂ 2w∂x2

]. (3.3)

• Momentum equation in z direction:

ρA(x)dx

(∂ 2w∂ t2

)=−Q(x, t)+

[Q(x, t)+

(∂Q∂x

)dx

]=

(∂Q∂x

)dx . (3.4)

This results in the equation of motion:

∂ 2

∂x2

[EI(x)

(∂ 2w∂x2

)]+ρA(x)

(∂ 2w∂ t2

)= 0 . (3.5)

In the case of a constant cross section and a constant bending stiffness, this yieldsthe simplified equation

(∂ 4w∂x4

)+

(ρAEI

)(∂ 2w∂ t2

)= 0 . (3.6)

We introduce separation of variables according to BERNOULLI

w(x, t) =∞

∑i=1

wi(x)qi(t) = q(t)T w(x) , (3.7)

and then obtain for each individual summand

w(4)i qi +

(ρAEI

)wiqi = 0 (3.8)

and finally

(EIρA

)w(4)

i

wi=− qi

qi= ω2

i . (3.9)

Page 114: Introduction To Dynamics

106 3 Linear Continuous Models

The equality of the two ratios for all times and positions can only be achieved if theyboth take the same constant value ω2

i . Therefore, we can split the partial differentialequation into two ordinary differential equations

qi +ω2i qi = 0 , (3.10)

w(4)i − k4

i wi = 0 (3.11)

with the fundamental solutions

qi(t) = ai cos(ωit)+ bi sin(ωit) , (3.12)

wi(x) = Ai cos(kix)+Bi sin(kix)+Ci cosh(kix)+Di sinh(kix) (3.13)

and

k4i =

(ρAEI

)ω2

i . (3.14)

The four constants Ai, Bi, Ci, and Di are determined from the boundary conditions.There are various possibilities, but we focus on the following cases:

• Clamped at both ends:

w(0, t) = w(L, t) = w′(0, t) = w′(L, t) = 0 . (3.15)

x = 0 x = L

• Clamped at one end, simply supported at the other end:

w(0, t) = w′(0, t) = w(L, t) = 0 , (3.16)

M(L, t) =−EIw′′(L, t) = 0 . (3.17)

• Simply supported at both ends:

w(0, t) = w(L, t) = 0 , (3.18)

M(0, t) =−EIw′′(0, t) = 0 , (3.19)

M(L, t) =−EIw′′(L, t) = 0 . (3.20)

Page 115: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 107

• Clamped at one end:

w(0, t) = w′(0, t) = 0 , (3.21)

M(L, t) =−EIw′′(L, t) = 0 , (3.22)

Q(L, t) =−EIw′′′(L, t) = 0 . (3.23)

• Free at both ends:

M(0, t) =−EIw′′(0, t) = 0 , (3.24)

Q(0, t) =−EIw′′′(0, t) = 0 , (3.25)

M(L, t) =−EIw′′(L, t) = 0 , (3.26)

Q(L, t) =−EIw′′′(L, t) = 0 . (3.27)

For the four unknown constants Ai, Bi, Ci, and Di, we get four determining equa-tions. In the following we investigate the cantilevered case in more detail as anexample. The geometric boundary conditions wi(0) = w′

i(0) = 0 yield:

Ai +Ci = 0 and Bi +Di = 0 . (3.28)

We obtain as a solution:

wi(x) = Ai [cos(kix)− cosh(kix)]+Bi [sin(kix)− sinh(kix)] . (3.29)

If we embed the kinetic boundary conditions w′′i (L) = w′′′

i (L) = 0 into this solution,we get

0 =

(cos(kiL)+ cosh(kiL) sin(kiL)+ sinh(kiL)−sin(kiL)+ sinh(kiL) cos(kiL)+ cosh(kiL)

)(Ai

Bi

). (3.30)

This homogeneous system of equations for Ai and Bi only yields a nontrivial solutionfor a vanishing determinant. For (kiL), the eigenvalue equation reads

cos(kiL)cosh(kiL)+ 1 = 0 . (3.31)

There are infinitely many eigenvalues (Fig. 3.2), which can be approximated forlarge values (kiL) as follows:

(kiL) = (2i+ 1)π2. (3.32)

The corresponding eigen angular frequencies ωi follow from (3.14). The corre-sponding mode shapes are sketched in Fig. 3.3. They are obtained by substituting

Page 116: Introduction To Dynamics

108 3 Linear Continuous Models

cos(kL)

(kL)

− 1cosh(kL)

1,875

4,694 7,855

π2 π 3π

25π22π

Fig. 3.2. Solution of the eigenvalue equation (3.31).

x

L

ρ = const EI = const

x

w(x)

i = 1w1 = 1,8752

√EI

ρAL4

x

w(x)

i = 2w2 = 4,6942

√EI

ρAL4

x

w(x)

i = 3w3 = 7,8552

√EI

ρAL4

L

L

L

Fig. 3.3. Mode shapes of the bending beam.

Page 117: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 109

(kiL) in (3.30) and calculation of the coefficients Ai and Bi, which are finally insertedinto (3.29):

wi(x) = [cos(kiL)+ cosh(kiL)] [cos(kix)− cosh(kix)]

+ [sin (kiL)− sinh(kiL)] [sin(kix)− sinh(kix)] . (3.33)

The modes are orthogonal in the following sense (Section 3.3.1):

∫ L

0wi(x)wn(x)dx = 0 for i �= n , (3.34)

∫ L

0wi(x)wn(x)dx �= 0 for i = n . (3.35)

The solution for the beam vibration problem is obtained as the sum of all eigen-modes multiplied by the time-dependent coefficients qi(t):

w(x, t) =∞

∑i=1

qi(t)wi(x) = q(t)T w(x) . (3.36)

The coefficients ai and bi result from the initial conditions at time t = 0:

w0(x) =: w(x, t = 0) =∞

∑i=1

aiwi(x) , (3.37)

w0(x) =: w(x, t = 0) =∞

∑i=1

biωiwi(x) . (3.38)

Because of the orthogonality of the eigenmodes, multiplication by wi(x) and inte-grating over the beam length give

ai =

∫ L0 wi(x)w0(x)dx∫ L

0 w2i (x)dx

, (3.39)

bi =

∫ L0 wi(x)w0(x)dx

ωi∫ L

0 w2i (x)dx

. (3.40)

With (3.36), we see that we can incorporate an elastic component at any time asan element of a larger configuration without any change in the solution structure ofthe initial boundary value problem. In such a case, the values qi(t) are determinedfrom the motion of the overall system. They weight the modes such that forceddeformations occur.

3.2.2 Beam as a Bending Vibrator with an End Mass

The bending beam with an end mass (Fig. 3.4) is a technically important case, forexample, for the modeling of wind turbines. Its description follows almost exactly

Page 118: Introduction To Dynamics

110 3 Linear Continuous Models

L

z

mD

w(x, t)x

EI,ρA

Fig. 3.4. Bending beam with an end mass.

the one without an end mass, since the end mass is noticeable only in the boundaryconditions. We obtain the following relations:

• Equation of motion (3.5):

∂ 2

∂x2

[EI(x)

∂ 2w∂x2

]=−ρA(x)

∂ 2w∂ t2 . (3.41)

• Boundary conditions:

w(0, t) = w′(0, t) = 0 , (3.42)

w′′(L, t) = 0 , (3.43)

∂∂x

[EI(x)

(∂ 2w∂x2

)]L= mD

(∂ 2w∂ t2

)L. (3.44)

The inertia of the end mass must be compensated by the shear force.

We assume constant values EI and ρA along the beam. With separation of variablesaccording to BERNOULLI

w(x, t) =∞

∑i=1

wi(x)qi(t) = q(t)T w(x) , (3.45)

it follows

qi +ω2i qi = 0 , (3.46)

w(4)i − k4

i wi = 0 (3.47)

with

k4i =

(ρAEI

)ω2

i . (3.48)

For the solution, we choose the same approach as in (3.13):

Page 119: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 111

qi(t) = ai cos(ωit)+ bi sin(ωit) , (3.49)

wi(x) = Ai cos(kix)+Bi sin(kix)+Ci cosh(kix)+Di sinh(kix) . (3.50)

For the evaluation of the boundary conditions

wi(0) = w′i(0) = 0 , (3.51)

w′′i (L) = 0 , (3.52)

EIw′′′i (L) =−ω2mDwi(L) , (3.53)

we have to calculate the spatial derivatives:

w′i(x) = ki {−Ai sin(kix)+Bi cos(kix)+Ci sinh(kix)+Di cosh(kix)} , (3.54)

w′′i (x) = k2

i {−Ai cos(kix)−Bi sin (kix)+Ci cosh(kix)+Di sinh(kix)} , (3.55)

w′′′i (x) = k3

i {Ai sin(kix)−Bi cos(kix)+Ci sinh(kix)+Di cosh(kix)} . (3.56)

From the boundary conditions, we obtain

wi(0) = 0 ⇒ Ai +Ci = 0 , (3.57)

w′i(0) = 0 ⇒ Bi +Di = 0 , (3.58)

w′′i (L) = 0 ⇒ Ai [−cos(kiL)− cosh(kiL)]+Bi [−sin(kiL)− sinh(kiL)] = 0 ,

(3.59)

thus

Ai =− [sin(kiL)+ sinh(kiL)] , (3.60)

Bi = [cos(kiL)+ cosh(kiL)] . (3.61)

The mode shapes can be represented as follows:

wi(x) = [cos(kiL)+ cosh(kiL)] [sin (kix)− sinh(kix)]

− [sin(kiL)+ sinh(kiL)] [cos(kix)− cosh(kix)] . (3.62)

The eigenvalues (kiL) follow from the fourth boundary condition (3.53),

k3i

{− [cos(kiL)+ cosh(kiL)]

2 − [sin(kiL)+ sinh(kiL)] [sin(kiL)− sinh(kiL)]}

=−(ω2mD

EI

){[cos(kiL)+ cosh(kiL)] [sin(kiL)− sinh(kiL)]

− [sin(kiL)+ sinh(kiL)] [cos(kiL)− cosh(kiL)]} . (3.63)

With ω2i

EI =k4

iρA , we get the eigenvalue equation:

Page 120: Introduction To Dynamics

112 3 Linear Continuous Models

0 = 1+ cos(kiL)cosh(kiL)

− (kiL)

(mD

ρAL

)[sin (kiL)cosh(kiL)− cos(kiL) sinh(kiL)] . (3.64)

Given mD, the eigenvalues ki have to be determined from this relation numerically.For mD = 0, we get the same result as the one without an end mass. As in the lastsection, the initial conditions complete the overall solution:

w(x, t) =∞

∑i=1

wi(x)qi(t) = q(t)T w(x) . (3.65)

If we consider in addition the rotational inertia Θ of the end mass, the boundarycondition (3.43) has to be replaced by the moment of momentum equation

Θ w′(L, t) = M(L, t) =−EIw′′(L, t) . (3.66)

The general procedure for the determination of the coefficients stays the same.

3.2.3 Beam as a Torsional Vibrator with an End Mass

In machinery configurations, beams play an important role as torsional elements.We consider such a case as shown in Fig. 3.5 [1, 43, 69]. The torsion angle dependson the location x and the time t, ϕ = ϕ(x, t), GIp(x) is the torsional stiffness withthe polar area moment of inertia Ip, Jp = ρIp is the moment of inertia per unit lengthwith the density ρ , mD is an end mass, andΘ is the corresponding moment of inertia.We have the following equations:

• Elastostatics:

Mt(x, t) = GIp(x)

(∂ϕ∂x

). (3.67)

L

mD,Θ

x

z

dx

GIp

Mt(x, t) Mt(x, t)+( ∂Mt∂x )dx

ϕ(x, t) ϕ(x, t)+( ∂ϕ∂x )dx

dx

Fig. 3.5. Torsion element.

Page 121: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 113

• Moment of momentum equation:

Jp(x)dx

(∂ 2ϕ∂ t2

)=

[Mt(x, t)+

(∂Mt

∂x

)dx

]−Mt(x, t) =

(∂Mt

∂x

)dx . (3.68)

This results in the equation of motion:

∂∂x

[GIp(x)

(∂ϕ∂x

)]− Jp(x)

(∂ 2ϕ∂ t2

)= 0 . (3.69)

We assume constant values Ip, Jp along the torsion element and apply the separationof variables method according to BERNOULLI

ϕ(x, t) = q(t)Tϕ(x) =∞

∑i=1

ϕi(x)qi(t) . (3.70)

Two ordinary differential equations for ϕi(x) and qi(t) result

qi +ω2i qi = 0 , (3.71)

ϕ ′′i + k2

i ϕi = 0 (3.72)

with

k2i =

Jp

GIpω2

i . (3.73)

For ϕi(x), the fundamental solution is

ϕi(x) = Ai cos(kix)+Bi sin(kix) . (3.74)

The coefficients Ai and Bi are obtained from the boundary conditions:

ϕ(0, t) = 0 , (3.75)

GIp

(∂ϕ∂x

)L=−Θ

(∂ 2ϕ∂ t2

)L. (3.76)

At the left end of the beam, no deformation is present, at the right end of thebeam, the elastic torque is in balance with the inertia of the sheave. From ϕ(0, t)= 0,it follows Ai = 0 and therefore ϕi(x) = sin(kix), because the mode shapes are deter-mined up to a factor. The boundary conditions at the right end of the beam result inthe relationship

GIpki cos(kiL) = ω2i Θ sin(kiL) (3.77)

or after rewriting

Page 122: Introduction To Dynamics

114 3 Linear Continuous Models

ki

(GIp

ω2i Θ

)= ki

(GIp

ω2i Jp

)(Jp

Θ

)= tan(kiL) . (3.78)

With k2i from (3.73), an eigenvalue equation (Fig. 3.6) follows:

tan(kL)

(kL)π 2π 3π

( 1kL

)( JpLΘ

)

Fig. 3.6. Eigenvalue determination.

tan(kiL) =

(1

kiL

)(JpLΘ

). (3.79)

Taking into account the initial conditions, we get for the solution

ϕ(x, t) =∞

∑i=1

sin(kix)qi(t) = q(t)Tϕ(x) . (3.80)

The first three modes are shown in Fig. 3.7.For the case of a vanishing end mass (Θ = 0), at the free end it is

GIp

(∂ϕ∂x

)L= 0 , (3.81)

thus ϕ ′(L, t) = 0. This yields cos(kiL) = 0 and

(kiL) = (2i− 1)π2. (3.82)

Considering the initial conditions, the solution satisfies

ϕ(x, t) =∞

∑i=1

sin[(2i− 1)

π2

( xL

)]qi(t) . (3.83)

Page 123: Introduction To Dynamics

3.2 Simple Examples of Continuous Vibrations 115

ϕ(x)

i = 1

( JPLΘ ) = 1

ω1 = 0,8605√

GIPLΘ

ϕ(x)

i = 2ω2 = 3,4256

√GIPLΘ

ϕ(x)

i = 3ω3 = 6,4373

√GIPLΘ

x

x

x

L

L

L

Fig. 3.7. Modes of the torsion vibration.

3.2.4 Transverse Vibrations of a String

The application range for string vibrations and for their model representation isvery large, running from textile machines, belt drives in engines and continuouslyvariable transmissions to high voltage cables for power transmission and carriercables of cable cars.

By a string, we mean a one-dimensional elastic continuum, for example in theform of a thread, which is prestressed by a constant force (Fig. 3.8). The thread isclamped at x = 0 and at x = L. Oscillations are assumed to occur only in the direc-tion transverse to the line continuum, they are small. Additional displacements inlongitudinal direction due to the transverse vibrations are negligible. The vibrationdisplacement is w(x, t), and it depends on the location x and on the time t. The mass

x = 0 x = LL

x

w(x, t)x

F F

z

w(x, t)

dxF

Fw

ϕ(x, t)ϕ(x+dx, t)

Fig. 3.8. Transverse vibrations of a prestressed string.

Page 124: Introduction To Dynamics

116 3 Linear Continuous Models

per unit length, also called line density, is constant μ = ρA with density ρ and cross-sectional area A. Assuming small angles ϕ and applying the momentum theorem toa free piece of thread result in

μdx(∂ 2w∂ t2 ) = Fϕ(x+ dx, t)−Fϕ(x, t) . (3.84)

It is ϕ ≈ ( ∂w∂x ) and therefore ϕ(x+ dx, t)−ϕ(x, t) ≈ ( ∂

2w∂x2 )dx. This yields the so-

called wave equation:

μ(∂ 2w∂ t2 )−F(

∂ 2w∂x2 ) = 0 or (

∂ 2w∂ t2 )− (

Fμ)(

∂ 2w∂x2 ) = 0 . (3.85)

This differential equation is valid for a large number of vibration phenomena inphysics. The value (F

μ ) = c2 is the square of the wave velocity. The standard formof the initial boundary value problem for the wave equation is therefore

w(x, t)− c2w′′(x, t) = 0 , (3.86)

w(0, t) = 0 , w(L, t) = 0 , (3.87)

w(x,0) = w0(x) , w(x,0) = w0(x) . (3.88)

A fundamental solution is

w(x, t) =∞

∑i=1

wi(x) [ai sin(ωit)+ bi cos(ωit)] (3.89)

with

wi(x) = Ai sin(ωi

cx)+Bi cos(

ωi

cx) . (3.90)

From the boundary conditions, we get the eigenfunctions wi(x) and the eigenfre-quencies ωi:

wi(x) = sin(π i(

xL)), (3.91)

ωi =π iL

√Fμ

. (3.92)

The unknown values ai and bi can be determined from the initial conditions. Foran infinitely long prestressed string, the wave equation (3.85) can be solved by themethod of characteristics [76]. This corresponds to a left and a right moving wavewith wave speed c.

Page 125: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 117

3.3 Approximation of Continuous Vibration Systems

The vibrations of elastic continua often cannot be calculated analytically. In thiscase, we are looking for replacement systems that reflect the oscillatory behaviorof the continuum with sufficient accuracy or which lead to an easier handling ofthe descriptive equations. The best known and most universal example is the finiteelement method (FEM), in which the continuum is replaced by a finite number ofmaterial elements (discretization). In this book, we do not address the FEM [12, 9].Instead, we deal with two best practice methods for the treatment of continuousvibration systems, which moreover play a central role in the FEM: the methods ofRAYLEIGH-RITZ and of BUBNOV-GALERKIN.

Both methods have been developed from mechanical problems. In his book [56]from 1877, RAYLEIGH described vibrations of elastic continua using a series ofeigenmodes. RITZ abstracted the concept in 1908 [57] by formulating the equationsas a variational problem, which he traced back (by approximating the integrand withmode shapes) to the minimization of a function with several variables. Indepen-dently in 1915, GALERKIN dealt with balance problems for thin plates. He approx-imated the functions appearing in the partial differential equations by suitable peri-odic functions, or as we know it today by the method of weighted residuals, and thusreduced the problem to the solution of algebraic equations [17]. According to [18]however, BUBNOV discovered the method of weighted residuals for the first timein 1913 for the approximation of partial differential equations. Today, we usuallyapply an additional integration by parts starting from the method of weighted resid-uals; based on this, we call the usage of finite-dimensional approximation spacesthe BUBNOV-GALERKIN method (historically perhaps not totally correct). It is thestarting point of the finite element method [12, 9]; in 1927, COURANT used hatfunctions as trial functions for the first time [13]. We start by providing some basicfunctional analytical background.

3.3.1 Function Systems and Completeness

Each scalar p-periodic function f (x) = f (x+ p) can be represented as a FOURIER

series [76]. In practice, the series is often truncated after a finite number of elements;we obtain an approximation of f with the help of a trigonometric polynomial:

f (x) ≈ fN(x) =N

∑i=−N

qiejωix with ω =

2πp

, qi =1p

∫ p

0f (x)e− jωixdx .

(3.93)

The trigonometric polynomial is a linear combination of the linearly independenttrial functions wi(x) = e jωix. The trial functions are usually summarized in a so-called function system {wi}i. The set of all possible linear combinations, which alsoincludes fN(x), forms a linear space, a so-called vector space. This can be extendedby a scalar product: for elements of the function system, wi and wk, we define

Page 126: Introduction To Dynamics

118 3 Linear Continuous Models

< wi,wk >=1p

∫ p

0wi(x)w

∗k(x)dx , (3.94)

with w∗k(x) being complex conjugate to wk(x). Substituting wk = wi, it follows

‖ wi ‖=√< wi,wi > (3.95)

and one obtains a norm for this linear space.Convincing ourselves with the example of FOURIER series, it holds

< wi,wk >= δik . (3.96)

With respect to the scalar product defined above, the basis functions wi are orthonor-mal: they form an orthonormal system. The FOURIER coefficients satisfy

qi =< f ,wi > . (3.97)

They are obtained by minimizing the error

ΔN =‖ f − fN ‖2 (3.98)

between the exact solution f (x) and the approximate solution fN(x). For the mini-mum approximation error, we obtain

ΔN,min =‖ f ‖2 −N

∑i=−N

q2i ‖ wi ‖2 . (3.99)

The orthonormal system is called complete, if limN→∞ ΔN,min = 0, that is if the so-called PARSEVAL theorem is satisfied

‖ f ‖2=∞

∑i=−∞

q2i ‖ wi ‖2 . (3.100)

This is the property of the representability of any scalar p-periodic function f withthe help of the function system {wi}i as FOURIER series.

For vector functions fl (x1, · · · ,xn) with l ∈ {1, · · · ,m}, the trial functions arelinearly independent vector functions:

wi (x) =

⎛⎜⎝

wi,1 (x)...

wi,m (x)

⎞⎟⎠ with x = (x1, · · · ,xn)

T . (3.101)

Concerning vibration problems, we usually have dependencies on x and t and avector function f = f(x, t), that is qi = qi(t). We met similar separation of variableapproaches in Section 3.2.1; we now consider them for the calculation of approxi-mations.

