Sie sind auf Seite 1von 676

System Dynamics

Karl A. Seeler

System Dynamics

An Introduction for Mechanical Engineers

2123
Karl A. Seeler
Mechanical Engineering Department
Lafayette College
Easton, Pennsylvania
USA

ISBN 978-1-4614-9151-4    ISBN 978-1-4614-9152-1 (eBook)


DOI 10.1007/978-1-4614-9152-1
Springer New York Heidelberg Dordrecht London

Library of Congress Control Number: 2013951813

© Springer Science+Business Media New York 2014


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the 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 information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed. Exempted from this
legal reservation are brief excerpts in connection with reviews or scholarly analysis or material supplied specifically
for the purpose of being entered and executed on a computer system, for exclusive use by the purchaser of the work.
Duplication of this publication or parts thereof is permitted only under the provisions of the Copyright Law of the
Publisher’s location, in its current version, and permission for use must always be obtained from Springer. Permissions
for use may be obtained through RightsLink at the Copyright Clearance Center. Violations are liable to prosecution
under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and
regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of publication, neither
the authors nor the editors nor the publisher can accept any legal responsibility for any errors or omissions that may be
made. The publisher makes no warranty, express or implied, with respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

I began my engineering career with a bachelor’s and master’s in civil engineering, specializing
in geotechnical engineering, primarily in foundation design and underground construction. My
five years of practice was in what is termed “heavy construction.” The scale of the structures
and force and power of the machines needed to construct them made lasting impressions. It
was my fascination with machines, such as excavators which could lift 60,000 lb. and move
with life-like dexterity, which motivated me to return to school and earn a master’s and Ph.D.
in mechanical engineering. I found myself taking graduate level courses without having their
prerequisites, relying on textbooks for the coursework I lacked. Sadly, I found many textbooks
to be incomplete, leaving the reader to fill in missing steps and to search for other texts which
covered gaps. My career led me to Lafayette College in Easton, Pennsylvania, where I have
developed a sequence of three courses: Introduction to System Dynamics, Automatic Control
Theory, and a Systems Dynamics and Controls Laboratory. Colleges have no graduate students
and, therefore, no teaching assistants. As a result, I have a more thorough teaching experience
in these subjects than is typical for a university instructor. Fortunately, Lafayette College does
have an engineering machine shop which is better equipped and staffed than those at most uni-
versities. With these resources, I have designed and worked with the machinists to build three
generations of laboratory equipment for the Systems and Controls lab.
This text is a product of my experience as a student and teacher. It is complete. There are
no missing steps in any derivation or calculation. It also covers the entire breadth of system
dynamics, so that a student who begins study of this text with a shaky understanding of or-
dinary differential equations can master the modeling and simulation of dynamic systems.
The linear graph method, a circuit-like representation of dynamic systems, is introduced and
developed. A strength of the linear graph method is the ease with which “hybrid” systems,
i.e., machinery, can be modeled. Details of machine design are added, where appropriate, to
enrich the text and motivate the student. I drew the hundreds of illustrations which appear in
the text, as an aid for visual learners like me. Example problems are incorporated throughout
the text. It is my opinion that textbooks in system dynamics have moved too far from teaching
the principles, theory, and methods of their application toward teaching the use of simulation
software. Consequently this text incorporates the use of computational software, both Mathcad
and MATLAB, but neither package’s companion simulation software.
Introduction to System Dynamics for Mechanical Engineers begins with a review of the
concepts of energy and power, and goes on to present the steps of developing an engineering
model. The mathematical background of system dynamics students varies widely. I decided
to present nearly all of the mathematics needed for the entire text in Chap. 2. Consequently,
sections of Chap. 2 should be read or assigned, as needed. The solution of ordinary differential
equations with constant coefficients, by the method of undetermined coefficients, is reviewed
in Chap. 2. However, this text will emphasize the responses of dynamic systems more than
the solutions of differential equations. As a case in point, Chap. 3 uses the example of a mass
sliding on a lubricant film to present (1) the construction of various pulse-type input functions
by superposition of scaled and time shifted Heaviside step functions, and (2) construction

v
vi Preface

of the corresponding response functions by scaling and time shifting the unit step response.
Mathcad and MATLAB are introduced to perform the superposition and plotting. The energe-
tic properties of translational and rotational mechanical elements systems, illustrated by the
step response of first-order systems, are presented in Chap. 4. Chapter 5 emphasizes the ana-
logies between fluid, electric, and thermal elements. Chapter 6 concludes the compendium of
energetic elements with transformers and transducers, i.e., levers, linkages, belt drives, chain
drives, gear sets, DC motors and generators, fluid motors and pumps, that interface similar and
dissimilar energetic subsystems.
Chapter 7 reviews the fundamentals of linear algebra, and introduces state-space. Math-
cad’s and MATLAB’s Runge–Kutta solvers are used to solve state and output equations, and
plot the responses of the higher-order systems for various inputs. Difference equations and
programming the Euler method and Runge–Kutta algorithm in MATLAB are the topics of
Chap. 8. The appendix of Chap. 8 presents the fundamentals of programming in MATLAB for
students who have not had a programming course in C or MATLAB and as a reference for tho-
se who have. Chapter 9 introduces block diagrams and block diagram algebra, the basic con-
cepts of closed-loop feedback control, the s-plane, and partial fraction expansion. Chapter 10
presents frequency response. Chapter 11 considers AC circuit analysis as a specific application
of frequency response. Chapter 11 makes use of complex impedances and phasors to apply
frequency response to electrical transformers and AC motors.
There is more material in this text than can be addressed in an introductory course in system
dynamics. Some material is included for the benefit of those students, who plan to continue
their study of system dynamics and controls on the graduate level. Other material is included
to provide depth and completeness in specific topics, to allow instructors to tailor their indivi-
dual courses to suit their curriculum and their students’ preparation. The application of system
dynamics to machine design is emphasized throughout, to motivate students and strengthen
their ability to develop engineering models of real systems. Whether you are intrigued broadly
by the flow of energy in physical systems, or need to model a specific system for a particular
purpose, you will find this book useful.
I close this preface by acknowledging my debt to my family to whom I dedicate this text.
I express my gratitude to my wife, Rani, for her love, support, and editing; to our sons, Adim
and Felix, for their patience and cooperation; to our four cats, Mack, Lily, Fluffy, and Zelda,
for their late night company, but not for their traffic across the keyboard, the text of which I
hope I have removed; and lastly, our hound, Juno, who contributed to the preparation of this
book with a cold nose to my mouse arm.

Lafayette College Karl A. Seeler, Ph.D., P.E.


Contents

1  Introduction to System Dynamics�����������������������������������������������������������������������������   1


1.1 Introduction ���������������������������������������������������������������������������������������������������������    1
1.1.1 Why Study System Dynamics? ��������������������������������������������������������������    2
1.1.2 Engineering Models �������������������������������������������������������������������������������    4
1.1.3 Mathematical Models �����������������������������������������������������������������������������    6
1.1.4 Extraction of Mathematical Statements from Engineering Models �������    6
1.1.5 Formulation of a Mathematical Model from an Engineering
Model �����������������������������������������������������������������������������������������������������    7
1.1.6 Summary of Engineering Modeling and Analysis ���������������������������������    7
1.2 Mechanical and Energetic Models ���������������������������������������������������������������������    8
1.3 Energy, Mechanical Power, and Coenergy ���������������������������������������������������������  11
1.3.1 Strain Energy �����������������������������������������������������������������������������������������  11
1.3.2 Mechanical Power ���������������������������������������������������������������������������������  11
1.3.3 Strain Coenergy �������������������������������������������������������������������������������������  12
1.3.4 Energy Density ���������������������������������������������������������������������������������������  14
1.3.5 Kinetic Energy ���������������������������������������������������������������������������������������  14
1.3.6 Power Flows and Signs ���������������������������������������������������������������������������  16
1.3.7 Power Sources ���������������������������������������������������������������������������������������  16
1.4 Network Representation of Energy Flow �����������������������������������������������������������  19
1.4.1 Compatibility and Continuity Equations �����������������������������������������������  20
1.4.2 Compatibility Equations �������������������������������������������������������������������������  21
1.4.3 Continuity Equations �����������������������������������������������������������������������������  24
1.4.4 Summary of Compatibility and Continuity Equations ���������������������������  26
1.5 Overview of Engineering Modeling and Analyses ���������������������������������������������  26
1.5.1 Engineering Modeling and Analysis Process �����������������������������������������  27
1.5.2 Engineering Modeling and Analysis Examples �������������������������������������  31
Summary ���������������������������������������������������������������������������������������������������������������������  39
Problems ���������������������������������������������������������������������������������������������������������������������  40
References and Suggested Reading �����������������������������������������������������������������������������  44
2 Differential Equations, Input Functions, Complex Exponentials,
and Transfer Functions����������������������������������������������������������������������������������������������  45
2.1 Introduction ���������������������������������������������������������������������������������������������������������  45
2.2  Input Functions ���������������������������������������������������������������������������������������������������  46
2.2.1  Power Sources ���������������������������������������������������������������������������������������  46
2.2.2  Heaviside Unit Step Function �����������������������������������������������������������������  47
2.2.3  Unit Impulse �������������������������������������������������������������������������������������������  48
2.2.4  Unit Ramp �����������������������������������������������������������������������������������������������  48
2.2.5 Sinusoids �������������������������������������������������������������������������������������������������  49
2.2.6  Step Responses as Input Functions ���������������������������������������������������������  49
2.3 Linearity �������������������������������������������������������������������������������������������������������������  49

vii
viii Contents

2.4 Superposition ���������������������������������������������������������������������������������������������������    51
2.5  Method of Undetermined Coefficients �������������������������������������������������������������    51
2.6  Initial Conditions ���������������������������������������������������������������������������������������������    59
2.7  Complex Numbers and Variables ���������������������������������������������������������������������    60
2.7.1  Cartesian Form  Z = x + j y ���������������������������������������������������������������    60
2.7.2  Polar Form: Z = Z  Z �����������������������������������������������������������������������    63
2.7.3  Euler’s Equations ���������������������������������������������������������������������������������    64
2.7.4  Complex Exponential Form Z e jφ �������������������������������������������������������    65
2.7.5  Rotating Complex Exponential Unit Vector �����������������������������������������    68
2.7.6 Example: Solution of a Homogeneous Equation using Complex
Exponential Unit Vectors ���������������������������������������������������������������������    71
2.8 Solved Problems Illustrating the Method of Undetermined Coefficients ���������    74
2.8.1 Example Problem One: First-Order Step Response �����������������������������    74
2.8.2 Example Problem Two: Non-Oscillatory Second-Order
Step Response ���������������������������������������������������������������������������������������    77
2.8.3 Example Problem Three: Oscillatory Second-Order
Step Response ���������������������������������������������������������������������������������������    79
2.9 Eigenvalues and Response Characterization ���������������������������������������������������    84
2.9.1  First-Order Step Responses �����������������������������������������������������������������    84
2.9.2 Non-Oscillatory Second-Order Step Responses �����������������������������������    86
2.9.3  Oscillatory Second-Order Step Responses �������������������������������������������    88
2.9.4  Time Step for Response Calculations ���������������������������������������������������    89
2.10 Laplace Transformation and Transfer Functions ���������������������������������������������    90
2.10.1  Laplace Transformation ���������������������������������������������������������������������    90
2.10.2  The Inverse Laplace Transformation �������������������������������������������������    92
2.10.3  Final Value and Initial Value Theorems ���������������������������������������������    93
2.10.4  Transfer Functions �����������������������������������������������������������������������������    93
2.10.5  Partial Fraction Expansion �����������������������������������������������������������������    95
Problems �������������������������������������������������������������������������������������������������������������������  104
Chapter 2 Appendix ���������������������������������������������������������������������������������������������������  105
Table 2.3 Laplace Transform Pairs �������������������������������������������������������������������  105
Mathcad and MATLAB ���������������������������������������������������������������������������������������������  106
Plotting in Mathcad �������������������������������������������������������������������������������������������  106
Plotting in MATLAB ���������������������������������������������������������������������������������������  110
References and Suggested Reading ���������������������������������������������������������������������������  115
3 Introduction to the Linear Graph Method, Step Responses,
and Superposition�����������������������������������������������������������������������������������������������������  117
3.1 Introduction �������������������������������������������������������������������������������������������������������  117
3.2  Introduction to the Linear Graph Method ���������������������������������������������������������  117
3.2.1  Energetic Model �����������������������������������������������������������������������������������  118
3.2.2 Newtonian Formulation of the Force-Mass-Damper Model ���������������  121
3.2.3 Linear Graph Formulation of the Force-Mass-Damper Model �����������  121
3.2.4 Examples Illustrating the Linear Graph Method ���������������������������������  129
3.2.5 Summary of the Introduction to Linear Graphs �����������������������������������  135
3.3  The Heaviside Unit Step Function �������������������������������������������������������������������  136
3.3.1 Differentiation of the Heaviside Unit Step Function ���������������������������  137
3.4  Initial Conditions ���������������������������������������������������������������������������������������������  138
3.4.1  Initial Condition, First-Order System ���������������������������������������������������  139
3.4.2  Initial Conditions: Second-Order System ���������������������������������������������  140
3.5  First-Order Step Responses �������������������������������������������������������������������������������  142
3.5.1 Categorization of First-Order Step Responses �������������������������������������  144
3.5.2 Step Response of a First-Order Mechanical System ���������������������������  145
3.6  Time Shift �������������������������������������������������������������������������������������������������������    147
3.7 Superposition of Heaviside Step Functions �����������������������������������������������������  148
3.8 Superposition of First-Order Step Responses ���������������������������������������������������  149
3.8.1 Scaling, Time Shifting, and Superposing Unit Step Responses �����������  151
Contents ix

3.9 Superposition of Second-Order Step Responses �����������������������������������������������  154


3.9.1 Overdamped or Non-Oscillatory Step Response ���������������������������������  154
3.9.2 Underdamped or Oscillatory Step Response ���������������������������������������  156
3.10  Initial Condition of Energized Systems �����������������������������������������������������������  161
3.10.1 Example of Second-Order Pulse Response
of an Energized System ���������������������������������������������������������������������  162
3.10.2  Initial Value Method ���������������������������������������������������������������������������  165
3.11  Solved Problems �����������������������������������������������������������������������������������������������  169
3.11.1 Step Responses of Initially De-energized System �����������������������������  169
3.11.2 Step Responses of Initially Energized System �����������������������������������  171
3.11.3 Second-Order Step Responses and Pulse Responses �������������������������  174
Summary �������������������������������������������������������������������������������������������������������������������  182
Problems �������������������������������������������������������������������������������������������������������������������  190
Chapter 3 Appendix ���������������������������������������������������������������������������������������������������  197
Mathcad: Plotting Superposed Functions ���������������������������������������������������������  190
MATLAB: Plotting Superposed Functions �������������������������������������������������������  190
Time Shift �����������������������������������������������������������������������������������������������������������������  191
Superposition Using Nested Loops ���������������������������������������������������������������������������  191
References and Suggested Reading ���������������������������������������������������������������������������  193
4  Mechanical Systems �������������������������������������������������������������������������������������������������  196
4.1 Translational Mechanical System Elements �����������������������������������������������������  197
4.2  Modeling Translational Elements ���������������������������������������������������������������������  198
4.2.1  Mass, Kinetic Energy Storage Element �����������������������������������������������  199
4.2.2  Spring, Strain Energy Storage Element �����������������������������������������������  200
4.2.3  Effective Mass �������������������������������������������������������������������������������������  207
4.2.4 Damper: Viscous Friction Energy Dissipation �������������������������������������  210
4.2.5  Translational Mechanical Sources �������������������������������������������������������  213
4.2.6 Summary �����������������������������������������������������������������������������������������������  216
4.3  The Sign Problem of Mechanical Systems �������������������������������������������������������  216
4.4 Drawing Linear Graphs from Mechanical Schematics �������������������������������������  219
4.4.1  Linear Graph Symbols �������������������������������������������������������������������������  219
4.4.2 Force Source Acting on a Parallel Mass-Damper System �������������������  219
4.4.3 Force Source Acting on a System of a Mass and Two Dampers ���������  221
4.4.4 Force Source Acting on System of a Mass and Two Dampers �������������  221
4.4.5 Force Source Acting on a Mass-Spring-Damper System ���������������������  221
4.4.6  Viscoelastic Models �����������������������������������������������������������������������������  221
4.5  Rotational Mechanical System Elements ���������������������������������������������������������  223
4.6  Modeling Rotational Elements �������������������������������������������������������������������������  223
4.6.1  Mass Moment of Inertia �����������������������������������������������������������������������  223
4.6.2 Mass Moment of Inertia of Primitive Shapes ���������������������������������������  226
4.6.3 Mass Moment of Inertia Calculated from Area Moment of Inertia �����  226
4.6.4 Mass Moment of Inertia Calculated by Superposition �������������������������  226
4.6.5  Torsion Springs �������������������������������������������������������������������������������������  232
4.6.6  Rotational Damping �����������������������������������������������������������������������������  234
4.6.7  Rotational System Sources �������������������������������������������������������������������  236
4.7  Dynamic Tests ���������������������������������������������������������������������������������������������������  237
4.7.1 Components with a Single Unknown Energetic Parameter �����������������  237
4.7.2 Components with Multiple Energetic Parameters �������������������������������  238
4.7.3  First-Order System Step Responses �����������������������������������������������������  239
4.7.4  Second-Order System Step Responses �������������������������������������������������  247
4.7.5  Higher-Order System Responses ���������������������������������������������������������  254
4.8  Equivalent Elements �����������������������������������������������������������������������������������������  257
4.8.1  Dampers in Parallel �������������������������������������������������������������������������������  258
4.8.2  Dampers in Series ���������������������������������������������������������������������������������  258
4.8.3  Springs in Parallel ���������������������������������������������������������������������������������  258
x Contents

4.8.4  Springs in Series �����������������������������������������������������������������������������������  257


4.8.5 Equivalent Mass and Mass Moment of Inertia �������������������������������������  258
Summary �������������������������������������������������������������������������������������������������������������������  259
Problems �������������������������������������������������������������������������������������������������������������������  259
References and Suggested Reading ���������������������������������������������������������������������������  268
5  Fluid, Electrical, and Thermal Systems �����������������������������������������������������������������  269
5.1  Fluid Systems ���������������������������������������������������������������������������������������������������  269
5.1.1  Fluid Power Variables ���������������������������������������������������������������������������  269
5.2  Fluid Elemental and Energy Equations �������������������������������������������������������������  271
5.2.1  Fluid Energy Dissipation: Fluid Resistance �����������������������������������������  271
5.2.2  Kinetic Energy Storage Fluid Inertance �����������������������������������������������  271
5.2.3 Pressure-Based Energy Storage Fluid Capacitance �����������������������������  272
5.2.4  Fluid System Sources ���������������������������������������������������������������������������  275
5.3  Linear Graphs of Fluid Systems �����������������������������������������������������������������������  277
5.3.1  Example Fluid System Linear Graphs �������������������������������������������������  277
5.4 Calculating Fluid Element Parameters from Fluid Properties
and Geometry ���������������������������������������������������������������������������������������������������  282
5.4.1  Fluid Resistance �����������������������������������������������������������������������������������  282
5.4.2  Fluid Inertance �������������������������������������������������������������������������������������  283
5.4.3 Fluid Capacitance (Hydraulic Accumulators) �������������������������������������  284
5.5 Fluid Power System Hardware and Symbols ���������������������������������������������������  284
5.5.1  Metering or Flow Control Valves ���������������������������������������������������������  284
5.5.2  Check Valves ����������������������������������������������������������������������������������������  285
5.5.3  Multi Position Shuttle or Spool Valves �������������������������������������������������  285
5.6  Electrical Systems ���������������������������������������������������������������������������������������������  286
5.6.1 Analogies Between Fluid and Electrical Systems �������������������������������  286
5.6.2  Summary of Electromagnetic Phenomena �������������������������������������������  286
5.6.3  Electrical Units �������������������������������������������������������������������������������������  288
5.7  Electrical System Elements �������������������������������������������������������������������������������  288
5.7.1  Electrical Resistance �����������������������������������������������������������������������������  288
5.7.2 Electrical Capacitance: Energy Stored in an Electric Field �����������������  290
5.7.3 Electrical Inductance: Energy Stored in a Magnetic field �������������������  291
5.7.4  Electrical Systems ���������������������������������������������������������������������������������  293
5.8  Thermal Systems �����������������������������������������������������������������������������������������������  294
5.8.1  Thermal Power Variables ���������������������������������������������������������������������  294
5.8.2 Modes of Heat Transfer and Their Corresponding
Thermal Resistances �����������������������������������������������������������������������������  295
5.8.3  Thermal System Elements �������������������������������������������������������������������  297
5.8.4  Thermal Systems ����������������������������������������������������������������������������������  299
5.9 Equivalent Elements in Fluid, Electrical, and Thermal Systems ���������������������  300
5.9.1  Fluid, Electrical, or Thermal Resistances ���������������������������������������������  301
5.9.2  Fluid and Electrical Capacitance ���������������������������������������������������������  302
5.9.3  Fluid Inertance or Electrical Inductance �����������������������������������������������  303
5.9.4 Fluid Inertances or Electrical Inductances In Series �����������������������������  304
Summary �������������������������������������������������������������������������������������������������������������������  304
Problems �������������������������������������������������������������������������������������������������������������������  305
Chapter 5 Appendix ���������������������������������������������������������������������������������������������������  312
Engineering Electromagnetics �������������������������������������������������������������������������  312
Electromagnetic Force �������������������������������������������������������������������������������������  312
Electromagnetic Force between Two Current Elements �����������������������������������  314
The Magnetic Field B ���������������������������������������������������������������������������������������  314
Coils or “Solenoids” �����������������������������������������������������������������������������������������  318
Magnetic Moment and Engineering Approximations of Magnetic Field
Density �������������������������������������������������������������������������������������������������������������  318
Magnetic Permeability μ and Ferromagnetic Materials �����������������������������������  319
Contents xi