Page 127: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 119

3.3.2 Rayleigh-Ritz Method

We consider linear elastic continua, the behavior of which can be described by thevariational problem of HAMILTON’s principle (1.174)[28]

∫ t2

t1(T −V)dt → stationary . (3.102)

The expression

T =12

∫K

(uT u

)dm (3.103)

is the kinetic energy of the continuum K,

V =12

∫K

1E

(σ2

x +σ2y +σ2

z

)dxdydz

− 12

∫K

2μE

(σxσy +σyσz +σzσx)dxdydz

+12

∫K

1G

(τ2

xy + τ2yz + τ2

zx

)dxdydz (3.104)

is the potential energy of K (Fig. 3.9). Where E is Young’s modulus, G is theshear modulus, μ is Poisson’s ratio, and u(r, t) is a (linear-elastic) displacementof the position vector to the material point from its (stationary) reference position r

Ix

Iy

IzK(t)

K(= reference position)

r

σz

σy

σx

dmu(r, t)

material point in thereference position

material point afterdisplacement

Fig. 3.9. Elastic deformation of a material point.

Page 128: Introduction To Dynamics

120 3 Linear Continuous Models

(LAGRANGE view). The stresses σx, σy σz, τxy, τyz, and τzx depend on the displace-ments u(r, t) due to the material law [12, 6, 69].

According to RITZ, we do not solve the variational problem (3.102) for the dis-placement u exactly. Instead, the displacement field u is approximated with a finitelinear combination uN of trial functions wi (r) from a function system; therefore, thevariational problem is reduced to a system of ordinary differential equations. Withthe coefficients qi(t), we approach uN by

uN (r, t) =N

∑i=1

wi (r)qi(t) . (3.105)

The trial functions wi (r) are known. Then, T and V depend only on the coefficientfunctions qi(t), which will be determined such that the approximated functional(3.102) is stationary. This problem leads to LAGRANGE’s equations of the secondkind (1.151)

ddt

(∂T∂ qi

)− ∂T

∂qi+

∂V∂qi

= 0 (3.106)

for the coefficient functions qi. As uN = ∑i wi (r) qi(t), the kinetic energy T , whichis approximated by the trial functions wi, only depends on qi. Then, (3.106) reducesto

ddt

(∂T∂ qi

)+

∂V∂qi

= 0 . (3.107)

We consider only continuous oscillations, where the displacement field u is planar,that is

u(r, t) = (0,0, w(r, t))T . (3.108)

Then,

wN (r, t) =N

∑i=1

wi (r)qi(t) = q(t)T w(r) (3.109)

with real-valued trial functions wi (r):

qT = (q1,q2, · · · ,qN) , (3.110)

wT = (w1,w2, · · · ,wN) . (3.111)

Furthermore,

T =12

qT(∫

Kw(r)w(r)T dm

)q =

12

qT Mq , (3.112)

which according to (3.107) leads to the following equations of motion –

Page 129: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 121

RAYLEIGH-RITZ method:

Mq+

(∂V∂q

)T

= 0 with M =

(∫K

w(r)w(r)T dm

). (3.113)

The determination of the exact displacement field u (r, t) from the variationalproblem (3.102) is traced back to the determination of an approximate displacementfield uN (r, t), whose explicit representation is known with the solution of (3.113).

In the static case, it is T = 0 and we obtain the equilibrium conditions

(∂V∂q

)T

= 0 . (3.114)

The approximation method can be interpreted that among the infinitely many ap-proximation systems, the RAYLEIGH-RITZ method picks the one, which satisfiesthe principle of d’ALEMBERT in the dynamic case (the LAGRANGE’s equations)and the principle of virtual work (the equilibrium conditions) in the static case [77].

Example 3.1 (Cantilever beam according to RAYLEIGH-RITZ [69]). Thebending vibrations of a cantilever beam are treated analytically in Section 3.2.1(Fig. 3.10). We apply the RAYLEIGH-RITZ method. It is

x

L

Fig. 3.10. Cantilever beam.

T =12

∫ L

0ρA

(∂w∂ t

)2

dx , V =12

∫ L

0EI

(∂ 2w∂x2

)2

dx .

We choose the approach

w(x, t) = qT (t)w(x) .

The vectors q(t) and w(x) have only three components:

Page 130: Introduction To Dynamics

122 3 Linear Continuous Models

q(t) =

⎛⎝q1 (t)

q2 (t)q3 (t)

⎞⎠ , w(x) =

⎛⎜⎝(

xL

)2(xL

)3(xL

)4

⎞⎟⎠ .

The trial functions represent the static bending line for a constant uniform load.From

V =12

qT (t)

{EI∫ L

0w′′(x)w′′T (x)dx

}︸ ︷︷ ︸

K

q(t) ,

it follows for the derivative(∂V∂q

)T

= Kq .

This gives the equation of motion

Mq+Kq = 0 with M =

∫ L

0ρAwwT dx

and the matrices

M =

⎛⎝

15

16

17

16

17

18

17

18

19

⎞⎠(ρAL) , K =

⎛⎝4 6 8

6 12 188 18 144

5

⎞⎠(

EIL3

).

With the approach

q = qeλ t ,

we obtain the characteristic equation for the eigenvalues and the system of linearequations for the eigenvectors:

det(Mλ 2 +K

)= 0 ,

(Mλ 2 +K

)q = 0 .

We get the following eigenvalues:

RAYLEIGH-RITZ exact rel. error(k1L) 1,876 1,875 0,05 %(k2L) 4,712 4,694 0,38 %(k3L) 19,261 7,855 145 %

The last line confirms the experience that one cannot determine the nth eigenvaluewith n trial functions. The eigenvalues of the discrete system are larger than thoseof the continuous system, because the RAYLEIGH-RITZ method is consistently de-fined by an integral projection. The restriction to a finite number of trial functionsacts as a constraint and yields a stiffening of the system. For a comparison of theapproximated mode shapes

Page 131: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 123

wi(x, t) = eλit qTi w(x)

with the exact mode shapes see Fig. 3.11.

1

exact solutionapproximation with RAYLEIGH-RITZ

2

3

Fig. 3.11. Mode shapes (exact and approximation with RAYLEIGH-RITZ).

If we choose the mode shapes as trial functions in Example 3.1, we obtain a solutionup to the corresponding order and diagonal mass and stiffness matrices. In general,we do not know the mode shapes. The trial functions have to be chosen arbitrarily,but with the boundary conditions to be satisfied (Section 3.3.4). We focus on globaltrial functions; for the finite element method, one selects local trial functions.

3.3.3 Bubnov-Galerkin Method

The models in mechanics lead to differential equations that represent balances ofmomentum and/or moment of momentum changes, that is sums of forces and/ormoments. This also applies to continuum mechanics, where one has to deal withpartial differential equations.

The approximation method of BUBNOV-GALERKIN does not primarily concerna variational problem, like (3.102), but the so-called strong form, that is a partial oran ordinary differential equation

D [u] = 0 (3.115)

with the differential operator D, which we assume to be linear. We suppose that thedisplacement function u acts only in one plane and is thus defined by one componentw of u.

Example 3.2 (Prestressed vibrating string: differential operator D). For the pre-stressed string (Fig. 3.12) with pretension force F defined in the reference state K,line density μ and deflection w(r, t) at the location r = (x,0,0)T and at the time t,we have obtained the wave equation as a descriptive differential equation in Sec-tion 3.2.4:

Page 132: Introduction To Dynamics

124 3 Linear Continuous Models

F

z

x = 0

x

x = L

K(t)

K

r

w(r, t)

F

Fig. 3.12. Prestressed string.

∂ 2w∂x2 − μ

F∂ 2w∂ t2 = 0 .

Since the position vector r represents a line continuum, the material points are de-fined by specifying only one coordinate, that is the curve parameter x for the refer-ence configuration. The differential operator D is

D =∂ 2

∂x2 − μF

∂ 2

∂ t2 .

We now assume that wN is an approximation for the exact solution w of D [w] = 0and use a function system with known trial functions wi but with coefficient func-tions qi to be determined:

wN(x, t) =N

∑i=1

wi(x)qi(t) = q(t)T w(x) . (3.116)

To obtain equations for the determination of the functions qi, we require that thescalar product of the residual

rN = D [wN ]−D [w] = D [wN ] (3.117)

with given weighting functions gk (k = 1, · · · ,N) vanishes –

method of weighted residuals:

< rN ,gk >=< D [wN ] ,gk >= 0 . (3.118)

The scalar product < ·, · > is part of the function system of the trial functions(3.94). Up to a factor, equation (3.118) means to multiply the residual by the weight-ing functions and to integrate over the position variable.

If the weighting functions gk belong to a complete function space, then we canrepresent r = limN→∞ rN itself as a series of the weighting functions gk. Further,

Page 133: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 125

< r,gk >= 0 for (k = 1,2, · · ·) , (3.119)

that is r is orthogonal to every gk. This implies limN→∞ ‖ rN ‖= 0. So, the residualrN converges to 0 for N → ∞.

If the approximation wN satisfies all the boundary conditions (Section 3.2), whichare satisfied by the exact function w, then (under certain assumptions on the differ-ential operator D [9]) from the convergence rN → 0 follows also the convergencewN → w:

limN→∞

‖ wN − w ‖= 0 . (3.120)

The method of weighted residuals is an intermediate step. Depending on how theweighting functions gk are chosen, a number of methods can be derived. We focuson the classic BUBNOV-GALERKIN method and assume [76]

gk = wk . (3.121)

The weighting functions correspond to the trial functions with which the solution wis approximated. We split the differential operator into a time derivative componentand a spatial derivative operator K [w] and demonstrate the step to the BUBNOV-GALERKIN method with Example 3.2.

Example 3.3 (Prestressed vibrating string: BUBNOV-GALERKIN method). Thedifferential operator D is composed of a time derivative component

−μF

∂ 2

∂ t2

and a spatial derivative operator

K =∂ 2

∂x2 .

The method of weighted residuals for gk = wk is

0 =< D[N

∑i=1

wi(x)qi(t)],wk(x)>=

∫ L

0D[

N

∑i=1

wi(x)qi(t)],wk(x)dx

=N

∑i=1

∫ L

0−μ

Fwi(x)

∂ 2qi(t)∂ t2 wk(x)+K[wi(x)]qi(t)wk(x)dx

=−μF

N

∑i=1

∫ L

0wi(x)wk(x)dx

∂ 2qi(t)∂ t2 +

N

∑i=1

∫ L

0

∂ 2wi(x)∂x2 wk(x)dxqi(t) .

The trial and weighting functions wi and wk are known or chosen, such thatthe boundary conditions are satisfied for the system under consideration (Sec-tion 3.3.4). Then, the integrals can be calculated in advance and we obtain an

Page 134: Introduction To Dynamics

126 3 Linear Continuous Models

ordinary differential equation for the time-dependent coefficients qi in a similar wayas for the RAYLEIGH-RITZ method

Mq = Kq

with mass and stiffness matrices

Mki =μF

∫ L

0wi(x)wk(x)dx , Kki =

∫ L

0

∂ 2wi(x)∂x2 wk(x)dx .

A drawback is the lack of symmetry of the stiffness matrix because of the second-and zeroth-order spatial derivatives of the trial or weighting functions. This is amajor difference to the RAYLEIGH-RITZ method, which starts directly from thevariational principle (3.102). We shift derivatives from the trial functions to theweighting functions by integration by parts until we get the desired symmetry:

μF

N

∑i=1

∫ L

0wi(x)wk(x)dx

∂ 2qi(t)∂ t2 =−

N

∑i=1

⎡⎢⎢⎢⎣∫ L

0

∂wi(x)∂x

∂wk(x)∂x

dx+∂wi

∂xwk

∣∣∣∣L

0︸ ︷︷ ︸=0

⎤⎥⎥⎥⎦qi(t) .

The boundary integral vanishes in our example, since the string is clamped on bothsides and therefore the corresponding trial and weighting functions have to be cho-sen appropriately, we say they are admissible. We finally obtain the

BUBNOV-GALERKIN method

Mq+Kq = 0

with the mass and stiffness matrices

Mki =μF

∫ L

0wi(x)wk(x)dx , Kki =

∫ L

0

∂wi(x)∂x

∂wk(x)∂x

dx

like in the RAYLEIGH-RITZ method.

Generally, the BUBNOV-GALERKIN method results from the method of weightedresiduals by integration by parts, which is performed until the resulting differen-tial operator has as a maximum of symmetry with respect to trial and weightingfunctions. The boundary integrals include so-called natural boundary conditions.We discuss the treatment of boundary conditions for all methods in the followingsection.

Page 135: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 127

3.3.4 Boundary Conditions for the Rayleigh-Ritz andBubnov-Galerkin Method

The solution of a variational problem or a partial differential equation must be sup-plemented by initial and boundary conditions. We characterize boundary conditionsby an operator equation

R [w] = 0 (3.122)

with

R = (R1, · · · ,Rm)T . (3.123)

The operator R contains as many suboperators as the problem has boundary condi-tions.

Example 3.4 (Cantilever beam: boundary conditions). By analogy with Exam-ple 3.14, we consider Fig. 3.13. We model geometric boundary conditions on theleft and kinetic boundary conditions on the right:

R [w] =

⎛⎜⎜⎝

w(0, t)w′(0, t)

w′′(L, t)+ MEI

w′′′(L, t)+ FEI

⎞⎟⎟⎠= 0

}kinematic ,}kinetic .

We denote boundary conditions, where only the boundary values of w and w′ occur,as geometric, kinematic, or essential boundary conditions. Accordingly, we denoteboundary conditions, where also the boundary values for w′′, w′′′ occur, as free,kinetic, or natural boundary conditions.

For the convergence wN → w in the method of weighted residuals, we have therequirements in accordance with Section 3.3.3:

1. the N trial and weighting functions have to be from a complete function system,2. all boundary conditions imposed on w have to be satisfied also by wk.

reference

z

w(x, t)

x

F

M

Fig. 3.13. Cantilever beam with constant end load and constant end torque.

Page 136: Introduction To Dynamics

128 3 Linear Continuous Models

Since the method of weighted residuals is based on the differential equation itself,the trial functions have to be admissible in the sense that they meet both the geo-metric and the free boundary conditions.

The BUBNOV-GALERKIN method is constructed from the method of weightedresiduals by integration by parts. We choose gk = wk with gk = 0 and g′k = 0 atgeometric boundaries. Geometric boundary conditions vanish during the integrationby parts, free boundary conditions are preserved naturally. So, wN converges to w,if

1. the N trial and weighting functions wk are from a complete function system,2. trial functions wk satisfy the geometric boundary conditions.

The BUBNOV-GALERKIN method is based on the so-called weak form and theonly requirement is fulfilling the geometric constraints regarding admissibility. TheRAYLEIGH-RITZ method behaves accordingly. After defining the energy, for exam-ple, the corresponding potentials for the description of free boundary conditions,admissible trial functions must satisfy only the geometric boundary conditions.

Example 3.5 (Cantilever beam according to BUBNOV-GALERKIN [69]). In Ex-ample 3.1, we have treated the cantilever beam with the RAYLEIGH-RITZ method.Now we apply the BUBNOV-GALERKIN method. The differential equation is

D [w] = EI

(∂ 4w∂x4

)+ρA

(∂ 2w∂ t2

)= 0 .

As in Example 3.1, the boundary conditions satisfy

w(0, t) = w′(0, t) = 0 ,

w′′(L, t) = w′′′(L, t) = 0 .

As an approach, we choose w(x, t) = qT (t)w(x) with the trial functions of the staticbending line:

w(x) =

⎛⎝( x

L )2

( xL )

3

( xL )

4

⎞⎠ .

This approach satisfies the geometric boundary conditions. We insert it into thedifferential equation, multiply it by the weighting functions, and integrate over thelength of the beam:

∫ L

0

(EI

∂ 4wT

∂x4 q+ρAwT q)

wkdx = 0 .

Integration by parts with wk(0) = 0 and w′k(0) = 0 yields

Page 137: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 129

∫ L

0

(EIw′′

k∂ 2wT

∂x2 q+ρAwkwT q)

dx+

(EIwk

∂ 3wT

∂x3 q)

L︸ ︷︷ ︸force shifted by wk

−(

EIw′k∂ 2wT

∂x2 q)

L︸ ︷︷ ︸moment rotated by w′

k

= 0 .

The loads at the free boundary disappear in our example. Thus, we obtain the dif-ferential equation system of the RAYLEIGH-RITZ method

Mq+Kq = 0

from Example 3.1

3.3.5 Choice of Trial Functions

The efficiency of the approximation methods presented is highly dependent on thechoice of the trial functions w. Admissible trial functions must satisfy the geo-metric boundary conditions in each case. For the RAYLEIGH-RITZ or BUBNOV-GALERKIN method, the highest derivative degree is half the one occurring for themethod of weighted residuals due to energy expressions or integration by parts.Admissible trial functions have to have only half differentiability. Because of theenergy expressions or integration by parts, free boundary conditions are also in-cluded. These have to be provided for admissibility only in the method of weightedresiduals.

Since we have limited ourselves to global trial functions, no further rules can bespecified. We choose the trial functions to be as simple as possible. We try to ensurethat the trial functions are qualitatively consistent with the expected mode shapesin reality, as we can expect satisfactory mathematical convergence only in this case.Frequently, it may be useful to choose the mode shapes of a replacement problemas trial functions, for which the vibration characteristics are close to the continuouscomponent in the overall system. Such a replacement problem can be difficult todefine (e.g., a rotating rod). In such cases, it might be better to choose trial functionsfrom the physical intuition and adjust them iteratively.

In terms of automation and flexibility, finite element methods are preferable.They are based on the weak form and do not use global but local, mostly piece-wise polynomial, interpolation functions. These can be locally refined and adaptedautomatically to load and stress curves [12, 9].

3.3.6 Bending Vibrations of a Beam with Longitudinal Load

We consider the bending vibrations of a beam with longitudinal load as shown inFig. 3.14. Approximations are derived with the RAYLEIGH-RITZ and the BUBNOV-GALERKIN method.

Page 138: Introduction To Dynamics

130 3 Linear Continuous Models

L

x

z

F

w(x, t)

u(L, t)

dx

u(x, t)

w(x, t)

w′dx

ds

u(x+dx, t)

w(x+dx, t)

dx−

u′dx

Fig. 3.14. Beam with longitudinal load and beam element.

3.3.6.1 Rayleigh-Ritz Method

Three energy terms are required for the RAYLEIGH-RITZ method:

• kinetic energy

T =12ρA

∫ L

0w2(x, t)dx , (3.124)

• bending potential

V =12

EI∫ L

0w′′2(x, t)dx , (3.125)

• potential of the pressure force

V =−Fu(L, t) =−F∫ L

0

12

w′2(x, t)dx . (3.126)

The potential of the pressure force decreases the stiffness and thus the bending po-tential for positive F . We express the longitudinal displacement u(L) with the de-flection w. Therefore, we need the following relations:

u(x+ dx, t)≈ u(x, t)+ u′dx , (3.127)

w(x+ dx, t)≈ w(x, t)+w′dx . (3.128)

This yields (Fig. 3.14)

(dx− u′dx)2 +(w′dx)2 = ds2 (3.129)

Page 139: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 131

neglecting the longitudinal deformation ds ≈ dx

(1− u′)2 +w′2 = 1 . (3.130)

We obtain a formula for

u′ = 1−√

1−w′2 ≈(

w′2

2

)(3.131)

and finally

u(L, t) =∫ L

0u′dx ≈

∫ L

0

(w′2

2

)dx . (3.132)

The approach

w(x, t) = q(t)T w(x) (3.133)

is inserted into the energy expressions:

T =12

qT ρA∫ L

0w(x)wT (x)dx

︸ ︷︷ ︸M

q , (3.134)

V =12

qT[

EI∫ L

0w′′(x)w′′T (x)dx−F

∫ L

0w′w′T (x)dx

]︸ ︷︷ ︸

K

q . (3.135)

With LAGRANGE’s equations of the second kind, we obtain the equation of motion

Mq+Kq = 0 . (3.136)

The trial functions w(x) must satisfy the geometric boundary conditions:

w(0) = 0 , w(L) = 0 . (3.137)

If we select just one trial function, the matrices of the equation of motion becomescalars. The trial function

w1(x) = sin(πx

L

)(3.138)

satisfies the required boundary conditions. With this trial function, we obtain:

Page 140: Introduction To Dynamics

132 3 Linear Continuous Models

M = ρA∫ L

0sin2

(πxL

)dx =

12ρAL , (3.139)

K = EI∫ L

0

(πL

)4sin2

(πxL

)dx−F

∫ L

0

(πL

)2cos2

(πxL

)dx

=12

EIL(π

L

)4− 1

2FL

(πL

)2. (3.140)

Thus, the eigenvalue problem det(Mλ 2 +K

)= 0 transforms to

12ρALλ 2 +

(π2

2L

)[EI(π

L

)2−F

]= 0 . (3.141)

With λ 2 = −ω2 for complex conjugate eigenvalues, we obtain the eigenfrequency,which depends on the pressure force F :

ω =

√(π2

ρAL2

)[π2EI

L2 −F

]. (3.142)

We study some cases below:

• The critical buckling load is reached when the eigen angular frequency disap-pears (ω = 0). Then the re-acting force in the transverse direction vanishes:

Fkrit = π2(

EIL2

). (3.143)

Now, we can write

ω2 =

(π2

ρAL2

)(Fkrit −F) . (3.144)

• For F = 0, we obtain the eigen angular frequency of a beam which is simplysupported on both sides:

ω20 =

EIπ4

ρAL4 . (3.145)

• The eigen angular frequency of a prestressed string is obtained with vanishingbending stiffness EI = 0 and tensile forces F < 0:

ω2 =Fπ2

ρAL2 . (3.146)

Fig. 3.15 shows the eigen angular frequency as a function of the pressure force F .

Page 141: Introduction To Dynamics

3.3 Approximation of Continuous Vibration Systems 133

ω2

FFkrit

stringω2

0

EI =0, F

<0

Fig. 3.15. Eigen angular frequency as a function of the pressure force F .