 agnetic Flux Density B, Flux φ, and Applied Magnetic


M
Field Intensity H �����������������������������������������������������������������������������������������������  322
Magnetic Circuit Model �����������������������������������������������������������������������������������  323
References and Suggested Reading ���������������������������������������������������������������������������  331
6  Power Transmission, Transformation, and Conversion ���������������������������������������  333
6.1 Introduction to Power Transmission, Transformation, and Conversion �����������  333
6.1.1 Transformers �����������������������������������������������������������������������������������������  333
6.1.2  Ideal Transformers �������������������������������������������������������������������������������  334
6.1.3 Transducers �������������������������������������������������������������������������������������������  336
6.1.4  Ideal Transducers ���������������������������������������������������������������������������������  336
6.1.5  Block Model of Power Flows ���������������������������������������������������������������  336
6.1.6  Transformer and Transducer Equations �����������������������������������������������  337
6.1.7 Transformer and Transducers Linear Graph Symbol ���������������������������  337
6.2 Transformers �����������������������������������������������������������������������������������������������������  338
6.2.1  Mechanical Transformers Levers ���������������������������������������������������������  338
6.2.2 Gears Mechanical Transformers and Transducers �������������������������������  341
6.3 Transformer and Transducers Sign Conventions ���������������������������������������������  346
6.3.1 Example 1: The Linear Graph for a Rotational System
with a Transformer. �������������������������������������������������������������������������������  348
6.4 Transducers �������������������������������������������������������������������������������������������������������  350
6.4.1 DC Electric Motors: Electrical to Mechanical Transducers �����������������  350
6.4.2 Generators: Mechanical to Electrical Transducers �������������������������������  354
6.4.3  Pumps: Mechanical to Fluid Transducers ���������������������������������������������  354
6.4.4 Hydraulic Motors Fluid to Mechanical Transducers ���������������������������  356
6.4.5 Linear Hydraulic Motor or Hydraulic Piston-Cylinder �����������������������  356
6.4.6  Rotational Hydraulic Motors ���������������������������������������������������������������  358
6.4.7 Example 3: Linear Graph of a Fourth-Order Fluid-Mechanical
System ���������������������������������������������������������������������������������������������������  359
6.4.8 Rotational to Translational Mechanical Transducers ���������������������������  360
6.4.9 Translational to Rotational Mechanical Transducers ���������������������������  362
6.5  Multiport Transformers and Transducers ���������������������������������������������������������  362
6.5.1 Example 1: A Lever with Three Attachments ���������������������������������������  363
6.5.2  Example 2: A Pinion Driving Two Gears ���������������������������������������������  365
6.5.3  Example 3: A Belt Driving Two Pulleys �����������������������������������������������  367
6.6 Floating Sources, Transformers, and Transducers �������������������������������������������  368
6.6.1  Floating Mechanical Sources ���������������������������������������������������������������  368
6.6.2  Floating and Multiple Fluid Sources ���������������������������������������������������  369
6.6.3  Floating and Multiple Electrical Sources ���������������������������������������������  369
6.6.4  Floating Transformers and Transducers �����������������������������������������������  373
6.7 Equivalent Elements in Systems with Transformers and Transducers �������������  374
6.7.1 Equivalent Elements in a System with a Transformer �������������������������  374
6.8  Example Problems �������������������������������������������������������������������������������������������  377
6.8.1 Example Problem 1: Linear Graph of a Hybrid Rotational
Translational System �����������������������������������������������������������������������������  377
6.8.2 Example Problem 2: A Pivoted Beam (Lever) Acting on Three
Elements �����������������������������������������������������������������������������������������������  379
6.8.3 Example Problem 3: A Serpentine Belt Driving Two Elements �����������  384
6.8.4 Example Problem 4: Rotational System with Compound Gears ���������  388
6.8.5 Example Problem 5: Hybrid Electric, Rotational, and Translational
System ���������������������������������������������������������������������������������������������������  391
6.8.6 Example Problem 6: Hybrid Rotational and Fluid System �����������������  394
Summary �������������������������������������������������������������������������������������������������������������������  399
Problems �������������������������������������������������������������������������������������������������������������������  400
References and Suggested Reading ���������������������������������������������������������������������������  409
xii Contents

7 Vector-Matrix Algebra and the State-Space Representation


of Dynamic Systems �������������������������������������������������������������������������������������������������  411
7.1 Overview �����������������������������������������������������������������������������������������������������������  411
7.2  Vector-Matrix Algebra �������������������������������������������������������������������������������������  411
7.2.1  Matrix Addition �����������������������������������������������������������������������������������  412
7.2.2  Matrix Multiplication ���������������������������������������������������������������������������  413
7.3 Operating on a Vector-Matrix Expression with a Linear Operator �������������������  414
7.3.1 Laplace Transformation of Matrix or a Vector-Matrix Expression �����  415
7.4  Transpose of a Matrix ���������������������������������������������������������������������������������������  415
7.5  Matrix Inversion �����������������������������������������������������������������������������������������������  416
7.5.1  Calculation of an Inverse Matrix ���������������������������������������������������������  418
7.5.2  Determinant of a Matrix �����������������������������������������������������������������������  418
7.5.3  Cofactor of a Matrix �����������������������������������������������������������������������������  420
7.5.4  Adjoint of a Matrix �������������������������������������������������������������������������������  421
7.6 State-Space Representation of Dynamic Systems �������������������������������������������  422
7.6.1  State Variables ��������������������������������������������������������������������������������������  422
7.6.2 Example Second-Order Dynamic System RLC Circuit �����������������������  423
7.6.3  State Equations �������������������������������������������������������������������������������������  424
7.6.4  Output Equations ���������������������������������������������������������������������������������  425
7.6.5  Vector-Matrix Form of the State Equations �����������������������������������������  426
7.6.6 Vector-Matrix Form of the Output Equations ���������������������������������������  427
7.6.7 Numerical Solution of the State and Output Equations �����������������������  427
7.7 Example Derivations of State and Output Equations ���������������������������������������  428
7.7.1 Third-Order Dynamic System Example: A Rotational Mechanical
System ���������������������������������������������������������������������������������������������������  428
7.7.2 Fourth-Order Dynamic System Example: A Spring-Mass-Damper
System ���������������������������������������������������������������������������������������������������  430
7.7.3 Fourth-Order Dynamic System Example: A Fluid-Mechanical
System ���������������������������������������������������������������������������������������������������  433
7.8  Why “State-Space” is called “State-Space” �����������������������������������������������������  437
7.9 Expression of Systems Equations in State-Space ���������������������������������������������  438
7.9.1 Algorithm to Express a Higher-Order System Equation
Without Differentiation of the Input as State Equations �����������������������  438
7.9.2 Algorithm to Express a Higher-Order System Equation
with Differentiation of the Input as State Equations �����������������������������  440
7.10  Eigenvalues and Eigenvectors �����������������������������������������������������������������������  443
7.10.1 Eigenvalues �����������������������������������������������������������������������������������������  443
7.10.2 Eigenvectors ���������������������������������������������������������������������������������������  444
Summary �������������������������������������������������������������������������������������������������������������������  444
Problems �������������������������������������������������������������������������������������������������������������������  445
Chapter 7 Appendix ���������������������������������������������������������������������������������������������������  457
Mathcad’s Runge–Kutta Solver �����������������������������������������������������������������������  457
Step Response of a Linear Mass-Damper System �������������������������������������������  457
Step Response of a Mass-Damper System with a Non-Linear Damper �����������  458
Response of a Linear Mass-Damper System Subjected to a Pulse Train ���������  459
Response of a Mass-Damper System to Sinusoidal Inputs �����������������������������  459
Step Response of a Spring-Mass-Damper System �������������������������������������������  460
MATLAB’s Runge–Kutta Solver ode45() �����������������������������������������������������������������  461
Step Response of a Linear Mass-Damper System �������������������������������������������  462
Step Response of a Non-Linear Mass-Damper System �����������������������������������  462
Response of a Linear Mass-Damper System Subjected to a Pulse Train ���������  463
Response of a Mass-Damper System to Sinusoidal Inputs �����������������������������  463
Step Response of a Spring-Mass-Damper System �������������������������������������������  464
References and Recommended Reading �������������������������������������������������������������������  465
Contents xiii

8  Finite Difference Methods and MATLAB �������������������������������������������������������������  467


8.1 Finite Difference Approximation of Differential Equations �����������������������������  467
8.2 Euler Method, Forward Stepping, Finite Difference Algorithm �����������������������  468
8.2.1 MATLAB Programming of the Euler Method, First-Order System �����  469
8.2.2 Euler Method Solution of Second-Order State Equations �������������������  471
8.2.3 Example: Euler Method Solver, Non-Linear State Equation ���������������  473
8.3  User-Written MATLAB Functions �������������������������������������������������������������������  475
8.3.1  Static-Kinetic Coulomb Friction Model �����������������������������������������������  475
8.3.2  Programming a Function in MATLAB �������������������������������������������������  476
8.4  Runge–Kutta Method ���������������������������������������������������������������������������������������  479
8.4.1 Two-State, Fourth-Order Runge–Kutta Algorithm and Code ���������������  479
8.4.2 Three-State, Fourth-Order Runge–Kutta Algorithm and Code �����������  480
8.5 Programming Non-Linearities and Input Functions �����������������������������������������  481
8.5.1  Common Non-Linearities ���������������������������������������������������������������������  482
8.5.2 Input Functions, Non-Linearities and the Runge–Kutta Algorithm �����  483
8.5.3  Trapezoidal Integration �������������������������������������������������������������������������  484
Summary �������������������������������������������������������������������������������������������������������������������  484
Problems �������������������������������������������������������������������������������������������������������������������  485
Chapter 8 Appendix ���������������������������������������������������������������������������������������������������  493
Introduction to Programming and MATLAB ���������������������������������������������������  493
A Brief History and Classification of Computer Programming Languages �����  493
Fundamentals of Procedural Programming �����������������������������������������������������  494
A Brief History of Computer Memory �������������������������������������������������������������  495
Base Conversion �����������������������������������������������������������������������������������������������  497
Computational Error on Conversion from Natural
Decimal to Natural Binary Fractions ���������������������������������������������������������������  498
Data Types �����������������������������������������������������������������������������������������������������������������  499
Integer and Signed Integer Variables ���������������������������������������������������������������  499
Floating Point Variables �����������������������������������������������������������������������������������  500
Boolean or Logical Variables ���������������������������������������������������������������������������  500
Procedural Logic and Flow charts �����������������������������������������������������������������������������  500
Logic Loop �������������������������������������������������������������������������������������������������������  501
Nested Loops ���������������������������������������������������������������������������������������������������  501
Flowchart Rules and Guidelines �����������������������������������������������������������������������  501
MATLAB �������������������������������������������������������������������������������������������������������������������  503
MATLAB “Environment” �������������������������������������������������������������������������������  503
Variables �����������������������������������������������������������������������������������������������������������  504
Scalar, Matrix, and Array Variables �����������������������������������������������������������������  505
Programming Statement Syntax �����������������������������������������������������������������������  509
Assignment Statements �������������������������������������������������������������������������������������  509
Control Flow Statements ���������������������������������������������������������������������������������  509
Comments ���������������������������������������������������������������������������������������������������������  511
plot Statement ���������������������������������������������������������������������������������������������������  511
Programming a Function in MATLAB �������������������������������������������������������������  513
Reading From and Writing To Files �����������������������������������������������������������������  513
MATLAB’s step() and impulse() Functions �����������������������������������������������������  515
Vector Calculations in MATLAB ���������������������������������������������������������������������  516
References and Suggested Reading ���������������������������������������������������������������������������  517
9  Transfer Functions, Block Diagrams, and the s-Plane �����������������������������������������  519
9.1  Linear Operators and Transfer Functions ���������������������������������������������������������  519
9.1.1  Linear Operators �����������������������������������������������������������������������������������  519
9.1.2  Properties of Linear Operators �������������������������������������������������������������  520
9.1.3  Incrementally Linear Functions �����������������������������������������������������������  521
9.1.4 Differential Equations and Transfer Functions as Linear Operators�����  521
9.2 Laplace-Domain Solution of a Set of State and Output Equations �������������������  522
xiv Contents

9.3  Block Diagrams �����������������������������������������������������������������������������������������������  526


9.3.1  A Block �������������������������������������������������������������������������������������������������  527
9.3.2  Cascaded Blocks �����������������������������������������������������������������������������������  527
9.3.3 Differentiation, Integration, and Transfer Functions Blocks ���������������  528
9.3.4  Summation Junctions ���������������������������������������������������������������������������  529
9.3.5  Branch Points ���������������������������������������������������������������������������������������  529
9.3.6  Block Diagram Algebra �����������������������������������������������������������������������  529
9.3.7  Feedforward and Feedback Loops �������������������������������������������������������  530
9.4 Time-Domain Block Diagrams of Differential System Equations �������������������  535
9.4.1 Block Diagram Without Differentiation of the Input ���������������������������  535
9.4.2 Block Diagram with Differentiation of the Input ���������������������������������  537
9.5 State Equations as a Time-Domain Block Diagram �����������������������������������������  542
9.5.1 Drawing a Block Diagram of Existing State Equation �������������������������  542
9.5.2 Drawing a Block Diagram from the Energetic Equations �������������������  542
9.5.3 Drawing a Block Diagram of State Equations from
a System Equation �������������������������������������������������������������������������������  543
9.6  The s-Plane �������������������������������������������������������������������������������������������������������  545
9.6.1 Poles, Zeros, and Pole-Zero Transfer Function Form ���������������������������  546
9.6.2 Stability �������������������������������������������������������������������������������������������������  547
9.6.3  Real Component σ, the Decay Rate �����������������������������������������������������  549
9.6.4 Imaginary Component ω, the Observed, Damped Frequency �������������  549
9.6.5  Damping Ratio ζ �����������������������������������������������������������������������������������  549
9.6.6 s-Plane Plots of Transfer Function Poles and Zeros �����������������������������  551
9.6.7  Example Pole-Zero Plots ���������������������������������������������������������������������  551
Summary �������������������������������������������������������������������������������������������������������������������  552
Linear Operators �����������������������������������������������������������������������������������������������  552
Time-Domain Block Diagrams �������������������������������������������������������������������������  553
Transfer Functions and Laplace-Domain Block Diagrams �����������������������������  553
s-plane ���������������������������������������������������������������������������������������������������������������  553
Problems �������������������������������������������������������������������������������������������������������������������  553
Chapter 9 Appendix ���������������������������������������������������������������������������������������������������  560
Inverse Laplace Transformation Using Manual Partial Fraction Expansion �����  560
References and Suggested Reading ���������������������������������������������������������������������������  576
10  Frequency Response�������������������������������������������������������������������������������������������������  577
10.1 Overview of Sinusoidal Excitation and Frequency Response �������������������������  577
10.1.1 Calculating Magnitude and Phase Angle from a Response ���������������  580
10.1.2 Transient and Steady-State Response of a First-Order System ���������  582
10.1.3 Transient and Steady-State Response of a Second-Order System �����  584
10.2  Frequency Response Relationship �������������������������������������������������������������������  588
10.2.1 Example Frequency Response Calculation ���������������������������������������  588
10.3 Derivation of the Frequency Response Equation ���������������������������������������������  590
10.4 Fourier Series Approximation of Periodic Signals �������������������������������������������  593
10.5  Bode Plots ���������������������������������������������������������������������������������������������������������  595
10.5.1  Review of the Properties of Logarithms ���������������������������������������������  595
10.5.2 Decibels ���������������������������������������������������������������������������������������������  596
10.5.3  Interpolating on a Logarithmic Scale �������������������������������������������������  597
10.5.4  Log-Magnitude Bode Plots �����������������������������������������������������������������  597
10.5.5  Phase-Angle Bode Plots ���������������������������������������������������������������������  602
10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot �������������������������  605
10.6.1  Asymptotic Approximation Form �������������������������������������������������������  607
10.6.2 Asymptotic Approximation of First-Order Factors �����������������������������  608
10.6.3 Asymptotic Approximation of Underdamped
Second-Order Factors �������������������������������������������������������������������������  610
10.6.4  Integrator and Differentiator Factors �������������������������������������������������  613
Contents xv

10.6.5  Gain Factors ���������������������������������������������������������������������������������������  614


10.6.6 Sketching Asymptotic Approximation Log-Magnitude
Bode Plots �������������������������������������������������������������������������������������������  615
10.6.7 Bode Phase-Angle Plots of Transfer Functions ���������������������������������  620
10.6.8  A Complex Number’s Argument �������������������������������������������������������  623
10.7  Nyquist Polar Plots �������������������������������������������������������������������������������������������  623
Summary �������������������������������������������������������������������������������������������������������������������  625
Problems �������������������������������������������������������������������������������������������������������������������  626
Chapter 10 Appendix �������������������������������������������������������������������������������������������������  630
Drawing Bode Plots in Mathcad �����������������������������������������������������������������������  630
Drawing Nyquist Plots in Mathcad �������������������������������������������������������������������  632
Drawing Bode Plots in MATLAB �������������������������������������������������������������������  632
Drawing Nyquist Plots in MATLAB ���������������������������������������������������������������  634
References and Suggested Reading ���������������������������������������������������������������������������  634
11  AC Circuits and Motors�������������������������������������������������������������������������������������������  635
11.1 Introduction �������������������������������������������������������������������������������������������������������  635
11.1.1 Alternating Current �����������������������������������������������������������������������������  635
11.1.2 AC Power �������������������������������������������������������������������������������������������  636
11.1.3 Root-Mean-Square (RMS) or Effective Values ���������������������������������  636
11.1.4 Sinusoidal and RMS Voltages �������������������������������������������������������������  637
11.1.5 Three-Phase Alternating Current �������������������������������������������������������  637
11.2 Frequency Response of Electric Circuits ���������������������������������������������������������  638
11.3 Complex Impedance �����������������������������������������������������������������������������������������  639
11.3.1 Complex Impedance of a Resistor �����������������������������������������������������  639
11.3.2 Complex Impedance of a Capacitor ���������������������������������������������������  640
11.3.3 Complex Impedance of an Inductor ���������������������������������������������������  640
11.3.4 Complex Admittance �������������������������������������������������������������������������  640
11.3.5 Reduction of the RLC Circuit Using Complex Impedances ���������������  640
11.3.6 Driving Point Impedance �������������������������������������������������������������������  642
11.3.7 Graphical Reduction of Networks of Complex Impedances �������������  642
11.4  Phasors and Phasor Operators ���������������������������������������������������������������������������  642
11.4.1 Reactance and Resistance �������������������������������������������������������������������  643
11.5 Electrical Transformers �������������������������������������������������������������������������������������  644
11.5.1 Model of a Transformer with a Resistive and Capacitive Load ���������  644
11.6 Three-Phase Power �������������������������������������������������������������������������������������������  645
11.6.1 Line-to-Line Voltage ���������������������������������������������������������������������������  646
11.7 Physical Principles of Three-Phase AC Motors �����������������������������������������������  647
11.7.1 Rotating Magnetic Vector �������������������������������������������������������������������  647
11.7.2 Three-Phase Synchronous Motors �����������������������������������������������������  648
11.7.3 Three-Phase Induction Motors �����������������������������������������������������������  648
11.7.4 Variable Frequency Motors �����������������������������������������������������������������  649
11.8 Three-Phase AC Circuits ���������������������������������������������������������������������������������  649
11.8.1 Wye (Y) and Delta (Δ) Three-Phase Connections ����������������������������  649
11.8.2 Wye Connected AC Machines �����������������������������������������������������������  650
11.8.3 Delta Connected Machines ���������������������������������������������������������������  650
11.8.4 Example 1 Three-Phase Delta Connected Motor Phase Voltage �������  651
11.8.5 Example 2 Three-Phase Delta Connected Motor Line Voltage ���������  651
11.9 AC Power Calculation �������������������������������������������������������������������������������������  652
11.9.1 Example AC Power Calculations Using Motor Specifications ���������  653
11.10 Single-Phase AC Motors ���������������������������������������������������������������������������������  655
11.10.1 Capacitor Start ���������������������������������������������������������������������������������  655
11.10.2 Series Wound �����������������������������������������������������������������������������������  655
11.11 Mechanical Design Considerations �����������������������������������������������������������������  656
xvi Contents

11.11.1 NEMA ���������������������������������������������������������������������������������������������  656


11.11.2 US Department of Energy Efficiency Data �������������������������������������  656
Summary �������������������������������������������������������������������������������������������������������������������  656
Problems �������������������������������������������������������������������������������������������������������������������  656
Chapter 11 Appendix �������������������������������������������������������������������������������������������������  657
Evaluation of the Root-Mean-Squared Integral ���������������������������������������������  657
Reference and Suggested Reading �����������������������������������������������������������������������������  659
Index���������������������������������������������������������������������������������������������������������������������������������  661
Introduction to System Dynamics
1

Abstract
System Dynamics is the study of the change of the power variables within an energetic
system. As a system interacts with a power source, energy flows across the system bound-
ary, between storage modes within the system, and is dissipated as heat by friction, electri-
cal resistance, or magnetic hysteresis. Modeling energetic systems as circuit-like networks
simplifies the accounting of power flows in the system. Creation of a mathematical model
of an energetic system represented as a network is straightforward elimination by substitu-
tion, guided by a drawing of the network, once a complete set of mathematical statements
of the physical truths of network-like system has been written.

1.1 Introduction vibrate, energy flows back and forth between the kinetic en-
ergy of motion and the potential energy of elastic deforma-
Webster’s New World Dictionary defines “dynamic” as “re- tion. What we will model as energy storage elements are,
lating to energy or a physical force in motion: opposed to in fact, properties of physical objects. Often, a single physi-
static”…“relating to or tending toward change.” “System” cal object will have multiple “energetic” properties. We will
is defined as “a set or arrangement of things so related or represent each property in our system models as a different
connected as to form a unity or organic whole: as a solar sys- element, Fig. 1.1.
tem, irrigation system, supply system.” Our focus will be on Eventually, oscillations die down, as the flowing energy is
physical systems within mechanical engineering, specifical- dissipated as heat by friction, viscous shear, electrical resis-
ly, the machines and processes which mechanical engineers tance, or magnetic hysteresis. System dynamics does not in-
design and maintain. The changes we are interested in are clude the study of heat engines. Heat engines merit the entire
changes to a system’s internal physical “power” variables, discipline of thermodynamics. Consequently, energy dissi-
such as force, velocity, pressure, fluid flow rate, current, pated as heat is not converted back into other forms of energy.
voltage, temperature, and heat flow rate. Our objective will It must remain as heat. However, for a system to function
be to predict the dynamic response of systems comprising indefinitely, the energy dissipated as heat must be removed
mechanical, fluid power, and electrical elements. from the system, or the system’s temperature will rise to dam-
Energy is a quantity which flows. Dynamical systems aging levels. Therefore, although system dynamics does not
change due to the flow of energy across the boundary sepa- include thermo, it includes the basics of heat transfer.
rating the system from the surrounding environment, and the Dynamical systems change over time. We will work with
flow of energy between elements within the system. Energy time scales which range from microseconds to hours. Re-
flows into a system from a power “source,” such as a battery, gardless of the time scale, we will define time t = 0 to be the
a pump, a moving belt, or a spinning shaft. The amount of instant an input or action is applied to a system. We will then
energy which flows from the source is determined by the divide time into three periods relative to time t = 0: before,
response of the system to the supply of energy. shortly after, and long after. The latter two are more formally
Oscillations, such as pressure fluctuations and mechani- known as the transient time period, and the steady-state time
cal vibrations, are created by energy flowing between energy period. The changes of a system’s power variables during the
storage elements within a system. When mechanical systems transient period are due to (1) the energy in the system before

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_1, © Springer Science+Business Media New York 2014 1
2 1  Introduction to System Dynamics

Fig. 1.1   An energetic system


consists of interacting energy System boundary
storage, dissipation, and conver-
sion elements separated from the
surrounding environment by the  in Energy
storage
system boundary Energy
element storage
Energy flow into the element
system from the source