3.3.6.2 Bubnov-Galerkin Method

To apply the BUBNOV-GALERKIN method, we first derive the equations of motionusing the beam element in Fig. 3.16. We need the following relations:

x

z

∂w∂x dx

M+ ∂M∂x dx

F

Q

Q+ ∂Q∂x dx

w

M

dx

F

Fig. 3.16. Beam element.

• moment of momentum equation with neglected rotational inertia (about upperendpoint):

(∂M∂x

)dx−Qdx−F

(∂w∂x

)dx = 0 , (3.147)

• bending moment:

Page 142: Introduction To Dynamics

134 3 Linear Continuous Models

M =−EI

(∂ 2w∂x2

), (3.148)

• momentum equation in the z-direction:

ρA

(∂ 2w∂ t2

)dx =

(∂Q∂x

)dx . (3.149)

Finally, we obtain the partial differential equation:

ρA∂ 2w(x, t)

∂ t2 +EI∂ 4w(x, t)

∂x4 +F∂ 2w(x, t)

∂x2 = 0 . (3.150)

Integration by parts in the method of weighted residuals

∫ L

0wk

(EIw(4) +Fw′′+ρAw

)dx = 0 (3.151)

yields

0 =

∫ L

0EIw′′

k w′′dx+(EIwkw′′′)L

0︸ ︷︷ ︸=0

−(w′kM)L

0

−∫ L

0Fw′

kw′dx+(Fwkw′)L

0︸ ︷︷ ︸=0

+

∫ L

0ρAwkwdx . (3.152)

The trial functions

w(x, t) = q(t)T w(x) (3.153)

must satisfy the geometric boundary conditions

w(0) = 0 , w(L) = 0 . (3.154)

Inserting the free boundary conditions (M)L0 = 0 yields

ρA∫ L

0wwT dx

︸ ︷︷ ︸M

q+[

EI∫ L

0w′′w′′T dx−F

∫ L

0w′w′T dx

]︸ ︷︷ ︸

K

q = 0 . (3.155)

This gives the same equation of motion as for the RAYLEIGH-RITZ method:

Mq+Kq = 0 . (3.156)

Page 143: Introduction To Dynamics

3.4 Vibrations of Elastic Multibody Systems 135

3.4 Vibrations of Elastic Multibody Systems

Most mechanical systems in practice consist of both, bodies, which we can assumeto be rigid, and bodies, the deformation of which must be considered. From themany techniques for the representation of such systems, we choose, seen from anengineering point of view, the one that offers maximum transparency and closenessto reality with minimum effort. A measure is the minimum number of degrees offreedom, which just describe the motion of the system adequately.

The division into rigid and nonrigid bodies already guarantees to minimize thedegrees of freedom in a first step, because by modeling a part of the system com-ponents as rigid bodies the smallest possible number of degrees of freedom isused. In the second step, the nonrigid system components have to be modeled witha minimum number of degrees of freedom such that satisfactory realism can beachieved. For linear elastic systems, the presented methods of RAYLEIGH-RITZ

and BUBNOV-GALERKIN are based on global (modal) trial functions, because onlya small number of trial functions and thus additional elastic degrees of freedomare necessary due to always existing structural damping and its increasing effect onhigher eigenmodes (dissipation energy). One important criterion for selecting theelastic modes are the frequencies of operation of a machine or a structure. Onlythe elasticity of components is of interest, the eigenfrequencies of which are in therange of these operating frequencies.

The equations of motion of these systems can be derived with the methods pre-sented in Chapter 1. The equations for rigid multibody systems must be extendedby the nonrigid components. For the deformations of linear-elastic systems an ap-proximate approach in the sense of RAYLEIGH-RITZ or BUBNOV-GALERKIN canalways be found, which accurately describes the individual deformations. Withoutgoing into the mathematical details, the procedure will be outlined below:

1. Setting up a mechanical model.2. Defining the rigid and elastic (nonrigid) bodies.3. Finding the best coordinate systems.4. Setting up the free-body diagram with relevant forces.5. Evaluating the relative kinematics with all deformation effects (absolute veloc-

ities and accelerations, transformation matrices).6. Discretization of the elastic components (choice of trial functions, number of

elastic degrees of freedom).7. Choice of generalized coordinates for the rigid bodies.8. Determination of JACOBIANs of translation and rotation.9. Derivation of the projected equations of motion.

10. Analytical and numerical treatment of the equations of motion (first integrals,linearization, analytical or numerical solution).

A detailed consideration of such problems can be found in [11, 10, 59, 51].Fig. 3.17 shows a typical example of an elastic multibody system [55]. The

Ravigneaux planetary gearset is used in automatic transmissions of motor vehicles

Page 144: Introduction To Dynamics

136 3 Linear Continuous Models

clutch / brakewedge shaft

torsional radial

toothing

roller bearingaxial radial

S1S2

P2

P1

F

A

y

S2 P1

PT

S1

H

H

P2

z PT

Fig. 3.17. Mechanical model of a Ravigneaux planetary gearset [55].

primarily because of the small installation space and the variety of switching op-tions. By selecting different drive shafts, different overall transmission ratios can berealized.

The main components of the Ravigneaux planetary gearset are a small sun S1 anda large sun S2 (small central gears), a ring H (large central gear), a planet carrierPT (rack) with short planets P1 and long planets P2, the drive shafts A, and the freecoupling shafts F . The large sun S2, the planets P2, and the ring H together withthe planet carrier PT represent a simple planetary gear. The inner planets P1 aremounted on the same rack. The power is transmitted through the long planets P2 tothe common ring H.

In the planetary gear shown in Fig. 3.17, the output goes via the ring H. For thecoupling of the ring H to the output shaft, several design variations exist. Very often,the ring H is welded to the output shaft. In this case, the planar elastic deformationof the ring is very small due to the radial stiffening; it is therefore allowed to modelthe ring as a rigid body. However, if a thin-walled ring is coupled to the outputvia a gear teeth set according to Fig. 3.17, it may deform due to the presence ofradial and flank clearance and must be modelled as an elastic body. Analogously tosimple planetary gearsets, the gears and the rack PT are modelled as rigid bodieswith 4 or 6 degrees of freedom. For the reduced planetary gearset, the coupling ofthe individual bodies with each other and with the input and output is modelled bytooth couplings, bearings, and wedge shaft connections.

Page 145: Introduction To Dynamics

Chapter 4Methods for Nonlinear Mechanics

4.1 General Remarks

The motion of any mechanical system composed of rigid bodies or discretized elas-tic bodies is always described by a set of nonlinear ordinary differential equationsof second order of the form (Chapter 1)

M(q, t) q = h(q, q, t) . (4.1)

The vector h contains force or torque expressions, which might depend on veloci-ties q of maximum second order. In spite of the simplicity of these equations, it isimpossible to find a general solution or to find some superposition principle, whichallows us to combine some fundamental solutions like in the linear case (Chapter 2).Though, such a general solution can be evaluated for one-dimensional systems, forwhich the equation of motion is of RICCATI type [33]. For systems with many de-grees of freedom and also for continuum mechanical problems, one cannot avoidnumerical integration routines and/or approximations.

With respect to approximations, one makes use of special features of the dif-ferential equations or of the mechanical system under consideration. This concernssystems with weak nonlinearities or periodical systems. Considering the first case,we may develop the equations of motion with respect to a small parameter resultingin a sequence of differential equations, where each new set represents an improveddescription of the problem compared with the preceding set, convergence antici-pated [47, 48, 26].

Considering nonlinear vibration systems, we make use of their periodicity

q(t) = q(t +T ), (4.2)

where the period T usually is given by nonlinear equations. A couple of methods areavailable for their solution, particularly in connection with limit cycles. We discusssome simple examples.

We distinguish quantitative and qualitative methods. Qualitative methodsestablish some statements concerning general solution properties by geometric

© Springer-Verlag Berlin Heidelberg 2015 137F. Pfeiffer and T. Schindler, Introduction to Dynamics,DOI: 10.1007/978-3-662-46721-3_4

Page 146: Introduction To Dynamics

138 4 Methods for Nonlinear Mechanics

considerations of the equations of motion. Some aspects are stability of motion,periodical behavior, bifurcations of the solution, and possible chaotic developmentsof the motion. They are connected with names like POINCARÉ, LYAPUNOV, THOM,THOMPSON, ARNOLD, and others. Also in the future, geometric considerations ofthe nonlinear equations of motion will contribute significantly to the understandingof the phenomena included by them and from this also to many important applica-tions of practical relevancy [71, 3, 70, 75].

Such a textbook cannot treat nonlinear dynamics in an exhaustive way. We rec-ommend special literature [71, 3, 24, 26, 33, 39, 44]. However with respect to a firstintroduction, we consider some important aspects, which might help when lookingat practical problems. In a first step, we apply four methods to a simple nonlinearoscillator with one degree of freedom (DOF) only. These four methods are

• piecewise exact solutions,• weighted residuals,• harmonic balance,• least squares.

As an example of the qualitative methods, we consider LYAPUNOV’s stability theory.With respect to practice, analytical solutions and approximations make sense only

if they help to understand the problems under consideration by a deeper insight intotheir structures and their mechanical features. This requires a careful choice of meth-ods, especially in the case of nonlinear systems. Not always, it will be advisable touse numerical methods leading sometimes to problems of interpretation. Very sim-ple models often generate information that is hidden in the results of large mod-els requiring long investigations. This situation justifies our presentation of simplemethods applied to a simple example with one degree of freedom.

4.2 Phase Space

For a better understanding of dynamics we often use phase portraits drawn in aphase plane. For this purpose, we select from the state space two state trajectoriesover time and sketch them with respect to each other (projected phase curve). Usu-ally, we select velocity and position states. This type of presentation is also wellsuited to nonsmooth systems. For a one-DOF example with the equation of motion

x =−ω2x , (4.3)

the initial position x0 and vanishing initial velocity, we discuss three methods forachieving a phase portrait. In this case, this corresponds to the whole set of phasecurves.

1. Solution of the equations of motion and elimination of time

x(t) = +x0 cos(ωt) , (4.4)

x(t) =−x0ω sin(ωt) (4.5)

Page 147: Introduction To Dynamics

4.2 Phase Space 139

x

x

Fig. 4.1. Phase portrait of a 1-DOF oscillator.

results in(

xx0

)2

+

(x

ωx0

)2

= cos2(ωt)+ sin2(ωt) = 1 . (4.6)

These equations describe ellipses in the phase plane (Fig. 4.1). For the stationarypoint

(x x

)T=(0 0

)T, the corresponding trajectory remains there for all time.

The sense of rotation of these trajectories follows from the functional relationof x and x: for example, the value of x increases for a positive velocity x > 0.The phase curves will not have an intersection as long as the right-hand sideof the initial value problem is smooth enough. Every phase point in Fig. 4.1 ispart of one phase curve only. Therefore, Fig. 4.1 completely describes the phaseportrait. However, we should keep in mind that phase portraits represent a planecut of the total phase space. In general, intersecting projected phase curves arepossible (see for example the book cover).

2. Integration of the slope and conservation of energyThe slope of a phase curve

dxdx

=xx=−ω2x

x(4.7)

can be remodeled by separation of variables:

ω2xdx =−xdx . (4.8)

Integration results in

12ω2x2 +

12

x2 =12ω2x2

0 =Em

, (4.9)

which is the already considered ellipse equation. Every ellipse in the phase spaceindicates conservation of energy.

Page 148: Introduction To Dynamics

140 4 Methods for Nonlinear Mechanics

3. Isoclines (phase velocity field)According to (4.7), it is

(dxdx

)T (ω2xx

)= 0 , (4.10)

which says, that(dx dx

)Tis perpendicular to

(ω2x x

)T. This feature allows

one to construct a phase portrait. The differential change(dx dx

)T for a point(x x

)Tis indicated as a short line in Fig. 4.1. By considering all these small

lines, we also come out with elliptic phase curves.For an intersection point

(x x

)T=(x 0

)T, we get

dx =− xω2x

dx = 0 , (4.11)

which means that phase curves intersect the x-axis always in a perpendicular way.

4.3 A 1-DOF Nonlinear Oscillator

We consider a 1-DOF oscillator with a nonlinear restoring force r(x) (Fig. 4.2). Noother forces are applied, and the equations of motion write

mx =−r(x) . (4.12)

With x = dxdt (definition velocity) and dt = 1

x dx (elimination of time), we get

mxdx =−r(x)dx . (4.13)

After the separation of x and x, an integration is possible and results in

12

mx2 = E0 −∫

r(x)dx (4.14)

with E0 := 12 mx2

0. From this

nonlinear spring r(x)

mass m

r(x)

m

x x

Fig. 4.2. A 1-DOF nonlinear oscillator.

Page 149: Introduction To Dynamics

4.3 A 1-DOF Nonlinear Oscillator 141

x =

√2m

(E0 −

∫r(x)dx

)(4.15)

and finally

t(x) = t0 +∫ x

0

dxx

, (4.16)

we determine the reversal x(t).

r(x)

x

Fig. 4.3. Signum function.

These relations are general and will be used for the four methods discussed in thefollowing. For this purpose, we choose a specific nonlinear restoring force

r(x) = r · sgn(x) , (4.17)

with constant r. It is depicted in Fig. 4.3.

Example 4.1 (Piecewise constant restoring force and corresponding solution).The following three nonlinear oscillators possess restoring forces according to(4.17). It is very easy to test that by simple experiments. The first example of arod rocking on a block from left to right and from right to left can easily be calcu-lated, if we neglect the impact losses accompanying this oscillation (Fig. 4.5). Themoment of momentum equation about R is

JRα =−mgb2

. (4.18)

S

α

Fig. 4.4. From the left: rod on block, coin on plane, ball on plane.

Page 150: Introduction To Dynamics

142 4 Methods for Nonlinear Mechanics

If the rod is rocking on point L we have, using a relevant definition of the angle α ,

JLα =mgb

2. (4.19)

The moments of inertia are

J = JR = JL = JS +m(b2)2 =

ml2

12

(1+ 3(

bl)2)

, (4.20)

and combining these relations, we get the 1-DOF oscillator with a nonlinear restor-ing force according to (4.17)

Jα =−(

12

mgb

)sgn(α) . (4.21)

First, we consider (4.18):

S

R

α

g

b2

L

l

block

rod m, J

Fig. 4.5. Rod on block.

α =− rJ=−6bg

l2

(1

1+ 3(

bl

)2

). (4.22)

Formal integration results in

α =− rJ(t − t0)+C1 , (4.23)

α =−12

rJ(t − t0)

2 +C1 (t − t0)+C2 . (4.24)

The constant magnitudes C1 and C2 are determined by the initial conditions at timet = t0, at α|t0 = α0, and at α|t0 = 0. They give C1 = α0 and C2 = 0. Introducing thevelocity, eliminating the time, and separating the variables result in α dα

dα = α =− rJ

and from this

α2 =−2rJα+α2

0 . (4.25)

Page 151: Introduction To Dynamics

4.3 A 1-DOF Nonlinear Oscillator 143

Equation (4.25) represents a parabola

α =

√α2

0 − 2(rJ)α (4.26)

in the phase plane (α, α), the second half of which can be added by the correspond-ing solution of (4.19) (Section 4.3.1 and Fig. 4.6).

4.3.1 Piecewise Exact Solution

With respect to the piecewise constant feedback function r(x) = rsgn(x), there existonly two areas of motion, each of which is characterized by equations of motionwith constant coefficients. They are

mx =−r if x > 0 , (4.27)

mx =+r if x < 0 . (4.28)

For these simple equations an analytical solution by elementary integration is feasi-ble, and the two solutions must then be put together. For the case x > 0, we get

x =−(rm) , (4.29)

x =−(rm)t + x0 , (4.30)

x =−12(

rm)t2 + x0t + x0 . (4.31)

Elimination of time from the velocity solution and regarding x0 = 0 and t0 = 0 resultin

x = (m2r

)(x2

0 − x2) , (4.32)

which defines a parabola in the phase plane (x, x) (Fig. 4.6). The transient responsedescribed by (4.31) represents a parabola, too, with the characteristic magnitudes(Fig. 4.6)

t1 =2x0m

r, tm =

t12=

x0mr

, A = x(tm) =x2

0m

2r. (4.33)

At point t = t1, the sign of x becomes negative (sgn(x) < 0), and the exact solutioncontinues by the same parabola but mirrored. From this result, the period satisfies

T = 2t1 =4x0m

r. (4.34)

Expressing the initial velocity x0 by the amplitude A

Page 152: Introduction To Dynamics

144 4 Methods for Nonlinear Mechanics

0

x

tt1tm

A

x

x0

x

Fig. 4.6. Phase plane (x, x) and time behavior x(t).

x0 =

√2Arm

(4.35)

comes out with

T =4mr

√2Arm

=

√32mA

r= 5,66

√mAr

. (4.36)

The period T depends on the amplitude A, which is a characteristic feature of non-linear oscillations (Fig. 4.7). An experiment with a thrown coin will confirm thatimmediately (Fig. 4.4). The dependence of the period on the amplitude may be animportant indication for the assessment of vibrations.

T ∼√A

T

A

Fig. 4.7. Characteristic relationship T (A).

4.3.2 Method of Weighted Residuals

The basic idea of this method has already been presented in Section 2.3 consideringexamples of linear systems. In the following, we demonstrate its applicability withrespect to nonlinear systems. For this purpose, we approximate the solution x(t) ofa time-dependent nonlinear differential equation

D [x] = 0 (4.37)

Page 153: Introduction To Dynamics

4.3 A 1-DOF Nonlinear Oscillator 145

by the ansatz

xN(t) = aT w(t) (4.38)

with the constant coefficients ai and the trial functions wi(t). The unknown coeffi-cients ai are evaluated in such a way, that the weighted residual

∫wk(t)D

[aT w(t)

]dt = 0 (4.39)

becomes zero. The method is very general and can be used for time-dependent aswell as for space-dependent problems.

We again consider the one-dimensional example mx = −r(x) with a piecewiseconstant function r(x) = rsgn(x). Applying the ansatz

x(t) = Asin(ωt) (4.40)

and applying the method of the weighted residuals, we arrive at

∫ 2πω

0[mx+ rsgn(x)] sin(ωt)dt = 0 . (4.41)

Considering the periodicity of the problem, we integrate over one cycle T = 2πω only.

Now, we separate the integral into the two parts of the signum function resulting in(Fig. 4.8)

r sgn(x)

x

xr sgn(x)

t

t

t

r

Fig. 4.8. Properties of the signum function.

Page 154: Introduction To Dynamics

146 4 Methods for Nonlinear Mechanics

∫ πω

0

[−mAω2 sin(ωt)+ r]

sin(ωt)dt +∫ 2π

ω

πω

[−mAω2 sin(ωt)− r]

sin(ωt)dt = 0

(4.42)

and after performing the integration

−2[mAω2

( π2ω

)]+ 2

[2rω

]= 0 . (4.43)

From this, we get the eigenfrequency as

ω2 =4r

πmA, (4.44)

T =2πω

=

√π3mA

r= 5,57

√mAr

. (4.45)

It results in an error in comparison with T = 5,66√

mAr in (4.36) of the piecewise

exact solution of only 1,6%.

4.3.3 Harmonic Balance

What we call harmonic balance is sometimes also described by harmonic lineariza-tion. This approach replaces the original nonlinear differential equation by a lineardifferential equation, which approximates in the best way the motion over one pe-riod. Starting again with the standard example of the previous sections, this meansto replace the equation

mx+ rsgn(x) = 0 (4.46)

by the approximation

mx+ cx = 0, (4.47)

where in this case the term cx stands for the piecewise constant function r(x). It isan odd function.

r(x) =−r(−x) , (4.48)

r(0) = 0 . (4.49)

Therefore, we also use an odd ansatz

x(t) = Asin(ωt) (4.50)

and expand a FOURIER series for the odd function r(x)

Page 155: Introduction To Dynamics

4.3 A 1-DOF Nonlinear Oscillator 147

r(x) =∞

∑ν=1

aν sin(νωt) . (4.51)

Considering only the fundamental harmonic

cAsin(ωt) = cx = r(x)≈ a1 sin(ωt) , (4.52)

we get a relation for the spare coefficient c

c =a1

A. (4.53)

As r(x) is an odd function, the spare coefficient follows from

a1 = 2(2T)

∫ T2

0rsgn(Asin(ωt))sin(ωt)dt . (4.54)

In the interval under consideration (Fig. 4.8), the sign of Asin(ωt) is positive,

a1 =4rT

∫ T2

0sin (ωt)dt =− 2r

πcos(

2πT

)

∣∣∣∣T2

0=

4rπ

, (4.55)

and from this, we come to

c =a1

A=

4rπA

(4.56)

and to the linear differential equation approximating the original one

mx+

(4rπA

)x = 0 . (4.57)

The eigenfrequency is

ω2 =4r

πmA, (4.58)

T =2πω

=

√π3mA

r= 5,57

√mAr

. (4.59)

It is the same result as in the case of the weighted residuals, originating from thefact, that behind all these approximations there is a more or less hidden least squareidea.

4.3.4 Method of Least Squares

As before we want to generate a linear spare differential equation, but now by min-imizing the difference between the original and the spare differential equation

Page 156: Introduction To Dynamics

148 4 Methods for Nonlinear Mechanics

Δ = [mx+ r(x)]− [mx+ cx] = r(x)− cx (4.60)

in the sense of least squares integrated over one period

Δ2 =1T

∫ T

0[r(x)− cx]2 dt → min . (4.61)

A necessary condition for a minimum is

0 =

(∂ Δ2

∂c

)=

2T

∫ T

0[r(x)− cx]xdt , (4.62)

which results in

c =

∫ T0 r(x)xdt∫ T

0 x2dt. (4.63)

As above, we choose the ansatz x(t)

x(t) = Asin(ωt) (4.64)

and insert it into (4.63)

c =

∫ T2

0 rAsin(ωt)dt − ∫ TT2

rAsin(ωt)dt∫ T

0 A2 sin2 (ωt)dt=

4rπA

, (4.65)

which comes out with the same results as before, because of the arguments givenabove.

4.3.5 Practical Example

A highly illustrative example of some of the remarkable properties of nonlinear os-cillations generated by gear rattling has been detected in an application connectedwith noise problems of a pleasure boat [52]. The boat is equipped with a drive trainsystem combining a 340-hp eight-cylinder engine with a reversal gear, which inone direction operates as a one-stage gear system and in the other direction as atwo-stage gear system (Fig. 4.9). To switch between rotation and counter-rotation,a clutch is used which possesses backlash. In addition, all gear meshes also havebacklashes. Therefore, and dependent on the switched state, the system can rat-tle, the rattling process being excited by large engine unbalances. The rattling wasoccasionally so loud that the drive system could not be sold. Several parameterstudies were performed applying classical theories, but parameter variations onlymarginally improved the noise behavior. A breakthrough occurred as a result of anidea of the company’s design engineering department, which used a special clutchwith a maximum angular backlash of up to 35◦. The experiments indicated a noise

Page 157: Introduction To Dynamics

4.3 A 1-DOF Nonlinear Oscillator 149

engine

fly wheeladditional mass

clutch

counterrotation

rotation

Fig. 4.9. Scheme of the ship turning gear.

breakdown for an angular play of about 17◦. Simulations have confirmed this mag-nitude. Figure 4.10a displays the principal situation.

One can explain this strange behavior by considering a very simple 1-DOF modelwith backlash. The system is periodically excited and has a nonlinearity due to aspring with backlash that corresponds to the mesh of the gear teeth (Fig. 4.10b).

The equation of motion in dimensionless form is

ξ ′′+Dξ ′+ϕ(ξ ) = ϕ0 cosτ (4.66)

with the magnitudes ξ = xv , τ = ωt, ( · )′ = d

dτ , D = dmω , ϕ(ξ ) = F(x)

mω2v, ϕ0 =

F0mω2v

,and η = c

mω2 . The nonlinear force law F(x) is then

ϕ(ξ ) = η(ξ − 12) for ξ >+

12, (4.67)

ϕ(ξ ) = η(ξ +12) for ξ <−1

2, (4.68)

ϕ(ξ ) = 0 for − 12≤ ξ ≤+

12. (4.69)

We approximate the nonlinear equation (4.66) by the linear one

Page 158: Introduction To Dynamics

150 4 Methods for Nonlinear Mechanics

nois

e

backlash

(a) Noise over backlash.

vcFx

F0(ωt)x

m

d

(b) 1-DOF oscillator with backlash.

Fig. 4.10. Analysis of the ship turning gear.

ξ ′′+Dξ ′+η0ξ = ϕ0 cosτ (4.70)

and evaluate η0 by a least square method, taking into account (4.67)-(4.69). Thisresults in

η0 =

(2ηπξ0

)⎡⎣ξ0 arccos

(1

2ξ0

)− 1

2

√1−

(1

2ξ0

)2⎤⎦ , (4.71)

where ξ0 =x0v is the gain of the solution of (4.70).

Gain and phase ψ are

ξ0

ϕ0= mω2 x0

F0=

1√(1−η0)2 +D2

, (4.72)

tanψ =D

1−η0. (4.73)

Figure 4.11 depicts the resonance curves for this approximation. They reveal someastonishing properties. The diagram on the left of Fig. 4.11 illustrates the well-known behavior of resonance curves for systems with backlash. Scanning the playcharacteristic (4.67)-(4.69) from zero force to nonzero force results in a resonancestructure that is “more than linear” (resonance peak turns to the right); passing fromthe nonzero force branch to zero force we obtain a structure that is “less than linear”(resonance peak turns to the left). Both effects can be seen in Fig. 4.11. With in-creasing dimensionless play v

F0/c , the jump phenomenon is significantly intensified

when increasing or decreasing the excitation frequency Ω√c/m

.

The diagram on the right of Fig. 4.11 indicates a very strong jump behavior de-pending on the backlash itself. If the backlash v is of the order of magnitude of thespring deflection F0

c caused by the excitation force amplitude F0, we get a steep de-scent of the amplification factor x0

F0/c , where the character of this descent depends

Page 159: Introduction To Dynamics

4.4 Stability of Motion 151

Ω√c/m

x0F0/c

vF0/c

0,1 0,5 1 20

5

15

10

(a) Nonlinear resonance curves.

Ω√c/m

x0F0/c

vF0/c

0,1 1 100

5

15

10

(b) Jump behavior.

Fig. 4.11. Important parameter influences concerning rattling of a ship turning gear.

on the excitation frequency. With decreasing Ω√c/m

, we see a significant shift of the

amplification peaks to the right. The amplification x0F0/c is proportional to the noise

amplitudes. Therefore, the results of Fig. 4.11 give a physical interpretation of therattling phenomena as measured in the ship reversal gear. In addition, this could beconfirmed by a simulation with a complete model, which indicates that the neglec-tions in the simplified model are correctly estimated.

4.4 Stability of Motion

A confusingly large number of stability terms exist, terms like linear, nonlinear,static, dynamic, global, total, weak, strong, or asymptotic stability [38, 27, 33, 45].In the following, we focus our attention on

• stability of states of rest,• stability of orbits.

The concept of some norm plays an important role with respect to general stabilitydefinitions. For deviations from a reference state, however it may be defined, weneed some non-negative measure. The analysis of the bounds of such measures al-lows some insight into the global behavior of motion of a dynamic system and thussome statements with respect to its stability of motion. Some examples of norms arethe Euclidean norm , the weighted Euclidean norm and the arithmetic norm :

‖ x ‖2 =√∑x2

i =√

xT x , (4.74)

‖ x ‖2,R =√

xT Rx with R = RT > 0 , (4.75)

‖ x ‖1 =∑ | xi | . (4.76)

Norms have the following properties:

Page 160: Introduction To Dynamics

152 4 Methods for Nonlinear Mechanics

‖ x ‖ ≥ 0 and ‖ x ‖= 0 ⇒ x = 0 , (4.77)

‖ αx ‖=| α |‖ x ‖ for α ∈ IR , (4.78)

‖ x+ y ‖ ≤‖ x ‖+ ‖ y ‖ . (4.79)

4.4.1 General Stability Definitions

The splitting of the usually nonlinear equations of motion into a reference motionand into a perturbed motion was presented in Chapter 2. The reference motion itselfmight be a state of rest, but more frequently, it is some nonlinear type of motion,depending on the problem:

q(t) = q0(t)+ηq(t) . (4.80)

With the norm of the perturbation vector ηq(t)

ηq,2

ηq,1

asymptotically stable

stable

unstable

ε δ

q1

q2

ε

δt0

t

perturbedoriginal

Fig. 4.12. Illustration of stability definitions: ‖ ηq (t0) ‖< δ implies ‖ ηq(t) ‖< ε .

‖ ηq ‖=√

ηTqηq , (4.81)

we are able to generalize the stability definition of LAGRANGE and DIRICHLET inChapter 2 [36] (Fig. 4.12, left side):

Stability definition of LYAPUNOV:

• A unperturbed motion (state of rest) q0 of a dynamic system is calledstable, if there exists for every real constant ε > 0 another real constantδ (ε) > 0, so that from ‖ ηq (t0) ‖< δ , we always get ‖ ηq(t) ‖< ε for allt ≥ t0.

Page 161: Introduction To Dynamics

4.4 Stability of Motion 153

• The unperturbed motion (state of rest) is called asymptotically stable, if itis stable and in addition limt→∞ ‖ ηq(t) ‖= 0.

• The unperturbed motion (state of rest) is called stable in the limit, if it isstable but not asymptotically stable.

Example 4.2 (Pendulum). Observation tells us that a pendulum (Fig. 4.13) has twostates of rest, namely for ϕ = 0 and ϕ = ±π . It is also well known that positionϕ = 0 will be stable, and position ϕ = ±π will be unstable. A small perturbationof position ϕ = ±π will cause the pendulum to start moving. These two pendu-lum states possess some typical features, which we consider in the following. Theequations of motion write

ϕ+ω2 sinϕ = 0 with ω2 =mgl

J, (4.82)

with the moment of inertia J about the point of suspension. Eliminating time byϕ = ( dϕ

dϕ )ϕ gives us

ϕdϕ+ω2 sinϕdϕ = 0 (4.83)

and after integration the energy equation

12ϕ2 −ω2 cosϕ = c1 . (4.84)

Considering only small perturbations ηϕ around the two states of rest, we may per-form a TAYLOR-series expansion for (4.82) and (4.84):

• lower state of rest ϕ0 = 0, ϕ = ηϕ :

l

mg

Fig. 4.13. Pendulum.

Page 162: Introduction To Dynamics

154 4 Methods for Nonlinear Mechanics

ηϕ +ω2ηϕ = 0 , (4.85)

η2ϕ +ω2η2

ϕ = c2 . (4.86)

• upper state of rest ϕ0 =±π , ϕ =±π+ηϕ :

ηϕ −ω2ηϕ = 0 , (4.87)

η2ϕ −ω2η2

ϕ = c2 . (4.88)

The solution of the state ϕ0 = 0 describes an oscillation with constant amplitude

ηϕ = ηϕ0e jωt . (4.89)

According to LYAPUNOV such an oscillation is stable in the limit. Equation (4.86)describes an ellipse in the phase plane (ϕ , ϕ). Such points of equilibrium are said tobe elliptic points. They are always stable.

The solution for the case ϕ0 = ±π results in an oscillation with increasingamplitude

ηϕ = ηϕ0eωt . (4.90)

The stability conditions of LYAPUNOV are not satisfied. The pendulum is unstablefor that case. Equation (4.88) describes a hyperbola in the phase plane (ϕ , ϕ). Suchpoints are hyperbolic points. They are always unstable. Figure 4.14 displays thephase curves of all possible energy levels of the physical pendulum. It illustratesthe areas of the stable elliptic points ϕ0 = 2πν as well as the unstable hyperbolicpoints ϕ0 = ±(2ν+ 1)π . The curves being on the top of these influence areas rep-resent such a large energy that the pendulum starts to rotate. These two areas are

ϕ

π

ϕ

0

Fig. 4.14. Phase portrait for the pendulum.

Page 163: Introduction To Dynamics

4.4 Stability of Motion 155

separated by curves intersecting the unstable points ϕ0 =±(2ν+ 1)π . They form aseparatrix.

The definitions of LYAPUNOV as given above refer to states of rest, so to say to apoint in the phase space. If we want to refer these stability definitions to orbits, werecognize that the phase curves in the phase space change with time. Neighboringpoints stay close. Orbital stability requires that whole curves stay close withoutregarding specific points (right side of Fig. 4.12). We will come back to this point.See also the following example.

Example 4.3 (Orbital stability of a spherical pendulum.). We consider Fig. 4.15and Example 2.1. A spherical pendulum will be stable with respect to a small per-turbation of ϑ :

| ϑ −ϑ0 |< εϑ .

However, a εψ with | ψ −ψ0 |< εψ for all t ≥ t0 does not exist, because ψ0(t)represents a time-dependent reference motion. Anyway, the perturbed orbit will stayin the neighborhood of the reference curve, the pendulum will be orbitally stable.

m

ψ

Fig. 4.15. Spherical pendulum.

4.4.2 Linear Stability

LYAPUNOV’s first method judges the stability of a nonlinear system from the linearterms of a TAYLOR series expansion of the equations of motion. Therefore, it is of-ten called stability in the first approximation. LYAPUNOV formulated three stabilityconditions [10, 36].

We start with the nonlinear equations of motion in the state space form

x = f(x, t), (4.91)

Page 164: Introduction To Dynamics

156 4 Methods for Nonlinear Mechanics

and we assume that f(0, t) = 0 represents a state of rest. The term h(x, t) of theTAYLOR expansion

x = f(x, t) = f(0, t)︸ ︷︷ ︸=0

+A(t)x+h(x, t) (4.92)

is a vector-valued function, which does not include any linear components. Devel-oping the nonlinear function f(x, t) makes sense, because the stability of a referencemotion is usually perturbed by only small deviations. With the restriction to au-tonomous, time-invariant systems

x = Ax+h(x) , (4.93)

we formulate the following

Stability theorems of LYAPUNOV:

• If all eigenvalues λi of A have negative real parts, then the state of restx = 0 is asymptotically stable, independent of h(x).

• If at least one eigenvalue λk of A has a positive real part, then the state ofrest x = 0 is unstable, independent of h(x).

These two statements justify the limitation on investigations using only the linearterm of a TAYLOR approximation. They also hold for the case of multiple eigenval-ues. Some care is necessary, if the requirements for the above two theorems are notmet:

• If the matrix A has no eigenvalues with positive real parts but with van-ishing real parts, then the stability behavior is determined by the nonlinearterm h(x).

An important question in connection with these theorems is the relation withthe properties of the eigenvalues of A. It can be answered applying the methods ofSection 2.4. For the nonlinear case with vanishing real parts various methods areavailable, for example the center manifold reduction [24] or the general stabilitytheory of LYAPUNOV.

4.4.3 Stability of Nonlinear Systems

For critical cases like the third theorem of the preceding section or for a form ofthe equations of motion, which does not allow a decomposition into linear and non-linear parts, we have to determine stability only by considering the equations of

Page 165: Introduction To Dynamics

4.4 Stability of Motion 157

motion directly. The crucial point is, that we want to achieve stability informationwithout being forced to solve the equations of motion completely. LYAPUNOV con-sidered and solved this problem in the year 1892 by developing his second (direct)method for the stability of motion [27]. He introduced a test function V (x), whichpossesses some similarity with the energy of the system. This test function has tosatisfy certain conditions thus giving statements of stability or no stability. Today,the test function V is called the LYAPUNOV function.

Before considering the geometry of such a test function more closely, we remindof the stability theorems of LAGRANGE and DIRICHLET from Section 2.4. Thesetheorems connect stability of a state of rest with the minimum of the system’s po-tential energy. In many cases, but definitely not in all cases, it is sufficient to choosethe total energy for the LYAPUNOV function. A good choice is still a question ofinsight and experience.

To achieve a good feeling for the present method, we consider V (x) as an n-dimensional cup in state space x ∈ IR2 f . This cup must satisfy certain features inconnection with the equations of motion, to be able to deliver information on stabil-ity (Fig. 4.16). We restrict ourselves again to an autonomous system

x = f(x) (4.94)

V(x)

x1x2trajectory

(c increasing)

V =V +ΔV

(c decreasing)

ΔV ∗ = ΔV cot γ =

V = c

γ

n-dimensional cup

x = f(x) , x ∈ IR2 f

( ∂V∂x Δx+ ∂V

∂ t Δ t)cotγ

V =V −ΔV

Fig. 4.16. n-dimensional cup V (x) in state space.

Page 166: Introduction To Dynamics

158 4 Methods for Nonlinear Mechanics

with a state of rest f(0) = 0. A nonautonomous system x = f(x, t) would define adeformed and with time moving cup. The solution to the nonlinear equations ofmotion is an orbit in state space, which we do not know, and which we do not argueabout. The cup picture includes the property, that the contour lines (V = c) of thecup appear as closed curves on the cup surface. In addition, they might be projectedonto the state space, which is indicated as a plane in Fig. 4.16.

With this picture in mind we are able to define a few conditions for V (x) to be aLYAPUNOV function and the dynamic system to be stable:

1. V (x),(

∂V∂x

)are continuous at the origin.

2. V (0) = 0 is an isolated minimum.3. V (x)> 0 around the origin (V is positive definite).4. The derivation with respect to time along the solution trajectories x = f(x) satis-

fies (dVdt

)=

∂V∂x

dxdt

=∂V∂x

x =∂V∂x

f(x)≤ 0 . (4.95)

These features are immediately plausible in connection with the

stability propositions of LYAPUNOV:

• If we find a smooth positive definite function V (x) with ( dVdt ) ≤ 0 in the

neighborhood of 0, then 0 is stable.• If we find a smooth positive definite function V (x) with ( dV

dt ) < 0 in theneighborhood of 0, then 0 is asymptotically stable.

• If we find a smooth positive definite function V (x) with ( dVdt ) > 0 in the

neighborhood of 0, then 0 is unstable.

As already indicated, the contour lines of the cup V (x) are closed curves on thecup surface. Projecting these closed curves into the state space, which is displayedas a plane in Fig. 4.16, again results in closed curves around point x = 0. If in thestate space, the solution trajectories of x = f(x) run in such a way, that they cross theprojected contour curves from outside to inside, then they are going to the referencestate x = 0 and represent stable solutions (V < 0). If such a projected contour line isitself a solution trajectory, then the solution is stable, but not asymptotically stable.A solution trajectory going from inside to outside with (V > 0) represents an unsta-ble motion. This is the meaning of the above four conditions and the three stabilitypropositions. Applying these rules, a good choice of the LYAPUNOV function V ismandatory. There exist no clear rules for such a choice, but certain examples andexperiences:

1. We could choose the total energy and try to extend it by suitable terms.2. Sometimes, we may find some known integrals of the equations of motion and

use them for the LYAPUNOV function [48, 71].

Page 167: Introduction To Dynamics

4.4 Stability of Motion 159

• For conservative systems, we know the energy integral

T (q, q)+V (q) = E0 . (4.96)

• Again for conservative systems some momentum integrals may exist for thecase that qs does not appear in some terms for the energy (see Example 1.11).From the LAGRANGE’s equations of the second kind, we get for example

(∂T∂q

)qs

= 0 ,

(∂V∂q

)qs

= 0 (4.97)

and from this

ddt

(∂T∂ q

)qs

= 0 , (4.98)

which means (∂T∂ q

)qs

= pTs = const. (4.99)

The magnitude qs is said to be a cyclic coordinate and ps a cyclic momentum.It is

(∂T∂ q

)T

= p (4.100)

a generalized momentum. It is a constant for cyclic coordinates and wellsuited as a possible part of a LYAPUNOV function.

In spite of the difficulties involved in finding a good LYAPUNOV function, at least insome cases, it should be kept in mind, that LYAPUNOV’s methods are not only veryingenious, but also offer a quick possibility for a substantial stability analysis, evenfor complicated systems. We consider some simple examples in the following.

Example 4.4 (Nonlinear restoring forces and a critical case). We consider a mo-tion with cubic restoring forces:

x =

(x1

x2

)=

(0 1−1 0

)(x1

x2

)+ a

(x3

1x3

2

).

With respect to the state of rest x = 0, this is a critical case, because the linear termsalone result in a characteristic equation λ 2+1= 0, that is λ =± j with (ℜ(λ ) = 0).Again for the linear terms alone the energy budget writes

x21 + x2

2 = c2 .

Therefore, we select a test function

Page 168: Introduction To Dynamics

160 4 Methods for Nonlinear Mechanics

V (x) = x21 + x2

2

with the following properties:

• V (0) = 0.• V (x)> 0 around the origin (V is positive definite).• The derivation along the solution trajectory with respect to time gives

dVdt

=∂V∂x1

x1 +∂V∂x2

x2 = 2x1(x2 + ax3

1

)+ 2x2

(−x1 + ax32

)= 2a

(x4

1 + x42

).

These features make V a LYAPUNOV function for a ≤ 0:

a = 0 ⇒ dVdt

= 0 stable (stable at the limit) ,

a < 0 ⇒ dVdt

< 0 asymptotically stable (damping) .

For a > 0 the derivative satisfies dVdt > 0 and the state of rest is unstable (excitation).

Example 4.5 (Center). We consider again a 1-DOF-oscillator with a nonlinearrestoring force (4.12), which is assumed to behave near the origin r(0) = 0 likea strictly monotonic straight line. Nonlinear springs of some clutches behave likethat. According to (4.12), we get

x+r(x)m

= 0 .

Expressing the acceleration by velocity, eliminating time, and separating the vari-ables give

xdx+r(x)m

dx = 0 .

Integrating this equation with x =(x1 x2

)T=(x x

)T around x = 0 results in theenergy equation

V (x) =x2

2+

1m

∫ x

0r(ξ )dξ = c2 .

The first term corresponds to the kinetic, the second term to the potential energy.Choosing the energy V (x) as a test function satisfies the conditions for a LYAPUNOV

function, because we assume xr(x) > 0 for x �= 0 and r(0) = 0. It is

∂V∂x

x = x1∂V∂x1

+ x2∂V∂x2

= x21m

r (x1)− 1m

r (x1)x2 = 0 .

Page 169: Introduction To Dynamics

4.4 Stability of Motion 161

x

x

Fig. 4.17. Center point.

Therefore, the curves V (x) = c2 are closed curves, and the state of rest is stable inthe limit (Fig. 4.17). The point with respect to the state of rest x = 0 is also said tobe a center point.

Example 4.6 (Nondegenerated singularities for the planar case). A nonlinear 1-DOF-oscillator can be described in state space by

x = f(x) = f(0)+Ax+ o(‖ x ‖2)

with

x ∈ IR2 , A :=

(∂ f∂x

)0=

(a bc d

)∈ IR2,2 , det(A) �= 0 .

For a state of rest x = 0 (singular point, equilibrium point), we have f(0) = 0 andfrom this

x = Ax .

The corresponding eigenvalues are

λ1,2 =12

tr(A)± 12

√tr(A)2 − 4det(A) =

12(a+ d)± 1

2

√(a− d)2 + 4bc

with the eigenvectors

x1 =

(1

−(

a−λ1b

))

, x2 =

(1

−(

a−λ2b

))

.

From this, the general solution with arbitrary constants c1 and c2 writes

Page 170: Introduction To Dynamics

162 4 Methods for Nonlinear Mechanics

x(t) = c1x1eλ1t + c2x2eλ2t ,

x(t) = c1λ1x1eλ1t + c2λ2x2eλ2t .

According to the solution behavior around the state of rest, we describe a couple ofvariants by the various eigenvalue possibilities of the matrix A.