Energy
dissipation
element

 out Energy dissipated as heat


is lost from the system

the input is applied, (2) the type, or “functional form,” of the 1.1.1 Why Study System Dynamics?
input acting on the system, and to (3) the fact that the input is
applied by a power supply, giving the system access to an en- Why study system dynamics? The short answer is that me-
ergy source. The system’s change due to the functional form chanical design requires the ability to predict the transient
or shape of the input, such as step change F (t ) = F0 u s (t ) or a and steady-state response of energetic systems to arbitrary
sinusoid, F (t ) = F0 sin(t ), is termed as the “forced” response. inputs. A slightly longer answer is that most of the analyses
The forced response corresponds to the particular solution you have performed so far in your study of mechanical en-
of a differential equation, as you may recall from your study gineering have tacitly assumed that the system is in steady-
of differential equations. In the ideal models we will cre- state. In other words, you have ignored the transient response
ate, the forced response will exist from the instant after the of systems you have analyzed, and, accordingly, your analy-
input is applied, at a time t = 0, and forever after (i.e., t = ∞). ses are incomplete.
The response of a system to change in its energy source is The best answer to the question “Why study system dy-
termed as the “natural” or “unforced” response. The natural namics?” is to prepare you for engineering practice. In many
response occurs whenever the system is “disturbed” from ways, engineering practice is the inverse of the undergradu-
its current energetic equilibrium by the application of a new ate study of engineering. As a practicing engineer, you will
input. A system’s natural or unforced response is always the understand the applicable engineering analyses. Your effort
same shape though it differs in amplitude. It corresponds to will be focused on determining what is assumed to be known
the homogeneous solution of a differential equation. The nat- already in textbook problems. Specifically, you will first
ural response dies out with time, as the energy flows due to need to create an engineering model, which is a simplified
the disturbance dissipates, and the system approaches a new representation of the component or system to be analyzed.
energetic equilibrium. When the transient period ends, the Next, you must establish the material properties, the operat-
amount of energy flowing into the system equals the amount ing conditions, and the forces, torques, pressures, velocities,
flowing out. The system is then said to be in “steady-state.” flow rates, voltages, or currents acting on the component or
Changes of the system’s power variables during the tran- system. Only now you are ready to begin your analysis. The
sient period consist of the sum or superposition of the per- study of system dynamics develops one’s ability to create
manent forced response and the temporary natural response. engineering models. An understanding of system dynamics
The shape or functional form of the natural response depends is needed to establish unknown material properties from dy-
only on the dynamic system and is independent of the shape namic tests. Finally, the variables which act on individual
of the input function. We will be particularly interested in the components within a system are the power variables of a
transient response of the system, which corresponds to the dynamical model of the system. System dynamics analyses
startup or shutdown of real machines. It is during the tran- yield predictions of how the power variables in a system
sient period that the amount of energy stored in the system vary with time, allowing the engineer to determine the criti-
changes and when the power variables, such as force, pres- cal loading for their analysis.
sure, and electrical current, reach maximum or minimum Many of the topics and analyses of system dynamics
values. Consequently, the design values used to size the ma- appear in other subjects and applications. The allied topic
chine components often occur during the transient period. of mechanical vibrations preceded system dynamics in
1.1 Introduction 3

historical development, and utilizes many powerful ana- the benefits of this input–output or input-response perspec-
lytical techniques, based on Newtonian analyses, energy tive is that it allows practicing engineers to make rational
methods, and virtual work. Electric circuits are dynamical decisions based on their understanding of how a system can
systems. The method we will develop in this course, the lin- and cannot behave when they lack enough information to
ear graph, is a network representation of energetic systems formulate the system equation, which is often the case.
which borrows heavily from electric circuit analysis. The The Laplace transformation, a topic in differential equa-
most direct application of system dynamics is to develop the tions, will be reviewed and then used in a slightly different
mathematical models of systems needed for automatic con- way than students may have done in the past. We will in-
trol theory. Automatic control is added to a system to change troduce vector-matrix algebra to more easily represent the
its behavior. There are two fundamental types of automatic complicated dynamic systems mathematically. This math-
control, Boolean or on-off control and closed-loop feedback ematics may be new to many students and will be presented
control. Boolean control sequences the operations of a ma- as such. Likewise, we need some mathematics from complex
chine. Boolean control is formulated in Boolean logic and analysis. This also will be new material to many students. A
can make machines appear intelligent. Closed-loop feedback limited understanding of complex analysis yields significant
control changes the dynamic response of a machine. The de- benefits in computational tools, including complex imped-
sign of closed-loop feedback control requires a mathematical ances and phasors.
model of the dynamic system to be controlled. The mathe- Numerical computation using Mathcad, MATLAB, or a
matical form of the model varies with the mathematical form package of your preference is an integral part of systems dy-
of the controller. Analog or continuous control design re- namics. Even for the limited class of differential equations
quires a differential system equation. Digital control requires we can solve analytically, evaluating the solution manu-
a difference equation. Modern or state-space based control ally (i.e., with a calculator) to present the results as a plot
requires a set of first-order differential equations presented can be tedious and error prone. Automated computation is
in vector-matrix form. As daunting as these mathematical much easier and readily provides graphical representation
terms may be on first reading, even the engineering student of the input–response relationship. Further, we will exploit
most doubtful of their mathematical ability can master the the property of “superposition” of linear systems so that we
concepts and techniques needed to describe a dynamical sys- can superimpose, or sum, many individual, simple input and
tem as a mathematical model. response plots to predict the response of systems to compli-
A wide variety of engineering mathematics is needed to cated inputs. We will find that once the syntax and methods
analyze dynamic systems. Although everyone enrolled in of the computation software are familiar, developing MAT-
systems dynamics has passed differential equations, many LAB programs to approximate and solve differential equa-
students have forgotten most of what they knew. Others may tions using finite differences is quite straightforward and
believe that they never understood differential equations will free us from the need to simplify important aspects of
even though they passed the course. To those who question a physical system to yield a system equation which can be
their capacity to understand engineering mathematics, take solved analytically.
heart in the often cited quote, “You can only understand what Although students are most apprehensive about the math-
you already know.” Accordingly, it is common for mechani- ematics of dynamics systems, it is a student’s ability with
cal engineering students to pass the prerequisite course, dif- engineering modeling and analyses which is the key to suc-
ferential equations, by learning the various solution methods cess in system dynamics. It is important for mechanical en-
and then later gain an understanding of differential equations gineering students to recognize that their curriculum pro-
with their subsequent use in system dynamics. A student gresses from the compartmentalized, sophomoric courses,
should not be embarrassed if this is their situation. He or which introduce basic concepts, to the upper-level courses in
she should accept it and make the effort needed to gain an which engineering concepts, principles, and knowledge from
understanding. previous and concurrent courses are assumed and used. This
We will review the portion of differential equations which necessary progression from narrow to broad course content
we will use, presenting the solution of differential equations can be difficult for some students, particularly those who
from an engineering perspective. Further, we will restrict adopt an algorithmic perspective in which they see, or more
ourselves to a limited number of techniques. Specifically, accurately, prefer to see, each problem as having one, and
we will emphasize that physical systems are capable of only only one, unique solution. Such a limiting perspective is a
a limited number of responses to common inputs. Conse- crippling error in judgment. We will emphasize the reality of
quently, “solving” the differential system equation will be engineering practice by highlighting the similarities between
performed by, first, identifying which type of response the apparently different engineering analyses, and working some
input to the physical system will produce and then “sizing” analyses using two or more of complementary techniques
the response, by scaling its amplitude and duration. Among which we must choose between.
4 1  Introduction to System Dynamics

The further engineering students progress through their laws to mechanical, fluid mechanical, and “hybrid” systems,
course work, the more choice they have when presented such as electromechanical systems. The linear graph method
with “open-ended” problems in which there is an objective, draws analogies between the modes of energy storage and
such as “design an X to satisfy criteria Y,” but no immedi- energy dissipation as heat in different energy “domains” (or
ately identifiable analyses which the student is expected to types), based on the manner in which the power variables
apply. It is natural, at first, for students to seek guidance of of an energy domain are measured. The name, linear graph,
the form, “What do you want me to do?” which instructors refers to the dispensing of distinct symbols to represent dif-
typically, and annoyingly, answer by reiterating, “I want you ferent types of energetic elements, as used in electric sche-
to design an X to satisfy criteria Y.” A more helpful response matics, and representing all elements as a line connecting
may be “I want you to identify the questions (design vari- two nodes with an arrow head denoting the assumed positive
ables) involved in designing an X to satisfy criteria Y, and direction for the variable, which flows or acts through the
then answer them (i.e., determine the design values).” element. The parameter abbreviation, i.e., R for resistance,
The instructor’s response is evasive, because open-ended and M for mass, is used to identify the type of element.
design problems are fundamentally different from problems We will also apply Newton’s method. You are already fa-
which illustrate and reinforce specific algorithms. In prac- miliar with Newton’s method, but are likely unfamiliar with
tice, engineering analyses are performed to either make de- its application to the vibration (or oscillation) of a system
sign decisions or to validate design decisions. The analyses of masses, springs, and “dampers.” Dampers are mechanical
used at the beginning of a design should be simple and quick, elements that dissipate kinetic energy into heat, by shearing
even at the expense of accuracy because the design is yet to a viscous fluid or deforming a viscoelastic material, typically
be defined. The final analyses will be the most precise since an “elastomer,” which is a synthetic rubber. You will find
they will be based on the final geometry of the design and the that Newton’s method will have many mathematical similar-
final predicted design loads, material properties, and operat- ities with the linear graph method, even though the approach
ing conditions. is fundamentally different.
Typically, engineers can choose among a number of meth- Lagrange’s and Hamilton’s energy methods provide a dif-
ods which will yield the result they need to make a decision. ferent perspective on dynamic system from the linear graph
Often, the various analyses were developed from different method we will develop. Lagrange’s and Hamilton’s energy
starting points, at different times, and for different applica- methods provide important insights and useful techniques
tions, and only later found to overlap. This is the case in sys- for the analysis of energetic systems. The student interested
tems dynamics. The choice of analyses depends on the type in dynamic systems is advised to study them in the future.
and nature of the system, the result needed, and the personal However, all texts must have a limit, so this text does not
preference of the engineer. When an engineer is dealing with include them.
a novel and important design, then the overlapping but dif-
ferent analysis is a relief because it allows the engineer to
check the result. If there were two analyses available, the 1.1.2 Engineering Models
engineer would use the easier of the first two because, de-
pending on the result, it may need to be repeated with revised A model is a simplified representation of reality. An engi-
design parameters or, perhaps, abandoned completely. Fol- neering model represents only one (or some) of the attributes
lowing the iterative design, the second, more difficult analy- of the real object, system, or process. Engineering models
sis would be performed as a check on the first. can be physical, analogous, or mathematical. An example of
System dynamics is a relatively late development within a physical model is a model airplane, which may only be a
mechanical engineering, dating to the 1940s. System dy- polymer shell. It is a “scale model,” or a replica of the geom-
namics was preceded by three very different theoretical ap- etry, of an actual airplane. This type of model would be used
proaches, which all yield a system equation or set of system in a wind tunnel to model the aerodynamic forces acting on
equations that relate the initial condition(s) and the input a full-sized airplane of the same shape. Alternatively, the
variable(s) to the output variable(s) using different but com- physical model of an airplane may be a scale model, which
plementary analyses and techniques. They are, in order of has motors, moveable control surfaces, radio control, and the
historical development, the following: capacity to fly. Clearly, these two different types of models
Newton’s Laws of Mechanics (1687); of the same airplane have different levels of realism and dif-
Lagrange Energy Methods (1788) and Hamilton’s Energy ferent uses. The same is true for engineering models. The
Methods (1834); and, lastly, level of realism varies with the intended use, which depends
Kirchhoff’s Circuit Laws (1845). on the stage of the design process the model is used in. In all
Our focus will be on a network method, the “linear graph” cases, an engineering model is a simplified representation of
method, which is an extension of Kirchhoff’s electric circuit the real entity.
1.1 Introduction 5

Fig. 1.2   Creation of an engineer-


ing model of an actual physical Actual Physical Modeling Process Engineering
object or system is a filtering Object or System Abstraction and Model
process, in which only simplified
approximations and the relevant Approximation of
properties are retained Relevant Properties

Engineering models are “abstractions” of real, physi- Although the term “engineering model” may be new, you
cal objects or systems. “Abstract” is a word, which intimi- began using engineering models in your first engineering
dates students because one of its meanings is “difficult to course. What is new is that we are now considering the pro-
understand,” as in “abstract mathematics.” However, the cess of developing the model, Fig. 1.2. Previously, you were
primary meaning of “abstract” comes from its Latin roots, presented a model as the starting point for learning a new
which means “to pull or drag away from.” Note that “ab- concept or computation. For example, in a statics problem,
stract” shares a root with “tractor” and “traction.” Pulling the geometry of an object is retained, but the material proper-
or extracting representative aspects from a whole to create a ties of an object are simplified (or abstracted) by “assuming”
simplified version is the meaning of “abstract” we will use. it is perfectly rigid. The “assumption of perfect rigidity” is
This meaning will be familiar to some, since it is the same better termed the “simplification of perfect rigidity” since
usage of abstract as the “abstract” at the beginning of this it is a choice we make to simplify the engineering model.
chapter, which is a short summary of it. Consequently, our How reasonable it may be to make the choice to simplify the
use of “abstract” will be close to “simplified.” For example, material properties by modeling the object as rigid is rela-
if our objective is to predict the translational acceleration of tive, and is determined by comparing the results computed
an object subjected to a force then only relevant property of using the simplified model, with results computed using a
the object is its mass. If our objective is to predict the trans- more realistic model, or, if available, with experimental data.
lational and rotational acceleration of the object, the relevant Needless to say, an engineering student first learning statics
properties are the geometry of the object and its density or, cannot develop a more realistic model; nor does he or she
if it varies, the distribution of its densities. It is important to have a basis of experience to make the judgment of how rea-
remember that every real object has mechanical, electrical, sonable an engineering model is. Consequently, engineering
and chemical properties. We often consider only one type or modeling is not taught in introductory courses; the models
class of properties at a time and forget that the real object has are simply presented.
all physical properties, some of which may affect the perfor- Engineering computations are performed to provide
mance or reliability of our design. a basis for design decisions, and that design is a series of
All engineering analyses require an abstract model. Let choices. Early in the design process, few choices have been
us “deconstruct” this last sentence. Although we commonly made, and a great deal of uncertainty exists about the design,
employ the word, “analyze,” to mean “work to understand,” the materials and manufacturing processes that will be used
the literal meaning of “analyze” is “to take apart or break to make it, and the operating environment it will be used
into pieces,” which is indeed the first step in understanding in. The initial engineering models must be correspondingly
anything complicated. “Abstract” can mean “hard to under- simple. Often, a simple model is sufficient to reject or accept
stand,” but it can also mean “simplified,” as in an abstract a design choice. In other cases, the results fall into a “gray”
painting. “Model” means “a representation of something,” area and the initial, simple calculations are inconclusive, or
which includes only some of its attributes. Engineering the cost or importance of the part merits refining the initial
analyses are based on abstract models, since each physical results, by repeating the analyses on a more realistic and,
phenomenon which we must engineer for a part, machine, consequently, complicated model. For example, we may
or system typically requires a separate calculation. Although first choose to model an object as perfectly rigid, in order to
the phenomena coexist in reality, we separate them, in order consider only the static equilibrium of the object. Later, to
to simplify the problem sufficiently, so that we can handle it increase the realism of our model, we may choose to model
with the mathematical tools we have available. The model the material as “elastic–perfectly plastic.”
used for a calculation depends on what we wish to know, The familiar “elastic–perfectly plastic” material is a
and contains only the properties and attributes of the cor- model, Fig. 1.3. There is no material which actually behaves
responding physical object or system needed to describe that like this, but it is a reasonable approximation for some impor-
phenomenon. When it is impossible to separate two different tant materials, and, consequently, is widely used. Although
phenomena because of their physical coupling, the engineer- the elastic–perfectly plastic material model appears simple
ing analysis is more difficult. since it consists of only two straight lines, it is a non-linear
6 1  Introduction to System Dynamics

Elastic Plastic 1.1.4 Extraction of Mathematical Statements


σ = Eε σ = σy from Engineering Models
σy
The physical aspects of an engineering model, which can be
expressed as mathematical statements, clearly depend on the
σ E
particular engineering model. However, these mathematical
statements, usually equations but sometimes inequalities,
fall into two broad categories:
1. Mathematical statements of fundamental physical facts
that must always be true, and
2. Simplified engineering models of materials and de-
εy ε vices, which are approximations and, therefore, never
completely true.
Fig. 1.3   Elastic–perfectly plastic material model Conservation of mass and energy, equilibrium of forces and
moments, and geometric compatibility of deformations and
displacements are physical facts which must be true under
material model and can greatly complicate engineering com- all circumstances. The accuracy of these equations is limited
putations. We will return to this important point but, briefly, only by our ability to measure the relevant quantities. Engi-
a linear material (or property or “constitutive”) model is a neering models, on the other hand, are always approximate
single straight line which passes through the origin. and vary in accuracy from application to application. For
example, the linear elastic material model, σ = E ε, is quite
accurate for steel, before the stress reaches the yield point.
1.1.3 Mathematical Models Conversely, σ = E ε is an approximation for aluminum,
which has no defined yield point.
A mathematical model is an equation or set of equations To describe anything real with great accuracy is, at best,
which form an input–output relationship, where the input difficult and costly in engineering time. It is often simply
variables are known and the output variables are what we impossible, because the information is not available, or the
want to know, Fig. 1.4. A mathematic model is based on an variability between specimens or samples is so great, that
engineering model. The various attributes and properties of the uncertainty of the model makes its increased precision
the engineering model can be described by simple math- meaningless. For example, most stress calculations are
ematical statements. We work to create a compendium, or performed assuming the material is ideally elastic, and the
list, of all equations which describe physical aspects of engi- stresses and strains are kept below yield to prevent perma-
neering model. These equations are then “reduced” to elimi- nent deformation. The material model for an elastic stress
nate the unwanted variables, leaving just the input which we calculation requires Young’s modulus E, a single datum. If
know, the output we wish to calculate, and the independent the stress will produce permanent plastic deformation, then
variables of time and position. For example, if we were siz- a non-linear relationship is needed. Typically, the ideally
ing a part such as a shaft with a circular cross-section, the elastic–perfectly plastic model is used. It requires two val-
properties would be the geometric constraints, such as the ues, Young’s modulus E, and either the yield stress, σy, or
required length and the maximum possible diameter, the ap- the yield strain, εy. The yield stress or strain of a material is
proximate material properties of the shaft material, and the not known with the certainty of Young’s modulus. Although
approximate relationship between stress and strain (i.e., the Young’s modulus for steel is independent of the specific
material model), the input would be the forces and torques alloy and heat treatment, the yield stress and strain are not.
acting on the part, and the response would be stress, strain, If the manufacturer or supplier does not provide test data for
or deformation of the part. these values, the designer must rely on nominal values found
in standards and handbooks. Nominal values of yield stress

Fig. 1.4   Mathematical engineer- Response or


ing models are input–output
relationships used to facilitate Input Variable Mathematical Output Variable
design decisions Engineering
Forces and Torques Model Stresses, Strains,
or Deformation
1.1 Introduction 7

and yield strain for a specific steel are not average values. The mass distributed throughout a volume is concentrated, or
Although engineers often use “nominal” to mean “normal,” lumped, as mass M at the mass center. Similarly, the elastic
“nominal” means “in name” only. In the case of yield stress, strain of a spring is lumped into a spring rate or spring con-
a nominal value is often below the expected minimum yield stant K.
stress. This is a useful value if yield must be prevented, but The fact of the matter is that we must use very simple
it is not the yield stress. equations to describe the physical aspects of the engineer-
A more accurate description of a stress–strain relation- ing model, or the resulting mathematical model is useless,
ship requires replacing the two straight lines of the elastic– because we cannot solve it. By and large, most equations
perfectly plastic model with a single, continuous curve. This which describe the physical aspects of an engineering model
type of non-linear model requires significantly more infor- are purely arithmetic—either sums, such as, an equilibrium
mation than a “piece-wise” continuous model composed of statement that the forces in the x-direction sum to zero, or
two straight lines. Non-linear stress–strain models are only linear algebraic equations, such as σ = E ε. The most intimi-
used in mechanical design where the economic value of the dating type of equation we will need to describe the phys-
improved design justifies the expense of the analysis, such ics of our models are linear, first-order ordinary differential
as the design of automobile bodies. Ultimately, the accuracy equations.
of a material model depends on the variability of the ma-
terial being modeled. Engineers deal with “populations” of
things, where the number in the population ranges from a 1.1.5 Formulation of a Mathematical Model
few to millions. There is always variation between individ- from an Engineering Model
ual specimens and samples drawn from the population. The
properties of populations are described stochastically using A mathematical model is derived from an engineering model,
probability to quantify the uncertainty. In order for a sophis- by reducing the set of equations representing the physical
ticated non-linear material model to yield useful results, the phenomena and properties of the engineering model to an
properties of the sample used to create the model must be equation which yields the value of an output variable (what
representative of the population of production parts. For ex- we wish to know) as a function of the input variable (what
ample, plastic forming of steel changes it properties. How do we do to the system) and the independent variable (usually
the properties of the steel in a location of an automotive body either time a position in space).
correspond to the test specimens? In general, even the most Reduction of the set of equations extracted from the engi-
sophisticated engineering calculations are best described as neering model to a mathematical model is conceptually the
approximations. easiest step of the process. We will reduce the set of equa-
The process of formulating a mathematical model from tions to an input–output relationship using “elimination by
an engineering model is aptly described as analysis, which, substitution.” We will begin, by choosing an equation from
again, strictly defined means “to take apart.” How one takes the set of equations that contains the input variable, the out-
an engineering model apart depends on whether the variables put variable, or both. We will then substitute for a variable
of interest vary continuously with position and time, or only in the equation, which is neither input nor output nor time.
with time. When materials are considered on the macroscopic Repeating this process eventually yields an equation whose
scale, i.e., we use a “continuum” model to represent materi- variables are only input, output, and time. The challeng-
als with distributed properties. Systems in which the dynamic ing aspect of the reduction is the number of substitutions
response of variables change continuously with both position required to eliminate all of the unwanted variables. Simple
and time, as is the case for pressure and velocity in fluids and algebra can become error prone when there are a number of
stress and strain in solids, are described by partial differen- steps unless done with care.
tial equations. Partial differential equations are significantly
more difficult to deal with than “ordinary” differential equa-
tions, which contain derivatives of only one variable (i.e., ei- 1.1.6 Summary of Engineering Modeling
ther time or position). In many circumstances, we can greatly and Analysis
simplify a model, and avoid the necessity of partial differ-
ential equations with negligible loss in accuracy, by concen- While each of your engineering courses presents new mate-
trating the distributed properties of a continuum into discrete rial, engineering analysis continues to build upon the same
elements. The modeling process of representing distributed framework, regardless of application. The commonality be-
material properties as discrete elements is known by the odd tween your various engineering courses is quite deep and
term, “lumping.” A distributed property becomes a “lumped permits understanding of the new through analogy to the old.
parameter.” A familiar example of a lumped parameter is the The method is summarized as follows:
concentration of the mass of an object, at its center of mass.
8 1  Introduction to System Dynamics

Fig. 1.5  a Model of a force


a b
acting on a mass supported on a W = Mg
frictionless surface. b Decom- +x,+v
position of the applied force into g FN
horizontal and normal compo-
nents
F
Fx
M M
Frictionless Surface

FR FR

1. Draw a picture(s) of what you have chosen to analyze. placement vector x. It is only the component of force, F, in
This is your model. Annotate the picture(s) to define all of the direction of the displacement, x, which does work. In
the variables and parameters in your model. order for force F to “do work on the mass” the point of appli-
2. Write mathematical statements of relevant physical truths. cation of the force, F, must displace Fig. 1.5. If the point of
3. Reduce the mathematical statements, eliminating all un- ­application of the force does not move, that is, if x = 0, then
known variables, except the input and output variables, the force does no work, and there is no energy transfer into
and solve the equation (or system of equations) for the the mass.
output variable(s). Energy is a quantity which can be accumulated in a physi-
4. Evaluate the solution. Iterate if necessary. cal object, transferred from one object to another, and con-
We shall explore the analogies between the application of verted from one form of energy to another, until the energy
engineering analysis to mechanical design and system dy- is ultimately dissipated as heat energy. Energy can flow. Fur-
namics in this chapter. ther, given the opportunity, energy will flow. Power is the
flow rate of energy, expressed as