1. Saddle point (2-tangent node)λ1 and λ2 have opposite signs and are real, which means

det(A) = λ1λ2 = ad− bc < 0 ,

Δ = tr(A)2 − 4det(A) = (a− d)2 + 4bc > 0 .

One part of the solution goes in the direction of the equilibrium point, the otherpart goes the opposite way (Fig. 4.18).

x1

x2

Fig. 4.18. Saddle point.

2. Node (2-tangent node)λ1 and λ2 have the same signs and are real, which means

det(A) = λ1λ2 = ad− bc > 0 ,

Δ = tr(A)2 − 4det(A) = (a− d)2 + 4bc > 0 .

The solutions go in the direction of the equilibrium point or they go the oppositeway depending on the sign of tr(A) (Fig. 4.19).

3. Focus / vortexλ1 and λ2 are conjugate complex (and therefore also the eigenvectors x1 and x2),which means

Page 171: Introduction To Dynamics

4.4 Stability of Motion 163

x2 x1

Fig. 4.19. Stable node (tr(A)< 0).

det(A) = λ1λ2 = ad− bc > 0 ,

Δ = tr(A)2 − 4det(A) = (a− d)2 + 4bc < 0 .

We assume tr(A) �= 0. The solutions go to the origin or they go away from theorigin, according to the sign of tr(A) (Fig. 4.20).

Fig. 4.20. Unstable node (vortex).

4. Center pointλ1 and λ2 are purely imaginary, which means, that similar to the node

det(A) = λ1λ2 = ad− bc > 0 ,

Δ = tr(A)2 − 4det(A) = (a− d)2 + 4bc < 0 .

Because of tr(A) = 0, we get a center (vortex) comparable to Example 4.6(Fig. 4.21).

Page 172: Introduction To Dynamics

164 4 Methods for Nonlinear Mechanics

Fig. 4.21. Center (vortex).

tr(A)

det(A)

Δ = 0

Δ > 0 Δ > 0

Δ < 0

sinks sources

x1

Fig. 4.22. Combined diagram (Δ = tr(A)2 −4det(A)).

It is possible to combine all these cases in one diagram (Fig. 4.22). We see that forcases 1, 2, and 3, the nonlinear system behaves approximately like a linear systemin the neighborhood of the singularity. This is not true for case 4. The nonlinearsystem of case 4 may also have a node (vortex), which finally is decided by termsof higher order in the Taylor series of f or by the general method of LYAPUNOV

according to Example 4.6. Such cases are called degenerated singularities or sin-gularities of higher order. The same situation might occur if at least one of theeigenvalues λ1 or λ2 is a null-eigenvalue or if the matrix A cannot be diagonal-ized. The above-mentioned classification of singularities was already established byPOINCARÉ more than a hundred years ago [24], whereas these degenerated singu-larities are still a matter of research today, especially with respect to singularities inmultidimensional systems.

Page 173: Introduction To Dynamics

Chapter 5Vibration Phenomena

5.1 Introduction

As we have seen in previous chapters, purely periodic oscillations can be character-ized by a periodicity condition

x(t) = x(t +T ) (5.1)

with the period T and the frequency

f =1T

(5.2)

or the angular frequency

ω = 2π f = 2πT

. (5.3)

The period need not be constant, but may depend on the amplitude (Fig. 4.7). In thefollowing, we deal with vibrations with one or more constant periods and investigatethe issue of vibration formation. In order to recognize the essentials, we restrictourselves in many cases to oscillators with one degree of freedom x(t).

Oscillations are characterized by periodically recurring states (x, x), where theintensity of these states may change depending on the energy in the system. This isreflected by the oscillation amplitude, which may be damped by dissipation, or alsobe excited. The recurring states can be realized by different physical mechanisms.

The simplest cases are free and forced oscillations as described in Chapters 2and 3. Characteristic of these oscillations are the eigenfrequencies f or the excita-tion frequencies Ω . While modes are created by a single external impulse, forcedvibrations require an (usually periodic) external excitation. Forced oscillations arethe cause behind most problems in machine dynamics.

As long as vibrations can be described by linear differential equations with con-stant coefficients, a more or less closed mathematical tool box to solve the problemsis available [1, 46]. However, some important oscillation phenomena are highly

© Springer-Verlag Berlin Heidelberg 2015 165F. Pfeiffer and T. Schindler, Introduction to Dynamics,DOI: 10.1007/978-3-662-46721-3_5

Page 174: Introduction To Dynamics

166 5 Vibration Phenomena

nonlinear. Their mathematical treatment is far more difficult. These include self-excited and parametrically excited oscillations. Self-excited vibrations are createdas a dynamic equilibrium state of power supply into the system and energy con-sumption within the system, where the energy is supplied at the frequency of thevibrations. Examples are flutter, friction oscillators, and the pendulum clock. Forparametrically excited oscillations, the periodic changes of one or more parametersof the system work as a kind of internal excitation, for example, the periodicallyvariable tooth stiffness due to the time-varying tooth contact in a gear [51, 42].

The frequencies of modes and self-excited oscillations are real eigenfrequencies,because they are determined by the vibration system itself (autonomous systems). Incontrast, the frequencies of forced and parametrically excited vibrations depend onexternal excitations or design-related time-varying processes (external excitation,heteronomous systems). Table 5.1 gives an overview of the oscillation phenomenato be treated. The classification of the vibrations according to formation principlesis certainly useful, but not the only possibility. Vibrations can also be distinguishedaccording to their properties, in linear and nonlinear oscillations. A third aspectcould be the degree of freedom, that is the antiquated distinction between mono-track and multiple-track oscillators (coupled vibrations). Although most machinesrepresent oscillators with many degrees of freedom, the restriction to one degree offreedom makes sense, because its characteristic behavior can also be found in largesystems.

Table 5.1. Classification of vibrations according to formation principles.

vibration type examples origin frequency equation of motion

pendulum, single eigen angular homogeneousfree tuning fork, external frequency ω x+ω2x = 0vibrations piano string impulse

modesfoundation external forces

forced vibration, or moments, excitation inhomogeneousvibrations vibrating screen, usually acting frequency Ω x+ω2x = F cos(Ω t)

vehicles on periodicallyrough groundclock, bell, self-control

self-excited string and wind with not about eigen nonlinearvibrations instruments, periodically angular frequency x+ f (x, x) = 0

flutter, acting ωwoodpecker toy energy sourcepiston engines, fractions or

parameter-excited propeller, periodically multiples of the periodicvibrations gear drive, changing parameter coefficients

pendulum with moving parameters frequency ωP x+ p(t)x = 0point of suspension

Page 175: Introduction To Dynamics

5.2 Free Vibrations 167

5.2 Free Vibrations

Free vibrations can be characterized by their eigenfrequencies, their eigenvectors ormodes, their damping behavior and energy balance. For a conservative system, thereis a periodic exchange of potential and kinetic energy (pendulum in Fig. 5.1):

T +V = E0 = const. (5.4)

x

m

mg

y

ϑ l

Fig. 5.1. Pendulum.

Modes may occur in linear and nonlinear systems. For a small initial deflection, thependulum in Fig. 5.1 is an example of a linear mode:

(ml2) ϑ =−mglϑ . (5.5)

With ω2 = gl , this yields

ϑ +ω2ϑ = 0 . (5.6)

The solution for this system is

ϑ(t) = ϑ0 cos(ωt)+ϑ0

ωsin(ωt) . (5.7)

With (5.6), introducing the velocity, eliminating the time, and separating the vari-ables gives

ϑdϑ +ω2ϑdϑ = 0 . (5.8)

In the phase portrait(ϑ , ϑ

), we obtain the ellipses

ϑ 2 +

(ϑω

)2

= ϑ 20 +

(ϑ0

ω

)2

. (5.9)

Page 176: Introduction To Dynamics

168 5 Vibration Phenomena

If we also allow large deflections ϑ , we obtain the equation of motion

ϑ +ω2 sinϑ = 0 (5.10)

and therefore the phase curves

(ϑω

)2

− 2cosϑ =

(ϑ0

ω

)2

− 2cosϑ0 , (5.11)

which degenerate to (5.9) for small ϑ . The corresponding phase portrait is shownin Fig. 5.2, from which we get the periodic change of the potential energy V (Ex-ample 4.2). Another example of a nonlinear mode is given by the rod on a block inFig. 4.12. Phase portrait and amplitude response are shown in Fig. 5.3.

For vibration processes with energy dissipation, the eigenfrequencies changeonly slightly, but the amplitude changes significantly. Considering the single degreeof freedom oscillator in Fig. 5.4, the following equations can be stated:

mx+ dx+ cx = F(t) = 0 , (5.12)

or with the eigen angular frequency of the undamped system ω20 = c

m and the LEHR

damping D = d2mω0

= d2√

cm :

V

ϑ−π π

ϑ

ϑ

0

−π0

π

Fig. 5.2. Phase portrait of the pendulum.

Page 177: Introduction To Dynamics

5.2 Free Vibrations 169

ω

ω ∼ 1√A

A

x

x

Fig. 5.3. Rod on a block – phase portrait and amplitude response (see Figs. 4.5 and 4.6).

F(t) F(t)

mm

c d cx(t) dx(t)

x(t)

Fig. 5.4. Single degree of freedom oscillator.

x+ 2Dω0x+ω20 x = 0 . (5.13)

With the ansatz x = xeλ t , we obtain the eigenvalues

λ1,2 =−Dω0 ± jω0

√1−D2 . (5.14)

The classification of the vibration types is thus given as in Fig. 2.5a-2.5d. Dependingon the value of D, five cases occur.

x

t

(a) Aperiodic case.

x

t

(b) Damped vibration.

t

x

(c) Periodic vibration.

Fig. 5.5. Stable oscillation types.

Page 178: Introduction To Dynamics

170 5 Vibration Phenomena

1. Aperiodic case, no oscillation (D > 1, Fig. 5.5a)

λ1,2 = ω0

(−D±

√D2 − 1

). (5.15)

2. Damped vibration (0 < D < 1, Fig. 5.5b)

λ1,2 = ω0

(−D± j

√1−D2

). (5.16)

3. Periodic vibration, limiting case (D = 0, Fig. 5.5c)

λ1,2 =± jω0 . (5.17)

4. Excited vibration (−1 < D < 0, Fig. 5.6a)

λ1,2 = ω0

(−D± j

√1−D2

). (5.18)

5. Unstable aperiodic case (D <−1, Fig. 5.6b)

λ1,2 = ω0

(−D±

√D2 − 1

). (5.19)

x

t

(a) Excited vibration.

x

t

(b) Unstable aperiodic case.

Fig. 5.6. Unstable oscillation types.

5.3 Forced Vibrations

Forced vibrations are characterized by the eigen behavior of the vibration system, bythe type of excitation, and the resulting output in the form of amplitude- and phasefrequency response functions [42, 41]. We consider an oscillator with one degree offreedom as in (5.12) with right-hand side f (t) �= 0

mx+ dx+ cx = f (t) = cAcos(Ω t) (5.20)

Page 179: Introduction To Dynamics

5.3 Forced Vibrations 171

resulting from a harmonic kinematic excitation Acos(Ω t). With the eigen angularfrequency of the undamped system ω0 and the LEHR damping, we obtain the differ-ential equation:

x+ 2Dω0x+ω20 x =

cAm

cos(Ω t) . (5.21)

The solution consists of a homogeneous and an inhomogeneous part. The homoge-neous component corresponds formally to the solution of (5.13) and physically tothe modes. If we assume a damped system, the modes initiated only once decreaseand disappear after a while. In contrast, the continuously excited oscillations per-sist, because the periodic excitation supplies power and forces the vibration systemto oscillate with the same frequency but different amplitude and phase.

According to Section 2.3.1, the stationary solution results from the frequencyresponse function (ω2

0 = cm , D = d

2mω0= d

2√

cm , η = Ωω0

)

G( jΩ) =c

m(−Ω 2 + jΩ2Dω0 +ω2

0

) =1

1+ j2Dη−η2 . (5.22)

Amplitude amplification and phase shift satisfy

V = |G( jΩ)|= 1√(1−η2)2 + 4D2η2

, (5.23)

ψ (G( jΩ)) = arctan

(2Dη

1−η2

), (5.24)

such that the particular solution is

x = AV cos(Ω t −ψ) . (5.25)

The ansatz of the type of the right-hand side is based on the physically reasonableassumption that the input Acos(Ω t) creates an output, which has a phase shift ψand an amplitude changed by the amplification factor V . The maximum value of Vis derived from ∂V

∂η = 0:

ηmax =√

1− 2D2 . (5.26)

Plotting V and ψ over η , we obtain the amplitude frequency response function inthe first case (Fig. 5.7) and the phase frequency response function (Fig. 5.8) in thesecond case.

Such diagrams are indispensable tools for the estimation of vibrations, especiallyin systems with many degrees of freedom. We can calculate and measure the am-plitude and phase frequency response functions. The evaluation of complex systemsrequires carefully established models with clear relation to practice. Measurements

Page 180: Introduction To Dynamics

172 5 Vibration Phenomena

V

2

2

1

10

0 η

D = 0

0,25

0,50,75

1,0

2,0

Fig. 5.7. Amplitude frequency response function (5.23).

η

ψ

1 200◦

90◦

180◦D = 0

0,25

0,5

0,752,0 1,0

Fig. 5.8. Phase frequency response function (5.24).

demand care and a clever (meaningful) arrangement of the measuring points. In eachcase, the curves inform on

• resonance points,• damping behavior,• phase of the degrees of freedom with respect to each other,• assessment of the resonances,• impact of construction parameter variations on position and amplitude of reso-

nances.

Page 181: Introduction To Dynamics

5.3 Forced Vibrations 173

We discuss forced oscillations with the actual example of a ship propulsion. Thefive-bladed propeller is driven by a gas turbine via a planetary gear and via a spurgear (Fig. 5.9). The most important external excitation is caused by the propeller. Ittakes effect with five times the propeller speed (= output speed), since each propellerblade produces a perturbation of the mean propeller thrust due to the hydrodynamicprocesses at the ship’s stern. Although the axial thrust bearing absorbs the totalthrust and the majority of the thrust variations, the rotational nonuniformity of thepropeller shaft as well as a part of the axial thrust variations remain and negativelyinfluence the entire drive train.

spur gear

planetary gear

clutches

gas turbine

1xpr

op.

2xpr

opel

ler

ring

2xpi

nion

sun 4x

ring

5xri

ng

ampl

itud

e

frequency

Fig. 5.9. Ship propulsion, measuring point located at the planetary gear.

Page 182: Introduction To Dynamics

174 5 Vibration Phenomena

In addition to the external excitations, there are parametric excitations in thetoothing of the two gears. As there is never the same number of teeth in contact andthe toothing effects the overall dynamics of the system as an elastic coupling, weobtain a time-varying, that is not constant, stiffness characteristic in such couplingpoints. The deviation from the constant mean depends on the degree of coverageof the gears (Section 5.5). The time-varying elastic couplings generate parametricexcitations in the system that lead to additional resonances. These parametric ex-citations are usually locally limited to the environment of their origin, especiallyin large powertrains [35]. Since the spectrum in Fig. 5.9 has been recorded at theplanetary gear, we can only observe the influence of parametric excitations at theplanetary gear and not those at the spur gear.

The influence of the propeller excitation occurs only in the lower frequencyrange. At higher frequencies, the parametrically excited vibrations and the modesof all the components of the planetary gear dominate. These can be more dangerouswith respect to the amplitude increase than the lower frequency resonance (Sec-tion 5.5). The amplitude frequency response function (spectrum) for the main har-monic components (FOURIER coefficients) of a vibration system does not containdirect information on the energy of the resonances. This topic is addressed in moredetail in [42, 43].

spur gear

freewheel

planetring

sun

gas turbine

double tooth coupling

thrust bearingpropeller

P(t) MAb

Fig. 5.10. Model of the ship propulsion (Fig. 5.9).

The analysis of a spectrum as in Fig. 5.9 may be difficult. Often, we have tofind a mechanical vibration model. Fig. 5.10 shows such a model for the drivetraindepicted in Fig. 5.9. Basically, it serves to represent the torsional vibrations of theoverall system and some more detailed phenomena in the gears. All coupling pointssuch as bearings and toothings are modelled as spring-damper elements, wherebywe have to take into account the time-varying stiffness and damping in the toothings(Section 5.5). For such a mechanical model, the equations of motion are derivedaccording to the methods of Chapter 1. After that, simulations can be undertaken

Page 183: Introduction To Dynamics

5.4 Self-excited Vibrations 175

with the mathematical model. In the present case, the entire drivetrain system has38 degrees of freedom. The measured frequencies in Fig. 5.9 could be identified andproven without exception.

5.4 Self-excited Vibrations

Self-excited vibrations represent a very special and fascinating class of natural os-cillations. They take energy from an energy source at the frequency of the vibrationsand cover energy losses in the vibration system with this energy supply. If there isa balance of supply and loss, a stable periodic oscillation occurs that can be repre-sented by a limit cycle in the phase diagram. Self-excited oscillators need an energysource and a kind of switch, which can enable or disable the supply of energy fromthe source. Since the oscillation is periodic in its stationary state, the switch mustbe involved in the periodic process of the oscillator. For self-excited oscillators, wedistinguish two types, that is oscillator type and storage type. A typical example ofthe oscillator type is the electric bell (Fig. 5.11). It takes electric energy from thesource grid and thus causes an electromagnet to attract the clapper and hit the bell.This process interrupts the power supply such that the clapper swings back into itsoriginal position due to the deformation energy stored in the retaining spring. Then,the process begins again. Switch, energy storage, and oscillator is the clapper withits elastic attachment, while the energy source is the grid [42]. Other examples ofoscillator-type self-excited oscillators are the pendulum clock, the violin string, flut-ter, wind instruments, and the KÁRMÁN vortex street. The violin string stands for awhole class of self-excited oscillators, the friction oscillators.

Self-excited oscillators of storage type are often relaxation oscillators on a hy-draulic basis. A vessel is filled and emptied at a certain filling level through a tube(Fig. 5.12), or mechanical tilting occurs.

For the basic understanding of self-excited vibrations, we have to analyze theenergy balance. Linear and nonlinear conservative oscillators have modes that are

U

energy source oscillator switch

Fig. 5.11. Self-excited oscillator: oscillator type, example bell.

Page 184: Introduction To Dynamics

176 5 Vibration Phenomena

outflow

energy source storage switch

Fig. 5.12. Self-excited oscillator: storage type

characterized by a periodic exchange of potential and kinetic energy (5.4). The keyenergy magnitudes are T and V . In contrast, for self-excited vibrations the motionis determined by the supplied energy ΔEZ and by the dissipated energy ΔED in thesystem. During a period, we have to supply the same amount of energy (+ΔEZ) asis dissipated in the system (−ΔED). The following cases are possible:

• ΔEZ < ΔED (damping):More energy is dissipated than supplied. The amplitude decreases.

• ΔEZ > ΔED (excitation):More energy is supplied than dissipated. The amplitude increases.

• ΔEZ = ΔED (limit cycle):Just as much energy is supplied as is dissipated. This is the limiting case ofperiodic motion.

To get a rough quantitative estimate of these ratios, we assume a periodic oscillationand a linear damping law:

x ≈ xcos(ωt) , (5.27)

FD ≈−dx . (5.28)

Then for an oscillator with one degree of freedom, we get the dissipated energy

ΔED ≈−∫ T

0FDxdt = d

∫ T

0x2dt = πdωA2 ∼ A2 . (5.29)

For simplicity, we consider an oscillator for which the supplied energy ΔEZ is con-stant. Then, we obtain the conditions of Fig. 5.13, which are typical for self-excited

Page 185: Introduction To Dynamics

5.4 Self-excited Vibrations 177

E

ΔEZ

ΔED

ΔEZ < ΔED

ΔEZ = ΔED

ΔEZ > ΔEDA

A1

x

x

ΔED = ΔEZ A1

periodic limiting case

damping

excitation

limit cycle

Fig. 5.13. Energy balance.

oscillators. The point with the same amount of supplied and of dissipated energydefines a stable limit cycle in the illustrated case, which separates the two regionsof excitation and damping. In the phase portrait, this means that the vibrations ap-proach the limit cycle from outside and from inside and remain stable on the limitcycle. There are also unstable limit cycles, for which the regions of excitation anddamping are reversed, that is the damping region is below and the excitation regionis above the limit cycle. We can observe such a case for the pendulum clock, forwhich the basic conditions are sketched in Fig. 5.14. The power supply is stronglyamplitude-dependent, causing a second unstable limit cycle [33, 42]. If the pendu-lum is initiated only very weakly, the pendulum clock mechanism (for example, theGRAHAM mechanism in Section 5.4.7) does not start and the pendulum gets to restdue to internal friction. With moderate initiation, the pendulum clock mechanismis started; it drives the pendulum amplitudes to the outer stable limit cycle, whichcorresponds to the continuous operation of the clock.

5.4.1 Hydraulic Oscillator

Hydraulic oscillators belong to the storage type. The basic structure of such a systemis shown in Fig. 5.12. For the hydraulic oscillator shown in Fig. 5.15, the storageis filled by a continuous stream of water. At the water level height of h2, a leverattached in the storage, that is a switch acting on the sink, becomes active and en-sures that the storage is emptied to the height h1. Because of the air entering thesystem, the emptying is interrupted. Subsequently, it once again begins to fill. The

Page 186: Introduction To Dynamics

178 5 Vibration Phenomena

E

ΔEZ

ΔED

AA2

x

x

AD

D

A1

A = excitationD = damping

unstable

stable

Fig. 5.14. Principle of the pendulum clock.

ener

gyso

urce

inlet

switchstorage

outlet

h 1h 2

Fig. 5.15. Hydraulic oscillator.

motion consists of a repetitive swinging of the water height h between the two limit-ing heights h1 and h2, where the oscillation time T is the sum of filling time TF andemptying time TE . The time behavior of the system is shown in Fig 5.16a, while thelimit cycle of the phase portrait belonging to the steady state is shown in Fig. 5.16b.

Page 187: Introduction To Dynamics

5.4 Self-excited Vibrations 179

h

th1

h2

TF TE

(a) time behavior

h

hh1 h2

(b) limit cycle

Fig. 5.16. Behavior of the hydraulic oscillator.

5.4.2 Drinking Bird

The well-known drinking bird toy (Fig. 5.17) is a self-excited system of storage-oscillator type. Apart from the energy source (environment) and the switch (riserpipe), both storage (head and body) and oscillator (pendulum arrangement) are nec-essary such that the thermal-mechanical system can perform self-excited oscilla-tions. An absorbent layer fitted to the head of the drinking bird, which can soakwater through the pecker, ensures that the head remains wet and cool. The neck isdesigned as a riser pipe, it extends into the body and is surrounded by ether or an-other low-boiling liquid. Both in the body and in the head, the steam of the ether isabove the liquid. The vapor pressure p1 in the body equals approximately the one ofthe environment. The pressure p2 in the head is lower because the head temperatureis about 0,3◦ below the room temperature due to the water evaporation. Because ofthe pressure difference, the liquid level in the riser pipe increases (height h). Thecenter of gravity S, which was initially below the pivot point D, shifts upwards andthe drinking bird tilts forwards slowly. In the nearly horizontal position, the peckerdips into the water. The amount of liquid in the interior of the bird is defined, such

S

p1h

p2

D S

D

Fig. 5.17. Drinking bird: schematic arrangement with D–pivot point, S–center of gravity andarrangement before the drinking bird sits up.

Page 188: Introduction To Dynamics

180 5 Vibration Phenomena

that now the lower end of the riser pipe is released (Fig. 5.17). The liquid quicklyflows back from the head to the body, the center of gravity S moves down, the drink-ing bird sits up again and swings back and forth for some time. These oscillationsstimulate the evaporation at the head and the heat transfer from the environment tothe body. Slowly, the liquid level in the riser pipe increases, the drinking bird tilts,and the game starts again.

5.4.3 Woodpecker Toy

An interesting example of a nonlinear oscillator with self-excitation is the wood-pecker toy, which represents a nonsmooth dynamic system [53]. The toy consists ofa rod, on which the swinging woodpecker slides down. In particular, it consists ofa sleeve, which slides on the rod with clearance, and of the woodpecker, which isconnected to the sleeve by a spring (Fig. 5.18a). The most important part of the toyis the sleeve with clearance, which induces a self-locking of the sleeve at specifictilting angles ±ϑk1. The kinetic energy of the downward movement is built-up fromgravity g due to sliding along the rod; at the lower self-locking position ϑ = +ϑk1(Fig. 5.18a), it is converted into vibrational energy for the woodpecker after subtrac-tion of shock losses, because the z-degree of freedom of the downward movementis suddenly stopped. The woodpecker is now swinging to a maximum deflectionand back. Reaching the tilting angle +ϑk1 again, the self-locking and therefore thez-degree of freedom is released. Under the influence of gravity, slipping occurs butthe woodpecker swings upwards

(ϑ < 0

). At ϑ = −ϑk1, self-locking blocks the

z-degree of freedom. For a specific ϑ = −ϑk2 (| ϑk2 |>| ϑk1 |), the woodpecker isforced into a rapid turnaround due to the pecker impact. Thus, at ϑ = −ϑk1, it isreleased from self-locking and moves down. Now, it swings downwards

(ϑ > 0

)and enters self-locking at ϑ =+ϑk1. The cycle then starts again.

y

z

m1

m2

I1

I2

ab

S2cϑ

ϑk1

S1

(a) schematic arrangement

pecker impact lower sleeve impact

upper sleeve impact

ϑk2

+ϑk1

−ϑk1

ϑmax

ϑ

ϑ

0

12

34

5

6 7

(b) schematic limit cycle

Fig. 5.18. Woodpecker toy.

Page 189: Introduction To Dynamics

5.4 Self-excited Vibrations 181

The described motion is a nonlinear, self-excited vibration of storage-oscillatortype with a stable limit cycle. The main switch is the sleeve with clearance, whichcontrols the transformation of translational energy into vibrational energy as wellas the build-up of translational energy from gravitation via self-locking. The peckerof the woodpecker is a further switch, which induces a fast reversion of the oscil-lation by the pecker impact. However, the pecker impact is not necessary for thefunctionality of the toy. The self-excitation mechanism is illustrated in Fig. 5.19.

energy source

gravitation

storage

mass

oscillator

woodpecker

switch

peckerϑ = ϑk2

switch

sleeveϑ =±ϑk1

turnaround ϑ

coupling z,ϑ

energy transfer due to impact

switch on/off, self-locking, friction

ΔV=mgΔz

Fig. 5.19. Self-excitation mechanism.

The oscillation consists of five phases, which are connected with or without im-pacting behavior due to constraints. The phases themselves are described by dif-ferential equations for one or two degrees of freedom. For the connections withimpacting self-locking, it is z = 0 (Fig. 5.18a). The five phases of motion can besummarized as shown in Table 5.2 (Fig. 5.18b). A solution of the equations of mo-tion connecting the different phases and using the periodicity results in the three-dimensional limit cycles according to Fig. 5.20.

The woodpecker toy is defined by b = 0,015m, a = 0,025m, m1 = 0,0003kg,m2 = 0,0045kg, I1 = 5 · 10−9 kgm2, I2 = 7 · 10−7 kgm2 as well as c = 0,0056Nm.The energy absorbed within the lower sleeve impact is about 79% of the supplied en-ergy. The difference between theory and measurement of the limit-cycle frequencyand of the falling height is only a few percent [53].

5.4.4 Friction Oscillator

Oscillation systems with dry friction (COULOMB friction) can be found in manytechnical applications, whereby the friction is often the source of the self-excitedoscillations. The formation principle is always the same, such that we can discussthe principle with the example of the friction pendulum (FROUDE pendulum).

The friction pendulum consists of an engine, where the shaft rotates with constantangular velocity ϕw (Fig. 5.21). A pendulum is mounted on the shaft such that it can

Page 190: Introduction To Dynamics

182 5 Vibration Phenomena

Table 5.2. Phases of motion.

Fig. 5.18b degree of freedom(ϑ , ϑ

)– begin

(ϑ , ϑ

)– end translation z

0 - 1 ϑ ϑ =+ϑk1 ϑ =+ϑk1 z = 0ϑ > 0 ϑ < 0

1 - 2 ϑ ,z ϑ =+ϑk1 ϑ =−ϑk1 z > 0ϑ < 0 ϑ < 0

3 - 4 ϑ ϑ =−ϑk1 ϑ =−ϑk2 z = 0ϑ < 0 ϑ < 0

5 - 6 ϑ ϑ =−ϑk2 ϑ =−ϑk1 z = 0ϑ > 0 ϑ > 0

6 - 7 ϑ ,z ϑ =−ϑk1 ϑ =+ϑk1 z > 0ϑ > 0 ϑ > 0

−4

4

8

−8

0

−0.10.2

0.4

0.1 0.2 −4

4

8

−8

0

−0.10.2

0.4

0.1 0.2 −4

4

−8

0

−0.10.2

0.4

0.1 0.2

ϑ

ϑ

z

ϑ

z

ϑ

point 0 point 0 point 0

a b c

ϑ

ϑ

z

Fig. 5.20. Results: a reference limit cycle, b approaching from outside, c approaching frominside [53].

rotate freely but it is taken due to the friction forces in-between the pendulum sleeveand the shaft. This effect is limited by the point, at which the pendulum torque canno longer compensate the gravitational torque due to sticking forces.

The friction pendulum is thus performing the following motion patterns. Forsmall sticking forces, the shaft will drive the pendulum up to a reversion point,where an equilibrium of the sticking forces, of the velocity-dependent friction forcesat the shaft, and of the gravitational forces breaks down. Then, the pendulum movesagainst the shaft rotation until another reversion appears, where the force balanceagain allows a driving effect. The other extreme, namely a large friction force,results in a pendulum rotation with the shaft. Starting from these two basic mo-tion patterns, an intermediate behavior is also possible which we do not discusshere [33, 42].

Page 191: Introduction To Dynamics

5.4 Self-excited Vibrations 183

ϕW

ϕ

Fig. 5.21. Friction pendulum.

In the following, we consider the motion of the pendulum with low sticking fric-tion. The characteristics of the dry friction decrease (Fig. 5.22). This allows self-excited oscillations to develop, because around the point of constant shaft velocityϕw, there is a larger friction torque for relative angular velocities 0 < ϕr < ϕw thanfor relative angular velocities ϕr > ϕw. This results in asymmetric oscillations. Ifthe pendulum in Fig. 5.21 moves in the rotation direction of the shaft with ϕ > 0,then the relative rotation satisfies ϕr < ϕw because of ϕr = ϕw − ϕ. According toFig. 5.22, a larger friction torque is introduced than for the case ϕ < 0 and ϕr > ϕw,that is energy is supplied to the pendulum by the friction torque.

The equation of motion for the friction pendulum in Fig. 5.21 is given by

Jϕ+ dϕ+mgssinϕ = R (5.30)

with the following parameters: J pendulum moment of inertia, d velocity propor-tional damping, m pendulum mass, g gravitational constant, s distance of the centerof gravity of the pendulum to its suspension point, R friction torque, and ϕ angle.The friction torque depends on the relative velocity ϕr:

R = R(ϕr) = R(ϕw − ϕ) . (5.31)

With the dimensionless quantities

R

+R0

−R0

ϕW

ϕr

Fig. 5.22. Characteristics of the dry friction.

Page 192: Introduction To Dynamics

184 5 Vibration Phenomena

dJ= 2δ ,

mgsJ

= ω20 ,

RJ= r , (5.32)

one obtains

ϕ+ 2δ ϕ+ω20 sinϕ = r (ϕw − ϕ) , (5.33)

which yields, by inserting the velocity and eliminating the time,

ϕdϕdϕ

= r (ϕw − ϕ)− 2δ ϕ−ω20 sinϕ . (5.34)

With a known friction function r, the phase portrait can be derived. First, we recog-nize an equilibrium position with ϕ = 0

r (ϕw)−ω20 sinϕ = 0 , (5.35)

hence

sinϕ0 =r (ϕw)

ω20

=R(ϕw)

mgs. (5.36)

Physically, this means that the pendulum adjusts to a value ϕ0 = ϕw > 0 due tothe friction between the pendulum sleeve and the motor shaft; this corresponds to abalance between friction torque R and gravitational torque (mgs).

If at larger pendulum amplitudes, the pendulum velocity equals the velocity ofthe shaft, that is ϕ = ϕw and ϕr = 0, this corresponds to a state of motion withmaximum friction R0 = R(0) or r0 = r(0). This state of static friction or stictionmay be left, if the damping and restoring forces just reach the value r0. Then, thependulum is released again from the shaft and starts to oscillate downwards. Such abreakpoint can be calculated using

r0 − 2δ ϕw−ω20 sinϕ1 = 0 , (5.37)

which results in

sinϕ1 =r0 − 2δ ϕw

ω20

. (5.38)

In the phase portrait, point ϕ1 is on the jump line ϕr = 0. For this value, the staticfriction jumps from −R0 to +R0 (Fig. 5.22). Each motion in the vicinity of the jumpline leads into it (Fig. 5.23). If the pendulum velocity ϕ is smaller than ϕw and ϕis increasing, the difference ϕw − ϕ will decrease, the friction torque will increase(Fig. 5.22), and the motion will tend to ϕ = ϕw. However, an overshoot with ϕ > ϕw

results in a negative friction torque opposite the direction of shaft rotation, which,first, cannot be compensated by the gravitational torque, and which, second, ensuresthat the motion stays on the jump line with ϕ = ϕw.

Page 193: Introduction To Dynamics

5.4 Self-excited Vibrations 185

ϕ

ϕ

ϕ0

ϕwϕ2 ϕ3 ϕ1

Fig. 5.23. Phase portrait for a friction pendulum.

The leftmost boundary of the jump line is a point satisfying ϕ = ϕw, r = −r0,hence the corresponding (negative) angle is given by

sinϕ2 =−(

r0 + 2δ ϕw

ω20

). (5.39)

All phase curves, which lead into the line ϕ2–ϕ1, go to point ϕ1 and spiral aboutthe equilibrium position defined by (5.39). With high internal damping, the phasecurves shrink to zero, with moderate internal damping, a limit cycle of the trajectoryϕ1–ϕ3 in Fig. 5.23 is created [33, 42].

5.4.5 Kármán Vortex Street

The periodic separation of vortices for round structures (wires, braces and the like),which are vertically passed by a fluid, can lead to resonance with the eigenfrequen-cies of the structure and to fractures. The most famous example is the collapse ofthe Tacoma Narrows Bridge on 7.11.1940 [32]. Less spectacular examples are vi-brations of power transmission lines or the ’singing’ of telephone lines [60]. In allcases, it is a self-excited oscillation, which takes its energy in cycles of vortex sep-arations from the flow.

Theodore von KÁRMÁN was the first to study these periodic vortex structures(1911 in Göttingen, as an assistant of Ludwig PRANDTL). He concluded that onlycertain staggered vortex configurations are stable (Fig 5.24). Such a stable configu-ration has the width-spacing ratio

hl= 0,283 . (5.40)

We consider a flow around a cylinder with diameter D. The velocity, for which thestable vortex formation and separation occur, is equal to the flow velocity V∞ minusthe so-called induced velocity u. This is caused by the circulation Γ

[m2/s

]of the

vortices:

Page 194: Introduction To Dynamics

186 5 Vibration Phenomena

V∞ unstable

stablehl

V∞

Fig. 5.24. Vortices in a flow following a circular cylinder.

Vw =V∞− u =V∞− 1√8

Γl. (5.41)

A new vortex pair is created after the time

T =

(l

V∞− u

). (5.42)

The staggered separation of vortices with frequency f may stimulate the body to adestructive oscillation, if its eigenfrequencies are excited. Therefore, we are inter-ested in estimating the frequency f for a stable vortex street.

A regular vortex street occurs only for REYNOLDS numbers Re= V∞Dν of about 60

to 5000 with the kinematic viscosity ν . For smaller REYNOLDS numbers, the flowis laminar, for larger REYNOLDS numbers, the flow is completely turbulent. In thegiven range of REYNOLDS numbers, there is a clear dependence of the REYNOLDS

number on the dimensionless STROUHAL number

S =f DV∞

(5.43)

due to measurements [60]. This relationship is shown in Fig. 5.25 as a summarizingresult of measurements and calculations. For larger REYNOLDS numbers, there is aconstant STROUHAL number S = 0,21.

For small cylinder diameters and moderate velocities, we evaluate frequencies inthe acoustic range. The well-known ’singing’ of telephone lines can be explained bythis. With an air velocity V∞ = 10 m/s and a wire diameter D = 2mm, the frequencysatisfies

f = 0,2110 m/s

0,002m= 1050s−1 . (5.44)

Page 195: Introduction To Dynamics

5.4 Self-excited Vibrations 187

101 2 4 6 8 102 2 4 6 8 103 2 4 6 8 104

0,22

0,20

0,18

0,16

0,14

0,12

S

S = f DV∞

Re = V∞Dν600

Fig. 5.25. STROUHAL number depending on the REYNOLDS number for the flow aroundcircular cylinders [60].

Thereby, the REYNOLDS number is Re ≈ 1200.Theodore von KÁRMÁN illustrated the discovery of the vortex street and the

work associated with the collapse of the Tacoma Narrows Bridge in his book DieWirbelstraße in a very illustrative way [32].

5.4.6 Flutter

Flutter is known from aircraft, helicopters, and turbines. Like the KÁRMÁN vortexstreet, the phenomenon belongs to the flow-induced self-excitation mechanisms [8].Flutter is a coupling effect between the flow conditions around the wing and the elas-tic properties of the wing; it is an aeroelastic problem that needs to be consideredin any design of wings or blades. For this, very extensive and complicated calcula-tions are necessary, which we can only indicate. We discuss the basic self-excitationmechanism with a conceptual model.

We model an airplane wing as a beam, which can perform bending and torsionalvibrations, and we take into account only the first eigenfrequency and the associatedfirst eigenmode of these bending and torsional vibrations (Chapter 3 and Fig. 5.26).The bending vibration ensures that the wing moves up and down, the torsional vi-bration generates a rotation of the wing cross section with respect to the inflow di-rection in each motion phase. Whether the wing gets to flutter or not depends on theeigenfrequencies of bending and torsion, as well as their relative phase. In Fig. 5.26,a case is shown in which bending and torsional vibration have the same eigenfre-quency but a 90o shifted phase. When the wing oscillates upwards, it is rotated atthe same time in such a way that a positive angle of attack and thus an additionalwing lift is produced. When the wing oscillates downwards, it is rotated such that anegative angle of attack and thus a negative lift is generated. In such conditions, the

Page 196: Introduction To Dynamics

188 5 Vibration Phenomena

position1

2

3

4

56

7

8

9

inflow

Fig. 5.26. Principle of flutter [8].

elastic vibrations of the wing are excited by the aerodynamics; unstable conditionsmay arise.

The self-excitation mechanism of flutter can be characterized as follows. Energyis taken from the flow energy of the surrounding fluid at the frequency of the elasticvibrations of the wing. This energy serves to maintain the oscillation process. Thetorsional vibrations act as a switch, setting the angle of attack of the wing at thecorrect moment to support the vibrations with additional positive or negative lift.The process depends on flight velocity. There are several critical velocities at whichdifferent forms of flutter can occur; also there are even more complicated flutterphenomena than the ones described above. To reduce the risk to the aircraft at suchvelocities, flutter must be reduced with appropriate design. For this, there are somegeneral approaches.

A classical approach consists in adapting mass and stiffness distributions in com-bination with the aerodynamics of the wing to shift the resonances either out of theoperational region or to avoid them. Against the background of modern computersoftware results are astonishing. A decoupling of elastic modes is difficult and theflutter problem not always completely solvable. A third important step is increasingdamping. As we saw in Fig. 5.13 and in Fig. 5.14, a self-excited oscillation is gen-erated by the balance of supplied and dissipated energy. The mean amplitude of thelimit cycle will be smaller, the more energy is dissipated. In the limit, vibration nolonger arises. As a general step in reducing oscillations, we require as much energydissipation in the system, that is the wing, as possible. This idea, of course, also hasconstructive and function-related limits.

5.4.7 Pendulum Clock

The pendulum clock with GRAHAM escapement offers a classic example of self-excited oscillation. The energy source is a torsional spring or a coiled weight, theoscillator is a pendulum, and the switch is the anchor with its input and outputpallets together with the escape wheel (Fig. 5.27). The escape wheel is held withina moment tension either by a spring or by a weight, such that it always tends tocontinue rotating in the sense of this external torque.

The GRAHAM escapement consists of an escape wheel with saw-shaped teeth,which is driven by the clock spring or by a weight, and an anchor with a fixed

Page 197: Introduction To Dynamics

5.4 Self-excited Vibrations 189

input pallet

output pallet

anchor

escape wheel

pendulum

ZL

Z′L

ZR

A1 A2

ϕ = q

S

L RV1 V2

U2 U10

E1

E2

Fig. 5.27. GRAHAM escapement of a pendulum clock [33].

connection to the pendulum. Its two pallets, the input and the output pallet, compriseparts of a ring which is concentric to the pendulum axis [7, 19, 33].

Suppose now that the pendulum is about to swing from its outermost position R tothe left; then, first, the tip ZL of a tooth of the escape wheel slides on the cylindricalportion of the outer edge of the input pallet. Thereby, the escape wheel remains atrest and does not exert any force on the pendulum, if we disregard friction forces.When the pendulum reaches position U1, which is on the right of the center 0, thetooth tip ZL of the escape wheel changes from the outer edge of the input pallet tothe transverse face E1E2 of this pallet. Now, the escape wheel can rotate further andexert a force on the pendulum, which results in a torque ML acting on the pendulumin the sense of its motion direction. This phase ends when the tooth tip ZL reachespoint E2, which corresponds to the pendulum position U2 on the left of the center

Page 198: Introduction To Dynamics

190 5 Vibration Phenomena

0. At this moment, the escape wheel continues to rotate without constraint, until thetip ZR of the tooth that is closest to the output pallet impacts the output pallet on itsinner edge. A further rotation of the escape wheel is prevented. We assume, that thependulum has not left position U2, when the tooth tip ZR hits the output pallet. Whenthe pendulum swings to the left to its maximum position L, the tooth tip ZR slideson the inner edge of the output pallet, whereby the escape wheel cannot rotate.

For the following half-cycle of the pendulum from position L to the right, thetooth tip ZR slides on the inner edge of the output pallet (beyond position U2, whichhas been reached by ZR with the previous half-cycle), until it reaches point A1 onthe transverse face of the output pallet. The pendulum position achieved, V1, is stillleft of the center 0. Now, the tooth tip ZR slides on the transverse face A1A2 of theoutput pallet. Thereby, the escape wheel rotates and exerts a force on the pendulum,which results in a torque ML acting in the direction of motion. When the tooth tip ZR

reaches point A2, which corresponds to a pendulum position V2 on the right of U1,the escape wheel rotates without restriction and very rapidly, until the next tooth tipZ′

L strikes the outer edge of the input pallet. Then, the escape wheel stops again. Aslong as the pendulum finishes its half-cycle to the right, the tooth tip Z′

L slides on the

U1

U2

V2

V1

RLq

q

−q2−q1

+q1+q2

driven

driven

free pendulum oscillation

free pendulum oscillation

qcos(ϑ)

Fig. 5.28. Simplified representation of the limit cycle of a pendulum clock with GRAHAM

escapement [33].