1.2 Mechanical and Energetic Models dE


P= (1.2)
dt

We will begin system dynamics, by reviewing the funda-
mental definitions of energy and power, as well as introduc- Figure 1.5a shows a force, F, acting on a mass supported by a
ing concepts needed for “energetic modeling” of dynamic frictionless surface. This drawing is an “energetic model” of
systems. The flow of energy into, out of, and within physical a real, physical system. Engineering models are greatly sim-
systems is what makes dynamic systems dynamic. Unfortu- plified representations of real objects or systems. An energet-
nately, most of the physical variables mechanical engineers ic model is a simplified representation of just the significant
work with cannot be directly observed. We must infer them, energetic properties or attributes of a system. The frictionless
by their effects which we can observe. For example, we can- surface in the model is an idealization, since there must be
not see force, but we can see the deformation of an object some friction between two surfaces. By modeling the surface
subjected to a force, or the motion of a mass accelerated by a as frictionless, we represent the friction force and the energy
force. Likewise, although we literally see light energy in the dissipating as heat, due to the friction between the mass and
form of photons striking our retina and feel heat, we cannot the surface being negligible. Significant and negligible are
sense any other forms of energy or power. We must infer the comparative adjectives that are always relative. In a New-
amount of energy stored in an object by its deformation or tonian model, whether a force is significant or negligible is
its motion. judged relative to the other forces present. In an energetic
A fundamental definition of energy is mechanical work model, whether the amount of energy stored or dissipated is
significant or negligible is judged relative to the other quan-
E ≡W = F⋅x (1.1) tities of energy in the system.
 In Fig. 1.5, the reaction force FR is approximately the same
where F is force and x is displacement. The dot or scalar magnitude as the applied force F. In a Newtonian model, the
product is the product of the projection: the force vector in reaction force FR would not be judged to be of negligible
the direction of the displacement vector, Fx, times the dis- magnitude. However, the reaction force FR is normal to the
1.2  Mechanical and Energetic Models 9

Round steel bar Round steel bar


Diameter D Diameter D
Young’s modulus E
Young’s modulus E Yield strength σy
Yield strength σ y Non-Uniform Non-Uniform
Stress Stress
F(t) F(t) F(t) F(t)
σ = Eε
Uniform Stress

L ~D ~D

L
Fig. 1.6   Axially loaded round steel bar
Fig. 1.7   Saint Venant’s ­principle for uniformity of stress. In an axially
loaded cylinder, non-uniform stress extends a minimum ­distance equal
displacement of its point of application. The reaction force to the diameter of the cylinder from the applied load
FR friction does no work against the mass and, therefore, is
negligible in an energetic model.
It is easy to recognize a frictionless surface as an ideal- fine the model used to represent the material of the bar as
ization. What is not apparent is that identifying an element elastic–perfectly plastic, Fig. 1.3. The elastic–perfectly plas-
of the energetic model as a “mass” is also an idealization. tic material model does not describe a real material. We shall
Note that we described the element as being a mass rather see that even this simplified model presents mathematical
than having mass. This semantic subtlety restricted the ele- difficulties because it is non-linear. It consists of two straight
ment to the single physical property of mass or inertia, at lines, not one straight line. A linear relationship is a straight
the exclusion of all other properties, such as stiffness, or its line, which passes through the origin. A linear material
inverse, compliance. Consequently, a mass is a mechanical model allows scaling (doubling the force doubles the stress)
element that is perfectly rigid. If it were not, then the ele- and superposition (sum of the stresses due to different forces
ment would have also have compliance and either store or equals the stress due to the sum of the forces). The elastic–
dissipate energy due to deformation. Describing the mass as plastic model is only linear in the elastic region.
a “mechanical” element further restricts it, by eliminating A second aspect of the model shown in Fig. 1.6 is so
all attributes of mass other than inertia, such as thermal ca- familiar, that it may escape notice. It is convention to rep-
pacitance. It is difficult at first to accept that the elements resent the force acting on a physical object as a vector. Yes,
of an energetic model describe real properties, but the ele- force is indeed a vector quantity. The vector representation
ments themselves are idealizations. No real object has only is physically accurate, but it is also abstract or simplified.
a single energetic property. However, many objects have a The modeling simplification is the omission of the details
dominant energetic property. For example, the springs of an as to how the force is applied to the object. The drawing
automobile’s suspension have mass, electrical conductivity, implies that the force is concentrated at a point which
and thermal capacitance, but their dominant energetic prop- certainly cannot be the case, since the stress at that point
erty is strain energy storage. Conversely, it is not unusual would be infinite. Why then represent the force applied as a
for a single real object to have multiple energetic properties vector? Models are simplifications, which contain only the
of approximately equal importance. When that is true, then essence of the physical system needed to perform a calcu-
the single object is modeled by multiple energetic elements. lation. A force applied to a machine element is represented
An ideal mass accelerated by a force is a dynamics problem. as a vector, when only the magnitude and direction of the
Energy is stored in the mass as kinetic energy, the energy of force are important. If an applied force is depicted as a vec-
motion. When material properties are included in the model, tor, then how that force is generated and transmitted to the
the motion of the point of application of a force is, in general, bar is unimportant to that model or unknown.
the sum of displacement and deformation. The deformation To calculate the stress in the bar, we must assume that
of a material can both store and dissipate energy, requiring stress is uniform on a transverse cross-section. The assump-
additional energetic elements in the model. tion of uniform stress sets a minimum distance from the con-
A model is a simplified approximation of the key aspects centration of stress of unknown magnitude where the force is
or attributes of actual materials, objects, processes, and phe- applied. Saint Venant’s principle is that the stresses are rela-
nomena. Consider the familiar problem of calculating the tively uniform at a distance of approximately one diameter
stress in a round rod due to an axial force F( t) using the from the applied force in an elastic object, Fig. 1.7.
model shown in Fig. 1.6. What assumptions and simplifica- If the stresses are uniform anywhere over the length of
tions are stated or implied in Fig. 1.6? In other words, what the bar, they will be uniform in the middle. A revised model
physical models and constraints are used on the variables? for the calculation of the stress in the bar which explicitly
The “callout” provides two material parameters, Young’s indicates the location of the stress calculation is shown in
modulus E, and the yield strength σy. These parameters de- Fig. 1.8.
10 1  Introduction to System Dynamics

Round steel bar


Diameter D x1, v1 x 2 , v2
Non-Uniform
Young’s modulus E
Yield strength σy F(t) K F(t)
Stress
F(t) σ = Eε Aσ(t) Translational Spring
Uniform Stress
Fig. 1.10   Schematic symbol for a translational spring K showing the
~D locations of distinct values of translational displacement and velocity at
L
_ the ends of the element
2
Fig. 1.8   Revised model showing the cross-section on which the stress
is assumed to be uniform wherein,

ΔL
 ε axial ≡ (1.4)
Elastic Deformation
L
∆L
L We can “lump” the distributed property of elasticity into a
F(t) F(t) single parameter by modeling the bar as a spring with the
elastic property, the spring constant, K. The deformation ΔL
Round steel bar of the bar is the difference in the displacements x1 and x2 of
x1 , v1 Diameter D
Young’s modulus E x 2 , v2 its two ends. The conventional notation for displacement is
Yield strength σy
the following: x1 − x2 = Δ x. We will use an alternative nota-
Fig. 1.9   Axially loaded elastic bar of length L with elastic deforma-
tion where the subject locations (or nodes) are identified by
tion ΔL. Locations of distinct values of translational displacement and subscripts. We will use the numbers, which identify the ends
velocity are denoted as x1,v1 and x2,v2 of the element. Hence,

x1 − x2 = ∆ x ≡ x12 (1.5)

The last aspect of the model we will consider is the mag-
nitude and mathematical functional form of the applied We will use the same notation to denote the difference be-
force, F( t ). We have noted that the vector representation of tween velocities, pressures, voltages, and other analogous
the force applied to the bar omits the source of the force, variables, which vary across elements.
as well as the manner in which the force is transferred to The elastic energy stored in the bar is recoverable. The
the bar, indicating that they are unimportant for the result of energy transferred to the bar to deform it plastically is dis-
the calculation. Are there physical limitations in the model sipated as heat, as the metal shears and is not recoverable.
which constrain the magnitude of the force applied or how it Although there are exceptions, such as head bolts in automo-
can vary with time? The magnitude is limited. The greatest bile engines and some other fasteners, the machine compo-
force which can be applied to the bar is that which causes nents are sized to keep stresses below the yield stress. When
yield. stress is below yield, all of the mechanical work done on
the component is stored as elastic strain energy and can be
Fmax (t ) = σ y A recovered. The schematic symbol for a significant amount
of strain energy storage in an energetic symbol is a spring,
Hence, the operating range of the applied force is limited to Fig. 1.10.
the following: The familiar equation for a spring is

F (t ) ≤ σ y A FK = K ∆ x (1.6)

The bar shown in Fig. 1.6 behaves as a spring, when its stress where K is called the spring rate in machine design and the
is in the elastic region below yield, and the axial stress is spring constant in system dynamics. Although it will seem
proportional to the axial strain awkward at first, we will use the notation of Eq. 1.4 to ex-
press deformation of the spring element as the difference in
 σ axial = E ε axial (1.3) displacements of the two ends of the spring, denoted as loca-
tions or “nodes” one and two.

FK = K x12 (1.7)

1.3  Energy, Mechanical Power, and Coenergy 11

1.3 Energy, Mechanical Power, and Coenergy


FK
1.3.1 Strain Energy K
1
Is there a dynamic constraint on the applied force, F( t), ap-
_
plied to the spring in Fig. 1.10? Specifically, is it physically  K= 12 Kx12
2

possible to instantaneously apply an axial force, F( t), to an


elastic member? No, it is not. There are two complemen- x12
tary perspectives, both of which lead to the same conclusion.
The Newtonian perspective involves forces, displacements, Fig. 1.11   Elastic strain energy stored in a linearly elastic spring is the
and their derivatives. If we were to posit a small but finite area between the FK(x) line and the displacement axis
force on a spring instantaneously, then we are saying that
the deformation of the spring due to that force can occur in-  1
E Spring = K x 2 max (1.11)
stantaneously. A finite deformation of the spring in zero time 2
requires infinite velocity of the ends of the spring, which is
physically impossible. A force–displacement plot of a linearly elastic spring is
The complementary perspective is to view the spring as shown in Fig. 1.11 where the force, FK, is the force acting
an energy storage element, and look for similar physical through the spring, and x12 is the deformation of the spring,
impossibilities. Recall that the differentials, d E and dt , are which is the difference in displacement of the two ends of
infinitesimal. They are infinitely small, compared to the “fi- the spring. The elastic strain energy is the area of this plot
nite” quantities, ∆E and ∆t. If there were a finite change of below the curve.
energy ∆E during an infinitesimal time, dt , then the power Eliminating the deformation of the spring x12 yields a
P would be infinite. less familiar, or, perhaps unfamiliar, expression for elastic
energy.
∆E ∆E
≈ ≈∞ → P ≈∞ (1.8) xmax FK max
F 
 dt 0 E Spring = ∫
0
FK dx12 → E Spring = ∫
0
FK d  K 
 K
Instantaneous deformation would require infinite power and
is, therefore, physically impossible. Note that the upper limit of integration was changed from the
Work is defined as the dot product of a force and the maximum displacement, xmax, to the maximum spring force,
point of application of the force (Eq. 1.1). Consequently, Fmax. Evaluate the integral
there must be displacement of the point of application of a
force to transfer energy into or out of an element or system. 1
FK max
1 2
FK max

The energy transferred to a spring equals the work done on E Spring =


K ∫ FK dFK → E Spring =
2K
FK
0
the spring. 0

1 2
E Spring = FK max (1.12)
xmax
 2K

E Spring = WSpring = ∫
0
FK ⋅ dx12 (1.9)