Page 199: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 191

outer edge of the input pallet. When a position corresponding to the initial positionR is reached, the whole procedure starts again with a new half-cycle to the left butthe escape wheel rotated by one tooth. The input and the output pallet at the anchorare set in such a way, that the pendulum positions U1 and V1 as well as U2 and V2

are symmetric to the center 0, where 0U2 = 0V 2 is larger than 0U1 = 0V 1 [33].Without going into detail on the extensive calculations connecting one phase to

the other, let us consider the limit cycle, which is the basis for the stable pendulummotion. The angles q belonging to U1, U2, V1, V2 are measured from the perpendicu-lar line and are denoted by +q1, −q2, −q1, +q2, with q2 > q1 > 0. Then in the limitcycle, the motion of the pendulum can be described as in the following section.

From position V2 (+q2), the pendulum swings to U1(+q1) via R (q = 0) describ-ing an elliptic arc. Along the arc from (+q1) to (−q2) corresponding to (U1 −U2)it is driven. On the left side, the pendulum swings from U2 (−q2) to V1 (−q1) viaL (q = 0) and is driven along the arc from V1 (−q1) to V2 (+q2). The connection ofthese phases shows a limit cycle, which is composed of elliptic spirals as a result of(dimensionless) dissipation ϑ on the pallet faces (Fig. 5.28 and [33]).

5.5 Parametrically Excited Vibrations

Parametrically excited vibrations arise from periodically time-varying parametersof the considered vibration system and influence machine dynamics only when thesystem is displaced from its undisturbed equilibrium position.

5.5.1 Overview

Typical practical examples of parametrically excited oscillations are asymmetries ofrotors, offset within drive shafts or the periodically time-varying tooth stiffness ingear transmissions [8, 35, 61]. The latter example shows all typical characteristicsof this vibration type. The tooth contact for a spur gear occurs along the so-calledcontact line, where the number of contacting teeth changes. This means that thetooth stiffness kv(t) fluctuates periodically about a mean value (Fig. 5.29). For a

gear wheel 1

gear wheel 2

dν (t)

kν (t)

kv(t)

kv0

t

given

approximation

Fig. 5.29. Model of a single stage, spur gear transmission, and the corresponding tooth stiff-ness function with approximation.

Page 200: Introduction To Dynamics

192 5 Vibration Phenomena

mechanical model, the time-varying gear contact can be represented with corre-sponding springs and dampers (Fig. 5.29). Deriving the equations of motion, we ob-tain time-varying damping and stiffness matrices as typical characteristics of para-metrically excited behavior.

Phenomena that can only be observed in parametrically excited systems and thatare sometimes critical are combination resonances. For excitation of the systemwith angular frequencies Ω in the vicinity of specific combinations of eigen angularfrequencies of the undamped time-invariant system, the amplitudes may increase.We obtain resonance phenomena and possible instabilities. These parameter andcombination resonances occur for angular frequencies of the following structure:

Ω =1p(qkωk ± qlωl) . (5.45)

It is (p,qk,ql) = 1,2,3, . . . and (k, l) = 1,2,3, . . . f with the number of degrees offreedom f . The eigen angular frequencies of the undamped, time-invariant systemare given by (ωk,ωl) with dv(t) = 0 and kv(t) = kvo.

We have to avoid these exciting angular frequencies, which correspond to (5.45).It is not sufficient to analyze the vibrational behavior of a parametrically excitedsystem only considering the eigen angular frequencies; this can lead to dangerousconclusions.

Sometimes parametrically excited vibrations are called rheonomic vibrations, areference to rheonomic (time-variable) constraints in analytical mechanics. Depend-ing on the type of system, we get rheo-linear or rheo-nonlinear vibrations.

5.5.2 Motion and Stability of Parametrically Excited Vibrations

5.5.2.1 Pendulum with Moving Suspension Point

If the suspension point A of a pendulum (Fig. 5.30 and Example 4.2) is moved upand down with the periodic time-variable acceleration a(t), reaction forces a(t)dmwill be created at the mass element of the pendulum. For such a physical pendulum,we get the equation of motion

Jϕ+ml (g+ a) sinϕ = 0 (5.46)

with moment of inertia J with respect to A and

a = a0 cos(Ω0t) . (5.47)

We linearize the equation of motion about the lower equilibrium point. With

τ =Ω0t , λ =mlg

Ω 20 J

, γ =mla0

Ω 20 J

,d

dτ= ( · )′ , (5.48)

Page 201: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 193

A

l

mg

ma

a(t)

Fig. 5.30. Pendulum with driven suspension point.

we get the dimensionless form [33]

ϕ ′′+(λ + γ cosτ)ϕ = 0 . (5.49)

5.5.2.2 Kane’s Baby Shoes

A Stanford student observed, that a pair of small baby shoes hanging from the mirrorof a car started to vibrate strongly in certain car situations and at certain frequencies.KANE used a model with two degrees of freedom to show that the rocking is aparametrically excited vibration [31]. Fig. 5.31 depicts the principal situation. Theequations of motion for this model are

h

l

ϕ

bx

y

z

δϑ

Fig. 5.31. Model for the baby shoes according to KANE [31].

Page 202: Introduction To Dynamics

194 5 Vibration Phenomena

(1+ acos2 ϑ

)ϕ− aϑ ϕ sin2ϑ +(1+ a)Ω 2

0 sinϕ = 0 , (5.50)

ϑ +ϕ2

2sin2ϑ = 0 . (5.51)

Thereby, we assume that the shoe of mass m and moment of inertia Ix ≈ Iz, Iy � 1about its center of gravity is hanging from a massless string, such that a = Ix

ml2 and

Ω0 =mlg

ml2+Ix. To get a first approximation, we further assume that

• the minimal coordinate ϕ oscillates harmonically: ϕ = ϕ0 cos(Ω0t),• the minimal coordinate ϑ is small: ϑ � 1.

From (5.51), we get

ϑ +

(ϕ2

0Ω20

2

)(1− cos2Ω0t)ϑ = 0 . (5.52)

After normalization with τ = 2Ω0t and ddτ = ( · )′, this yields

ϑ ′′+(λ + γ cosτ)ϑ = 0 (5.53)

with

λ =−γ =ϕ2

0

8. (5.54)

5.5.2.3 Mathematical Relationships

There are many more examples of parametrically excited vibrations [2, 10, 8, 26, 33,35, 42, 44, 61]. The equations of motion of type (5.49) and (5.53) are typical. In thefollowing, we discuss these ordinary, linear, and time-variant differential equationsin detail.

For the case of an oscillator with one degree of freedom and parametric excita-tion, we generally get:

x+ p1(t)x+ p2(t)x = 0 . (5.55)

With the ansatz

x = ye−12∫

p1(t)dt , (5.56)

equation (5.55) can be transformed into

y+P(t)y = 0 (5.57)

with P(t) = p2(t)− 12

ddt [p1(t)]− 1

4 p21(t). As p1(t) and p2(t) are assumed to be pe-

riodic, also P(t) is periodic:

Page 203: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 195

P(t +T) = P(t) . (5.58)

Equation (5.57) with (5.58) is called the HILL differential equation. For systemswith multiple degrees of freedom, we obtain the same structure, however y(t) is avector and P(t) a matrix.

A solution can be found with FLOQUET theory [61, 2]. In the case with onedegree of freedom, we choose the ansatz

y(t) =C1eμ1t y1(t)+C2eμ2t y2(t) . (5.59)

Thereby, y1 and y2 are periodic functions of time, C1 and C2 are constants, and μ1

and μ2 are so-called characteristic exponents of (5.57). These exponents depend onthe values in (5.57), but not on the initial conditions. They define the stability behav-ior of the solution. If one of the characteristic exponents has a positive real part, thesolution (5.59) increases without restriction; it is unstable. If the real parts of bothexponents are negative, then y decreases to zero; the solution is (asymptotically)stable. In the limiting case, the real part of one exponent (or both exponents) mayvanish. Then y is bounded without approaching zero asymptotically; y may be pe-riodic in this case. For the study of vibrations, real exponents are of interest. Then,the regions of stable solutions are separated from the regions of unstable solutionsby a limiting characteristic of pure periodic solutions. Thus, the search for unstableregions results in the determination of conditions for vanishing exponents, that ispure periodic solutions.

For some specific forms of the periodic functions P(t), the solutions of (5.57)have been systematically analyzed:

P(t) = P0 +ΔP cos(Ω t) , (5.60)

P(t) = P0 +ΔPsgn(cos(Ω t)) . (5.61)

For the first case, the parameter fluctuates according to a harmonic law, for the sec-ond case, the changes are abrupt, such that P(t) is a Meander function. With (5.60),the HILL differential equation transforms into a MATHIEU differential equation;with (5.61), it transforms into a MEISSNER differential equation.

Like in the above examples, we set

τ =Ω t ,d

dτ= ( · )′ , λ =

P0

Ω 2 , γ =ΔPΩ 2 . (5.62)

This yields the normal form of the MATHIEU and the MEISSNER differential equa-tion:

MATHIEU differential equation y′′+(λ + γ cosτ)y = 0 , (5.63)

MEISSNER differential equation y′′+(λ + γsgn(cosτ))y = 0 . (5.64)

The stability behavior of the MATHIEU differential equation has been analyzedby INCE and STRUTT [38]. The stability just depends on the parameters (λ ,γ),

Page 204: Introduction To Dynamics

196 5 Vibration Phenomena

whereby γ = 0 defines the periodic limiting case and γ �= 0 leads to unstable re-gions, which grow for increasing γ (Fig. 5.32).

stable

γ = λ

λ1 2 3 4 5

unstable

0

γ

4

3

2

1

γ =−λ

-4

-3

-2

-1

Fig. 5.32. Stability map according to INCE/STRUTT for the MATHIEU differential equation.

For negative λ and γ = 0, there are no parametric excitations and the system isunstable. If we consider an oscillator with constant, nonvanishing γ , its representa-tion in the stability map will move along a parallel to the λ -axis for changing λ .For λ > 0, unstable regions may be crossed. In practice, the oscillator is stable forγ = 0 and may become unstable for γ �= 0 and specific values of λ . The oscillatingpart with γ may have a stability reducing effect. In the case λ < 0, the oscillatoris unstable for γ = 0 and may become stable for γ �= 0. The oscillating part has astabilizing effect.

The tips of the unstable regions touch the abscissa (λ -axis) at the values

λ =(n

2

)2(n = 1,2, . . .) . (5.65)

The width of the regions – and therefore also their practical significance – reduceswith increasing n. This can be explained by damping mechanisms, which have notbeen considered within the present analysis but which occur for real oscillators.They yield decreasing unstable regions [42].

Page 205: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 197

In many cases, the area around the origin λ = γ = 0 of the stability map is ofinterest. Here, the boundary between the stable and unstable regions can be suffi-ciently approximated by functions λ = λ (γ). For the first three boundary curves, wesummarize without proof:

λ1 ≈−12γ2 , (5.66)

λ2 ≈ 14− γ

2, (5.67)

λ3 ≈ 14+

γ2. (5.68)

Fig. 5.33 shows an enlarged part of the stability map with the approximating bound-ary curves drawn as dotted lines.

0,2

0,250

γ = λ

λ

γ

γ =−λ

Fig. 5.33. Local stability map according to (5.66)-(5.68).

We return to the (linearized) examples. For the pendulum with moving suspensionpoint (Section 5.5.2.1), we have to distinguish two cases, the hanging and the stand-ing (inverted) pendulum. Equation (5.49) has been derived by linearization aboutthe lower equilibrium point; for the hanging pendulum, we have λ > 0. If the valuefor λ satisfies 0 < λ < 0,25, the oscillator will become unstable with increasing ex-citation amplitude γ (Fig. 5.32). Depending on the value of λ , the process will run

Page 206: Introduction To Dynamics

198 5 Vibration Phenomena

at different speeds. If we keep γ constant and vary λ with the frequency (Ω =Ω0),different stable and unstable regions are crossed along a horizontal line.

For the standing pendulum, the linearization has to be done about the upper equi-librium point: λ < 0. According to Figs. 5.32 and 5.33, only a small region of stableoscillations is possible. Without driving the suspension point (γ = 0), the pendulumis located in an unstable equilibrium point because of the center of gravity lyingabove the suspension point. It is remarkable that the unstable upper equilibriumposition of the pendulum for a resting suspension point can be stabilized by appro-priate vibration of the suspension point. This means that the pendulum may performstable oscillations for small displacements from this equilibrium position.

0

1

2

ϕ

Fig. 5.34. Explanation of the stabilizing effect for the pendulum with driven suspension point.

We give a physical explanation for this stabilizing effect (Fig. 5.34). We considera periodic motion between points 1 and 2, where the velocity vanishes in points 1and 2. It is positive in an upward direction and negative in a downward direction.Then, the acceleration in the area 0 – 2 – 0 is positive in an upward direction and theacceleration in the area 0 – 1 – 0 is negative in a downward direction. As a reaction,there is an acceleration of the pendulum mass in a downward direction for the area0 – 2 – 0 of the suspension point; for the area 0 – 1 – 0, there is an acceleration inan upward direction. For the latter case, the angle ϕ has a larger mean value thanfor the first case. As a consequence, there is always some rest acceleration in anupward direction and thus a torque left, which directs the pendulum to the upperequilibrium position. We can call this torque a jarring direction torque. If this torqueis larger than the torque due to the gravitational force, then the pendulum will stay inthe upper equilibrium positions and will not be destabilized by small perturbations.

The jarring direction torque may generate a shift of a beacon course when excit-ing in a corresponding direction. Thus, hanging measurement equipment might thenbe sensible in a direction of acceleration which is not desired.

For KANE’s baby shoes (Section 5.5.2.2), it is λ = −γ ≥ 0. The relevant lineγ = −λ is shown in Figs. 5.32 and 5.33. With increasing amplitude ϕ0, the doublependulum crosses different stable and unstable regions. If we transfer these regionsinto an amplitude stability diagram with ϕ0 as the polar angle, the single regions canbe highlighted; they can be easily tested experimentally (Fig. 5.35).

Page 207: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 199

ϕ

ϑ

180◦

156◦

115◦

66◦

0◦

stable

stable

ϕ0

Fig. 5.35. KANE’s baby shoes: typical stability regions for specific geometry and materialvalues.

5.5.3 Examples

In the following, we consider classical cases for parametrically excited oscillationsin practice.

5.5.3.1 Jeffcott Rotor with Stiffness Asymmetries

As an example of rotating machinery components with asymmetric behavior, weconsider the so-called Jeffcott rotor [73] with nonsymmetric shaft stiffness. Therotor is assumed to be a point mass, the mass of the elastic shaft is neglected oradded to the point mass. The shaft is mounted statically determinate and the speed Ωof the rotor is constant (Fig. 5.36). For the description, we use an inertial coordinatesystem I and a co-rotating coordinate system R with the same origin. If we consider

Ω t

Iy

Iz = Rz

Ix

Rx

Ry

Ω

Ω t

S

Fig. 5.36. Model of a Jeffcott rotor.

Page 208: Introduction To Dynamics

200 5 Vibration Phenomena

a stiffness asymmetry of the elastic shaft, the elastic restoring forces will act on thepoint mass of the rotor in the co-rotating coordinate system (Fig. 5.37):

RFc =−(

cx Rxcy Ry

). (5.69)

Neglecting internal and external damping forces, the equations of motion in the co-rotating coordinate system are

mRx− 2mΩ Ry+(cx −mΩ 2)

Rx = 0 , (5.70)

mRy+ 2mΩ Rx+(cy −mΩ 2)

Ry = 0 . (5.71)

With the abbreviations

εS =cx − cy

cx + cy, ω2

x =cx

m, ω2

y =cy

m, (5.72)

ω2 =ω2

x +ω2y

2, Ω0 =

Ωω

(5.73)

and the transformation

τ = ωt ,ddt

= ωd

dτ= ω (·)′ , (5.74)

these equations can be written as

Rz′′+ RG Rz′+ RK Rz = 0 (5.75)

with

Rz =(

Rx

Ry

), RG =

(0 −2Ω0

2Ω0 0

), RK =

(1+ εS −Ω 2

0 00 1− εS −Ω 2

0

).

(5.76)

For practical reasons, the dynamic behavior of the Jeffcott rotor in Fig. 5.36 is an-alyzed in the inertial coordinate system, which makes sense for estimating the load

Ry

Rx

Ix

Iy

Ω t

−cy Ry

−cx Rx

Fig. 5.37. Stiffness asymmetry and its effect on the rotor.

Page 209: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 201

on the bearings. The transformation from the co-rotating to the inertial coordinatesystem is a rotation about the common z-axis:

Rz = ARI Iz with ARI =

(cos(Ω0τ) sin(Ω0τ)−sin(Ω0τ) cos(Ω0τ)

). (5.77)

Inserting this transformation in (5.75), we obtain a linear system of differential equa-tions with periodic coefficients:

Iz′′+ IK Iz = 0 with IK =

(1+ εS cos2Ω0τ εS sin2Ω0τεS sin2Ω0τ 1− εS cos2Ω0τ

). (5.78)

This is only a formally parametrically excited oscillation of a system with two de-grees of freedom and, at the same time, the simplest possible representation of arotating machine component with nonsymmetric stiffness. Equation (5.78) could besolved with the help of FLOQUET theory [61]. However in this case, a stability pre-diction can also be obtained from the equations of motion (5.75) in the co-rotatingcoordinate system with constant coefficients. The ansatz Rz = Rz0 exp(λτ) yieldsthe characteristic equation

λ 4 + 2(1+Ω 2

0

)λ 2 +

[(1−Ω 2

0

)2 − ε2S

]= 0 . (5.79)

According to the STODOLA criterion (2.121), all the coefficients of the characteristicequation have to be positive. This is satisfied for the first two coefficients, (1) and(2(1+Ω 2

0

)). For the last expression, we have to require

ε2S <

(1−Ω 2

0

)2for stability . (5.80)

This result is sketched in Fig. 5.38. We recognize, that the system becomes unsta-ble for the undamped, nonsymmetric Jeffcott rotor and Ω0 =

Ωω → 1. However, the

always present internal and external dampings in practice ensure that a stable oper-ation is possible for Ω0 → 1. Therefore, the asymmetries must not be too large. Acase with damping is presented in Fig. 5.38 using a dotted line.

1

1

with damping

without damping

Ω0 =Ωω

unstable region

εS

Fig. 5.38. Stability map for the Jeffcott rotor with stiffness asymmetry.

Page 210: Introduction To Dynamics

202 5 Vibration Phenomena

Despite the simple model assumptions, equation (5.78) and the stability map inFig. 5.38 already show some essential features of vibrations of rotor systems. Ac-cording to (5.78), the effective shaft stiffness varies twice during a period betweenits largest value (1+ εS) and its smallest value (1− εS). This procedure results ina continuously changing load on the shaft and on the bearings, such that just be-cause of this effect the asymmetries have to be minimized. In addition according toFig. 5.38, the stability of the rotor system will be risky, if a critical speed (in thecase Ω0 = 1) has to be passed. Even with existing dampings, instability peaks occurin such situations, which go very far downwards and therefore also require minimalasymmetries (see next section).

5.5.3.2 Stability and Control of Rotors

A more sophisticated example of a rotor with asymmetric stiffness than in the lastsection has been treated in [2]. Fig. 5.39 shows an ultracentrifuge model, where therotor stiffness is unbalanced due to an elastic rectangular rod. The stability behaviorof this rotor with 18 degrees of freedom has been analyzed and improved by meansof some control. The control forces are generated by magnetic bearings, which arenot shown. Without going into the very complex modeling and design of a controller,we show only the most important results.

upper bearing

shaft

rotor section

link

control forces

rotor section

tip

lower bearing

figure axis

Fig. 5.39. Rotor with asymmetric stiffness [2].

Page 211: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 203

The diagram on the left of Fig. 5.40 shows the stability behavior of the rotor with-out control. We can think of it as a kind of INCE-STRUTT diagram for a system withmany degrees of freedom (analogous to Fig. 5.32). The abscissa indicates the speed,the ordinate gives the dimension εS (5.72) for the asymmetry of the rotor. One cansee some single and combination resonances A-H, for which the motion becomesunstable for minor asymmetries (tip down at A). A stable operation of the rotorwithout controller is no longer possible from speeds ω ≈ 210 1/s ≈ 2000Upm. Withan optimized control, the stability can be extended to ω ≈ 1800 1/s ≈ 17000Upm(diagram to the right in Fig. 5.40). The diagram on the lower right in Fig. 5.40 is aclose-up of the upper right diagram in the low speed range.

0,0

0,2

0,4

0,6

0,8

εS

0 100

200 300

500 1000 15002000,0

0,2

0,4

0,6

0,8

εS

0

0,0

0,2

0,4

0,6

0,8

εS

0 100

unstable

with control

with control

εS = 0,0039without controlΩ(1/s)

AB

CD

EF

GH

DG

D

G

B = 12 (ω1 +ω2)

C = 12 (ω1 +ω3)

E = 12 (ω1 +ω4)

A = ω1

D = ω3

F = 12 (ω2 +ω4)

H = ω5

I = ω6

G = ω4

IΩ(1/s)

Ω(1/s)

unstable

unstable

Fig. 5.40. Improving the stability of an asymmetric rotor with a control.

The results clearly show that vibrations of moving machine parts may pose aserious risk. Diagram 5.40 for the noncontrolled case proves that combination reso-nances (for ω1+ω3

2 ) can be just as dangerous as single resonances (for ω1).

5.5.3.3 Gear Drive

In the overview 5.1, the periodically time-varying stiffness of a gear drive was pre-sented as a typical example of a parametrically excited vibration. In the manualtransmission of a car as a component of the entire drive train, the teeth contacts ofthe locked gears create parametric excitation in the entire system. Fig. 5.41 showsthe mechanical model of such a drive train with 26 degrees of freedom and high-lights the detailed conditions for a single tooth contact (a locked gear consists of

Page 212: Introduction To Dynamics

204 5 Vibration Phenomena

M F M

ϕ ϕ

x

y

z ϕγ

κ

x

Fig. 5.41. Mechanical model of a passenger car drive train and a single gear ratio.

two meshings: constant and gear ratio). The problem of such a drive train is less aquestion of stability than a question of resonances, particularly because it is oper-ated over a large speed range. The parametrically excited vibrations yield additionalcritical resonances because of the possible combination resonances. The study ofthese resonances with the help of a mechanical model and the resulting simulationmodel is not treated. Fig. 5.42 shows a typical result of a parametrically excitedvibration (resulting from teeth contacts).

The time-varying tooth stiffness is the dotted curve. It shows about 10 down-ward tips during a full revolution. These stiffness peaks cause a shock to the teeth incontact and displace them abruptly. The structural damping ensures that the displac-ing impact almost completely decays until the next stiffness peak. Then the processbegins again.

5.5.3.4 Pendulum with Elastic String

The pendulum with elastic string (Fig. 5.43) shows a type of coupled oscillation,which cannot be calculated from simple superpositions. The elastic suspension pro-vides a kind of parametric excitation, which results in a periodic change of the strokemovement and swinging if the parameter values are adjusted appropriately. To un-derstand the process, we derive the equations of motion. The kinetic and potentialenergy of the pendulum are

Page 213: Introduction To Dynamics

5.5 Parametrically Excited Vibrations 205

toot

hst

iffn

ess,

toot

hdi

spla

cem

ent

speed

Fig. 5.42. Tooth stiffness (dotted) and tooth displacement (solid line) in a meshing during arevolution.

T =m2

[x2 +(R+ x)2 ϑ 2

], (5.81)

V = mg(R+ x)(1− cosϑ)+12

cx2 . (5.82)

Applying LAGRANGE’s equations of the second kind (1.178) for the generalizedcoordinates q = (x,ϑ)T and assuming small angles ϑ � 1 and velocities ϑ � 1,we obtain the equations of motion in the linearized form

ϑ

aR

xc

m

Fig. 5.43. Pendulum with elastic string.

Page 214: Introduction To Dynamics

206 5 Vibration Phenomena

x+ω2x x = 0 , (5.83)[

1+( x

R

)]ϑ + 2

(xR

)ϑ +ω2

ϑϑ = 0 (5.84)

with ω2x = c

m and ω2ϑ = g

R . The length R corresponds to that of the static equilibriumposition. As can be seen from (5.84), the harmonic oscillation of the x-coordinate

x = x0 cos(ωxt +ϕ) (5.85)

acts as a parametric excitation for the ϑ -oscillation. If we restrict to small oscillationamplitudes with the ansatz

ϑ ≈ ϑ0 cos(ωϑ t +ψ)+Δϑ , (5.86)

we obtain the relationship

Δϑ +ω2ϑΔϑ ≈ 1

2

(x0

R

)ϑ0ωϑ {(ωϑ − 2ωx)cos((ωx +ωϑ ) t +(ϕ−ψ))

+(ωϑ + 2ωx)cos((ωx −ωϑ ) t +(ϕ+ψ))} (5.87)

from (5.84) and (5.85) neglecting quadratic and higher terms in Δϑ and x0R . This can

be interpreted as an undamped, forced oscillation according to Section 5.3. Hence,it has the solution

Δϑ =V1 cos((ωx −ωϑ )t +(ϕ−ψ))+V2 cos((ωx +ωϑ ) t +(ϕ+ψ)) (5.88)

with

V1 =12

(x0

R

)ϑ0

(ωϑωx

)(ωϑ − 2ωx

2ωϑ −ωx

), (5.89)

V2 =12

(x0

R

)ϑ0

(ωϑωx

)(ωϑ + 2ωx

2ωϑ +ωx

). (5.90)

It can be concluded that for small oscillation amplitudes, the stroke and swing oscil-lations are decoupled as long as 2ωϑ �= ωx. No vibration excites the other, and Δϑis very small because of x0

R ϑ0 � 1. However, if 2ωϑ approaches ωx, that is

4( g

R

)≈ c

m, (5.91)

the amplification function V1 increases and the oscillation in the x-direction influ-ences strongly the ϑ -oscillation. There is a periodic exchange between stroke andswing motion. These phenomena can be realized and observed with a tuned pen-dulum according to Fig. 5.43. It can be seen that for large oscillation amplitudes,there is always a coupling between stroke and swing oscillation. In (5.83)-(5.84),the nonlinear couplings between both motion directions have been neglected.

Page 215: Introduction To Dynamics

References

[1] Angeles, J.: Dynamic Response of Linear Mechanical Systems. Springer, New York(2012)

[2] Anton, E.: Stabilitätsverhalten und Regelung von parametererregten Rotorsystemen,Fortschritt-Berichte VDI: Reihe 8, Mess-, Steuerungs- u. Regelungstechnik, vol 67.VDI Verlag, Düsseldorf (1984)

[3] Arnold, V.: Mathematical Methods of Classical Mechanics, 2nd edn. Springer, Berlin(1997)

[4] Bathe, K.J.: Finite element procedures. MIT, Cambridge (2007)[5] Bauchau, O.: Flexible Multibody Dynamics. Springer, Berlin (2010)[6] Becker, E., Bürger, W.: Kontinuumsmechanik. Teubner, Stuttgart (1975)[7] Berthoud, F.: Anweisung zur Kenntnis, zum Gebrauch und zur guten Haltung der Wand-

und Taschenuhren. Reprint, Zentralantiquariat der DDR, Meissen (1818)[8] Bishop, R.: Schwingungen in Natur und Technik. Teubner, Stuttgart (1985)[9] Braess, D.: Finite elements: theory, fast solvers, and applications in elasticity theory,

3rd edn. Cambridge University Press, Cambridge (2007)[10] Bremer, H.: Dynamik und Regelung mechanischer Systeme. Teubner, Stuttgart (1988)[11] Bremer, H.: Elastic Multibody Dynamics. Springer, New York (2008)[12] Brenner, S., Scott, R.: The mathematical theory of finite element methods. Springer,

Berlin (1994)[13] Courant, R.: Variational methods for the solution of problems of equilibrium and vibra-

tions. Bull. Amer. Math. Soc. 49:1–49:23 (1943)[14] Courant, R., Hilbert, D.: Methods of mathematical physics. Wiley, New York (1989)[15] Dresig, H., Holzweißig, F.: Dynamics of machinery: theory and applications. Springer,

Berlin (2010)[16] Fischer, U., Stephan, W.: Prinzipien und Methoden der Dynamik. VEB Deutscher

Verlag der Wissenschaften, Leipzig (1972)[17] Galerkin, B.: Rods and plates. series occurring in various questions concerning the elas-

tic equilibrium of rods and plates. Engineers Bulletin (Vestnik Inzhenerov) 19:897–19:908 (1915)

[18] Gander, M., Wanner, G.: From Euler, Ritz, and Galerkin to modern computing. SIAMRev 54, 627–666 (2012)

[19] Gelcich, E.: Geschichte der Uhrmacherkunst. Reprint, Zentralantiquariat der DDR,Weimar (1892)

[20] Geradin, M., Rixen, D.: Mechanical Vibrations. Wiley, New York (1997)

Page 216: Introduction To Dynamics

208 References

[21] Goldstein, H., Poole, C., Safko, J.: Classical mechanics, 3rd edn. Addison-Wesley,Reading (2001)

[22] Greiner, W.: Classical mechanics: systems of particles and Hamiltonian dynamics.Springer, Berlin (2008)

[23] Gross, D., Hauger, W., Schröder, J., Wall, W., Govindjee, S.: Engineering mechanics 3,2nd edn. Springer, Berlin (2014)

[24] Guckenheimer, J., Holmes, P.: Nonlinear oscillations, dynamical systems, and bifurca-tions of vector fields. In: Applied Mathematical Sciences, 7th edn., vol. 42, Springer,New York (2002)

[25] Hadwich, V.: Modellbildung in mechatronischen Systemen. Fortschritt-Berichte VDI,VDI Verlag (1998)

[26] Hagedorn, P.: Non-linear Oscillations. Oxford University Press, New York (1981)[27] Hahn, W.: Stability of Motion. Springer, Berlin (1967)[28] Hamel, G.: Theoretische Mechanik. Springer, Berlin (1978)[29] Huber, R., Clauberg, J., Ulbrich, H.: Herbie: Demonstration of gyroscopic effects by

means of a RC vehicle. In: Proceedings of the ASME 2011 IDETC/CIE Conference,Washington, August 29-31 (2011)

[30] Ibrahimbegovic, A.: On the choice of finite rotation parameters. Comput. Meth. Appl.Mech. Eng. 149, 49–71 (1997)

[31] Kane, T.: Mechanical demonstration of mathematical stability and instability. Int. J.Mech. Eng. Educ. 2(4), 45–47 (1974)

[32] von Karman, T.: Die Wirbelstraße. Hoffmann und Campe, Hamburg (1968)[33] Kauderer, H.: Nichtlineare Mechanik. Springer, Berlin (1958)[34] Kirchhoff, G.: Vorlesungen über Mathematische Physik, Mechanik. Teubner, Leipzig

(1876)[35] Kücükay, F.: Dynamik der Zahnradgetriebe, Modelle, Verfahren, Verhalten. Springer,

Berlin (1987)[36] La Salle, J., Lefschetz, S.: Stability by Liapunov’s direct method with applications.

Academic Press, New York (1961)[37] Laursen, T.: Computational contact and impact mechanics. Springer (2002)[38] Leipholz, H.: Stability Theory. Wiley, New York (1987)[39] Lichtenberg, A., Lieberman, M.: Regular and Stochastic Motion. Springer, Berlin

(1983)[40] Magnus, K.: Kreisel-Theorie und Anwendungen. Springer, Berlin (1971)[41] Magnus, K., Müller-Slany, H.: Grundlagen der technischen Mechanik, 7th edn. Leitfä-

den der angewandten Mathematik und Mechanik, Teubner, Wiesbaden (2005)[42] Magnus, K., Popp, K., Sextro, W.: Schwingungen, 8th edn. Vieweg, Wiesbaden (2008)[43] Meirovitch, L.: Analytical Methods in Vibrations. The MacMillan Company, New York

(1967)[44] Minorsky, N.: Nonlinear Oscillations. Krieger Publishing Company, Princeton (1974)[45] Müller, P.: Stabilität und Matrizen. Springer, Berlin (1977)[46] Müller, P., Schiehlen, W.: Lineare Schwingungen. Koch Buchverlag, Planegg (1982)[47] Nayfeh, A.: Problems in Perturbation. Wiley, New York (1993)[48] Nayfeh, A., Balachandran, B.: Applied Nonlinear Dynamics. Wiley, New York (1995)[49] Papastavridis, J.: Tensor calculus and analytical dynamics. Taylor & Francis, London

(1998)[50] Papastavridis, J.: Analytical Mechanics: A Comprehensive Treatise on the Dynamics of

Constrained Systems: For Engineers, Physicists, and Mathematicians. Oxford Univer-sity Press (2002)

Page 217: Introduction To Dynamics

References 209

[51] Pfeiffer, F.: Mechanical System Dynamics. Springer, Heidelberg (2008)[52] Pfeiffer, F., Fritzer, A.: Resonanz und Tilgung bei spielbehafteten Systemen. J. Appl.

Math. Mech. 4, 38–40 (1992)[53] Pfeiffer, F., Glocker, C.: Multibody Dynamics with Unilateral Contacts. Wiley, New

York (1996)[54] Popper, K.: Objektive Erkenntnis, ein evolutionärer Entwurf. Hoffmann und Campe,

Hamburg (1993)[55] Post, J.: Objektorientierte Softwareentwicklung zur Simulation von Antriebssträngen,

Fortschritt-Berichte VDI: Reihe 11, vol 317. VDI Verlag, Düsseldorf (2003)[56] Rayleigh, J.: The Theory of Sound. Macmillan, London (1877)[57] Ritz, W.: über eine neue Methode zur Lösung gewisser Variationsprobleme der math-

ematischen Physik. Journal für die reine und angewandte Mathematik 135:1–135:61(1908)

[58] Rudin, W.: Principles of mathematical analysis, 3rd edn. International series in pure andapplied mathematics, McGraw-Hill, New York (2008)

[59] Schiehlen, W., Eberhard, P.: Technische Dynamik, 3rd edn. Teubner, Wiesbaden (2012)[60] Schlichting, H.: Grenzschichttheorie. Springer, Berlin (2006)[61] Schmidt, G.: Parametererregte Schwingungen. VEB Deutscher Verlag der Wis-

senschaften, Berlin (1975)[62] Schwertassek, R., Wallrapp, O.: Dynamik flexibler Mehrkörpersysteme. Vieweg, Wies-

baden (1999)[63] Shabana, A.: Dynamics of multibody systems, 3rd edn. Cambridge University Press,

New York (2005)[64] Stoer, J., Bulirsch, R.: Introduction to numerical analysis. Texts in applied mathematics,

vol. 12. Springer, New York (2010)[65] Strang, G.: Introduction to Linear Algebra. Wellesley-Cambridge Press, Wellesley

(2009)[66] Synge, J.L.: Classical Dynamics, in Encyclopedia of Physics, Volume III/1: Principles

of Classical Mechanics and Field Theory. Springer, Berlin (1960)[67] Szabo, I.: Geschichte der mechanischen Prinzipien und ihrer wichtigsten Anwendun-

gen, Korr. Nachdruck der 3. Auflage edn. Birkhäuser, Basel (1996)[68] Szabo, I.: Einführung in die Technische Mechanik, 8th edn. Springer, Berlin (2003)[69] Tauchert, T.: Energy Principles in Structural Mechanics. McGraw-Hill, New York

(1974)[70] Thom, R.: Structural Stability and Morphogenesis. Westview Press, Boulder (1994)[71] Thompson, M., Stewart, B.: Nonlinear Dynamics and Chaos. Wiley, New York (2001)[72] Ulbrich, H.: Dynamik und Regelung von Rotorsystemen, Fortschritt-Berichte VDI:

Reihe 11, vol 86. VDI Verlag, Düsseldorf (1986)[73] Ulbrich, H.: Maschinendynamik. Teubner Studienbücher, Teubner, Stuttgart (1996)[74] Wriggers, P.: Computational contact mechanics, 1st edn. Wiley, Chichester (2002)[75] Zeemann, E.: Catastrophe Theory, Selected Papers 1972-1977. Addison-Wesley, Read-

ing (1977)[76] Zeidler, E., Hackbusch, W., Schwarz, H.R.: Oxford Users’ Guide to Mathematics.

Oxford University Press, New York (2004)[77] Ziegler, F.: Mechanics of solids and fluids, 2nd edn. Springer, Berlin (1998)[78] Zienkiewicz, O., Taylor, R., Zhu, J.: The Finite Element Method Set, 6th edn.

Butterworth-Heinemann, Oxford (2005)

Page 218: Introduction To Dynamics

Index

acceleration, 16absolute, 16applied, 19centripedal, 19Coriolis, 19relative, 16, 19rotational, 19

action, 57amplification factor, 171approximation, 117

Bubnov-Galerkin, 123error, 118of continuum system, 117Rayleigh-Ritz, 119

beam, 104bending vibration with longitudinal load

Bubnov-Galerkin, 133Rayleigh-Ritz, 130

bending vibrator, 104bending vibrator with an end mass, 109cantilever

Bubnov-Galerkin, 128Rayleigh-Ritz, 121

torsional vibrator with an end mass, 112Boltzmann axiom, 23boundary condition, 106

Bubnov-Galerkin, 127essential, 127free, 127geometric, 107, 127kinematic, 127kinetic, 107, 127natural, 127

Rayleigh-Ritz, 127boundary integral, 126Bubnov-Galerkin method, 123

calculus of variations, 9Cardan angles, 13Cauchy’s stress tensor, 23characteristic equation, 77configuration space, 14, 72constraint, 5

hidden, 6holonomic, 6kinematics, 5nonholonomic, 6rheonomic, 5scleronomic, 5

coordinate, 10cyclic, 49, 159generalized, 8, 14inertial, 11modal, 80natural, 80noninertial, 11normal, 85transformation, 15velocity-, 15

coordinate system, 10Coriolis equation, 17curve parameter, 124cut principle, 4

decoupling of equation of motion, 80decoupling of equations of motion, 80degree of freedom, 12

Page 219: Introduction To Dynamics

212 Index

differential principle, 57displacement, 119

real, 9virtual, 9, 29

drinking bird, 179Duhamel integral, 97dynamic Euler equation, 37

eigen angular frequency, 75eigenfrequency, 90eigenvalue, 77

algebraic multiplicity, 92complex conjugate, 77geometric multiplicity, 92multiple, 92purely imaginary, 77simple, 83

eigenvector, 78complex conjugate, 78generalized, 94linearly independent, 92orthogonal, 91real, 78

elementary rotation, 12elimination of time, 140energy, 25

kinetic, 26, 44potential, 26, 46rigid body, 46

equilibrium point, 73elliptic, 154hyperbolic, 154

Euler, 4Euler angles, 12Euler’s theorem, 17Euler-Lagrange equations, 57excitation

harmonic, 82periodic, 82, 97

Floquet theory, 195flutter, 187force, 4

active, 4applied, 4conservative, 46constraint, 4, 9external, 4generalized, 46

internal, 4non-conservative, 71nonconservative, 46passive, 4restoring, 159surface, 4volume, 4

force field, 25irrotationality, 26rotation, 26

Fourier coefficient, 118Fourier series, 117, 146frequency response function, 82

amplitude, 98, 171phase, 98, 171

friction oscillator, 181function system, 117fundamental vibration, 80

geardifferential, 51drive, 203

Graham escapement, 188gravitational constant, 49

Hamilton, 55canonical equation, 57function, 58principle, 57

harmonic balance, 146harmonic linearization, 146Hill differential equation, 195Hurwitz determinant, 99Huygens-Steiner rule, 34hydraulic oscillator, 178

initial boundary value problem, 109initial condition, 79, 109integral of motion, 27integral principle, 57inverse kinetics, 44, 60

Jacobian, 35, 43Jeffcott rotor, 199

with stiffness asymmetries, 199Jordan block, 92jump line, 184

Kane’s baby shoe, 193Karman vortex street, 185

Page 220: Introduction To Dynamics

Index 213

Kepler’s law, 48

Lagrangecentral equation, 56equation first kind, 40equation second kind, 44function, 56multiplier, 42

Laplace transformation, 97least square, 147Lehr damping, 168Lienard-Chipart criterion, 100limit cycle, 175

stable, 177unstable, 177

linearization, 69Lyapunov, 98

direct method, 157first method, 155function, 157second method, 157

mass, 3center, 11constant, 3elastic, 4positive, 3rigid, 4

mathematical model, 59Mathieu differential equation, 195matrix

circulatory, 71damping, 71fundamental, 84gyroscopic, 71mass, 41, 71of generalized force direction, 42stiffness, 71

mechanical interaction, 24Meissner differential equation, 195method of weighted residual, 124, 144modal behavior

with multiple eigenvalues, 92without multiple eigenvalues, 83

modal matrix, 78modal transformation, 80mode shape, 78model, 1

linear continuum, 103

linear discrete, 69mathematical, 3mechanical, 3numerical, 3

moment of momentum, 24equation, 34

momentum, 23cyclic, 159equation, 34generalized, 55, 159

multibody system, 35elastic, 135

Newton-Euler equation, 33Newtonian axiom, 23norm, 118

arithmetic, 151Euclidean, 151weighted Euclidean, 151

null space matrix, 43

orientation, 11orthonormal system, 118

complete, 118oscillation

aperiodic, 85decaying, 98forced, 97nonlinear with 1 DOF, 140period of , 90periodic, 85

pendulum, 6, 43double, 86friction, 181Froude, 181physical, 153spherical, 10, 50, 72wagon, 94with driven suspension point, 197with elastic string, 204with moving suspension point, 192

pendulum clock, 188perturbation, 70

vector, 69, 152perturbed motion, 152phase

plane, 138, 143portrait, 138

piecewise solution, 143

Page 221: Introduction To Dynamics

214 Index

Poisson’s ratio, 119polynomial

characteristic, 77trigonometric, 117

position, 11potential, 26potential function, 26principle

of d’Alembert, 28of Hamilton, 55of Jourdain, 31of least action, 57of virtual work, 30

Rayleigh damping, 71Rayleigh-Ritz method, 119reference motion, 69, 152reference vector, 69relative kinematics, 16reluctance, 4residual, 124resonance, 98

combination, 192robot, 61rod on block, 141roll condition, 21rotation matrix, 12Routh-Hurwitz criterion, 99

saddle point, 42scalar product, 117separation of variable, 105, 140separatrix, 155shear modulus, 119ship

propulsion, 173turnover gear, 148

singular point, 73singularity, 18

center, 161, 163degenerated, 164focus, 162node, 162saddle, 162vortex, 162

spare coefficient, 147spectrum, 174stability, 151

and control of rotors, 202

critically stable, 101general definition, 152linear, 155linear system, 98map according to Ince/Strutt, 196mechanical system, 101of Lyapunov, 152orbit, 151state of rest, 151

state space, 72, 138state vector, 72stationary point, 73, 139stationary state, 73Stodola criterion, 99string, 115

vibration, 123strong form, 123Strouhal number, 186superposition principle, 82, 97, 137symmetry, 126system

autonomous, 156conservative, 26decoupled, 85homogeneous solution, 96linear 1st order, 83linear 2nd order, 77particular solution, 96

theoremof Dirichlet, 102of Lagrange, 102of Lyapunov, 98of Parseval, 118

torque, 48jarring direction, 198

trial function, 117admissible, 128choice, 129

ultracentrifuge, 202

variation, 9variation of constants, 96variational calculus

fundamental lemma, 34, 42variational problem, 57, 117velocity, 16

absolute, 16, 24angular, 17

Page 222: Introduction To Dynamics

Index 215

applied, 17relative, 16rotational, 17translational, 16virtual, 32

vibrationaperiodic, 170decaying, 170degree of freedom, 166excited, 170forced, 170formation principle, 166free, 167

parametric, 191periodic, 170self-excited, 175

vibration node, 91

wave equation, 116weak form, 128weighting function, 124Woodpecker toy, 180work

physical, 25virtual, 29

Young’s modulus, 119