Note that the energy stored in a spring is positive, whether


We can dispense with the vector notation and the scalar prod- the spring is compressed or extended. The strain energy is
uct, when force and displacement are co-linear. positive, because the spring force and the deformation have
the same sign in Eqs. 1.6 and 1.7.
xmax
Newtonian analyses use the term, potential energy, for the

E Spring = ∫
0
FK dx12 (1.10)
energy stored in a spring, because that energy has the po-
tential of creating motion and being converted into kinetic
We can then eliminate either force or deformation from the energy. The phrase, strain energy, will be used in this text to
work integral, using the relationship between spring force differentiate it from other forms of stored energy, which can
and deformation. Eliminating the spring force, FK, yields the be converted into kinetic energy.
familiar expression for the energy stored in a spring

xmax xmax 1.3.2 Mechanical Power


E Spring = ∫ FK dx12 → E Spring = ∫ K x12 dx12
0 0 Real physical objects have multiple energetic properties.
Ideal energetic elements, such as an ideal mass, are abstract
12 1  Introduction to System Dynamics

Massless, rigid bar x,v quently, a transfer of energy. When we examine the second
K term of Eq. 1.14,
b1
F(t) dF

b2 dt
we find that the displacement is not changing, only the force
is. If the point of the application of force is not moving, there
is no transfer of energy. The second term does not contribute
Fig. 1.12   A bar modeled as massless and rigid acts as a “force spread- to mechanical power.
er” in a mechanical schematic. The cross-sections of a piston in a cylin-
der are symbols for dampers which dissipate mechanical power as heat
dx dF
by pumping fluid P�
= F· + x·
dt  dt
Does No
simplifications with a single energetic property. An ideal Work
mass’s single energetic property is the ability to store ki-
netic energy. In order to have this single energetic property, Mechanical power is the first term alone.
an ideal mass must be rigid. If an ideal mass were deform-
able, then it could store strain energy or dissipate energy as dx
P = F· (1.15)
heat through shear deformation, or both. It would have two  dt
or three energetic properties, not one. An object modeled as
rigid, but not given the attribute of mass, can neither store The time derivative of displacement x is velocity v:
nor dissipate energy.
We include ideal rigid bars and rods in models of mechan- dx
≡v (1.16)
ical systems to transmit force from one energetic element to  dt
another, Fig. 1.12. There must be displacement of the point
of application of a force to transfer energy into or out of an Thus, mechanical power is the dot product of force and the
element or system. If a rigid object is free to move, it can velocity of the point of application of the force.
transfer energy as well as force. Conversely, if a rigid object
cannot move, then it can provide a reaction force, but that P = F·v
(1.17)
force can do no work. It follows that energy cannot be trans-
ferred into or out of a rigid object fixed in position. Motion
is required to transfer energy. There must be displacement 1.3.3 Strain Coenergy
of the point of application of a force in the direction of the
force, in order to perform mechanical work. We concluded in Sect. 1.2.1 that it is physically impossible to
From the definitions of work, Eq. 1.1, and power, Eq. 1.2: instantaneously apply an axial force F to an elastic member,
because a finite energy transfer in infinitesimal time requires
dE dW d ( F ⋅ x ) infinite power. We also noted that the deformation x of the
P= → = (1.13)
 dt dt dt spring would occur instantaneously, which requires infinite
velocity. There is another perspective we can apply to that
The derivative of a dot product is evaluated as if the dot question. If it were possible to instantaneously apply force F0
product were conventional multiplication, yielding the sum to an elastic member, such as a spring, and then create defor-
of two terms. mation x0, the work done on the spring would be the product.

dx dF W = F0 x0
P = F· + x· (1.14)
 dt dt
This is the rectangular cross-hatched area in Fig. 1.13.
Examining the first term, We know that the strain energy that is actually stored
in the spring, when it is deformed x0 under force F0, is the
dx
F· triangular area under the curve, Fig. 1.11. The correspond-
dt
ing area above the curve is “coenergy.” Coenergy is a ficti-
we find that the displacement is changing with time. There tious mathematical entity. It does not exist, but it is useful
is motion of the point of application of the force and, conse- computationally. Coenergy is defined as an integral similar
1.3  Energy, Mechanical Power, and Coenergy 13

fixed, when the spring force changes by the infinitesimal


F0 amount, dFK. The expression,
Coenergy K
FK 1

x12 dFK ≠ dW (1.19)

Energy also illustrated in Fig. 1.14, does not represent an infinitesi-


mal amount of mechanical work, because the displacement,
x12 x0 x12, does not change. There must be motion of the point of
application of a force for that force to perform mechanical
Fig. 1.13   The work which would be done on a spring, if force F0 and work. Therefore, we can conclude that integral of the infini-
its resulting in deformation x0 could be applied instantaneously, is the tesimal product, Eq. 1.19, is not work, but something else,
cross-hatch rectanglular area, F0x0. Only the work represented by the even though the quantity has the units of force times dis-
triangular area below the FK(x) line is physically possible
placement. Notice that this logic is the same that we applied
to derive the expression of mechanical power, Eq. 1.15.
It is clear from the geometry of Fig. 1.13 that the energy
Coenergy and coenergy are equal in the case of a linearly elastic spring,
which has a spring constant K which is indeed constant. This
is not the case for any non-linear elastic spring, where the
Force, F dF x(F) spring rate K is a function of the spring displacement x, thus,
F(x) Energy
F ( x12 ) = K ( x12 ) x12 (1.20)
dx 
Non-linear springs can be designed to become either soft-
er or stiffer with displacement, Fig. 1.15b. For example, a
Displacement, x spring which becomes stiffer with displacement may have
the spring rate of Eq. 1.21, plotted in Fig. 1.14.
Fig. 1.14   Energy is the area below the force–displacement curve. The
fictitious but useful quantity, coenergy, is the area above the curve
K ( x) = K 0 + K1 x (1.21)

to Eq. 1.10, but with the derivative on the spring force, FK, Energy and coenergy are equal, when there is a linear re-
rather than on the spring displacement x. lationship between force and displacement. When the rela-
tionship between force and displacement is non-linear, then
Fmax
knowing either the energy or the coenergy allows calculation
COE Spring = ∫ x12 dFK (1.18)
of the other, by subtracting it from the area of the rectangle,
 0
FKmax xmax:
Although this integral resembles the intermediate result ob- FK max

tained deriving Eq. 1.12, which expresses the amount of en-



Energy = FKmax xmax − ∫ x( FK ) dFK (1.22)
ergy stored in a spring in terms of the force acting through 0

the spring, there is a subtle but important difference. Remov- and


ing the derivative from spring displacement x and placing xmax

it on the spring force FK indicates that the displacement is



Coenergy = FKmax xmax − ∫
0
FK ( x) dx (1.23)

Fig. 1.15a   Non-linear spring rate,


K ( x) = 5,000 N/m + 50,000 N/m 2 x. a 20 b 5
b Spring force vs. displacement
for the spring rate in Fig. 1.15a 4
15
K(x) FK(x) 3
__ 10
kN kN
2
m
5 1
0 0
0 0.1 0.2 0 0.1 0.2
x12 , m x12 , m
14 1  Introduction to System Dynamics

σ yield Elastic Plastic


σ = Eε σ = σy

σyield
Stress, σ
Energy Coenergy
density Stress, σ density
Energy
density

Strain, ε
Fig. 1.16   Energy density is the area below a stress–strain curve. It is ε yield ε max
the quantity of energy stored and/or dissipated deforming a unit volume Strain, ε
of material
Fig. 1.17   Coenergy density is the area between the stress–strain curve
and the stress axis

Equations 1.22 and 1.23 are Legendre transformations. A


Legendre transformation from thermodynamics is the re-
lationship between enthalpy, internal energy, pressure, and from the work piece. This is termed “spring back,” and ne-
volume. cessitates bending the work piece further than the desired
permanent bend. The mechanism of spring back can be seen
H = U + PV
(1.24) in Fig. 1.17. The energy density of a metal plastically de-
formed is dissipated as heat and cannot contribute to spring
back. However, the material stores the elastic strain energy,
1.3.4 Energy Density and that energy is recovered, when the bending force is re-
moved, causing spring back.
Strain energy and coenergy are expressed in terms of stress
and strain as “energy density,” which is the energy or coen-
ergy per unit volume of material. The area below a stress– 1.3.5 Kinetic Energy
strain curve is the energy stored or dissipated in deforming a
unit volume of the material. Energy density must be integrat- An old joke among engineering professors is that students
ed over the volume of a component to determine the amount part way through their engineering education believe they
of mechanical work stored as strain energy and dissipated as understand energy but do not understand entropy. Their en-
heat performing plastic deformation. gineering education is complete, when they realize that they
Calculating the energy density of the elastic–perfectly don’t understand energy either. Mass and energy are very
plastic material model shown in Fig. 1.16 is an example of mysterious at both the atomic and subatomic scales, where
the utility of coenergy. The energy density in the material is they are governed by quantum mechanics, and at extremely
calculated by subtracting the coenergy from the product of high energies on the macroscopic scale, where relativity
the yield stress, σy, and the maximum strain, εmax. governs. The transmutation of energy into mass and mass
into energy is a bizarre concept. However strange the im-
Energy Density = σ yield ε max − Coenergy Density (1.25) plications of relativity may be, they are not as confounding

as the reality of quantum mechanics. It is easier to accept a
where the coenergy density is the triangular area between the particle acquiring mass, as it is accelerated to great velocity,
stress–strain curve and the stress axis, Fig. 1.17. than it is to accept that the same particle must also be viewed
as being a wave having no definite location in space, and
1 may spontaneously appear elsewhere. Although some phe-
Coenergy Density = σ yield ε yield (1.26)
2 nomena mechanical engineers work with, such as electrical
 conduction and radiative heat transfer, require quantum me-
Energy density is useful in understanding the rupture of ma- chanics concepts to understand, most mechanical engineer-
terials and the behavior of materials during plastic forming, ing is done on a human size and energy scale or not far from
such as bending, stamping, and forging. A phenomenon ob- it. Consequently, the effects of both quantum mechanics and
served during bending metal is the recovery of some of the relativity are negligible, and we use Newtonian mechan-
bending deformation, when the bending force is removed ics without corrections. It is interesting to ponder over the
1.3  Energy, Mechanical Power, and Coenergy 15

mysteries of nature, and humbling to reflect on how little we  dvx


Fx = M (1.31)
actually understand. However, from the perspective of what dt
we can predict with confidence to make engineering deci-
sions, we can describe the relationship between mass and Similarly, we can eliminate the dot product in the expres-
energy with Eq. 1.1, the definition of energy as mechanical sion of mechanical power, Eq. 1.17, when the force and the
work, and two additional algebraic equations, Eqs. 1.27 and displacement of its point of application are restricted so that
1.28, and a differential equation, Eq. 1.30. they are colinear. We again assume that force and displace-
Both mass and energy can move or flow. Although mass ment are in the x-direction:
and energy often move together, we account for them sepa-
rately with two Conservation Laws, when we exclude atomic P = F · v → P = Fx vx (1.32)

reactions and relativistic velocities. Conservation of mass is
a statement that the difference between the mass that flows We now substitute for Fx using Newton’s Second Law:
into a system and that which flows out must be stored within
the system. dvx
P=M vx
dt

∑ mass In − ∑ mass Out = massStored (1.27)
Multiplying both sides by dt and rearranging the right side,
Conservation of energy is expressed similarly.
P dt = Mvx dvx

∑ energyIn − ∑ energyOut = energyStored (1.28)
dE
using the definition of power P = rearranged as
In both Eq. 1.27 and 1.28, we have defined flow into a sys- P dt = d E and then integrating yields dt
tem as positive and flow out of a system as negative.
The remaining equation needed to describe the relation- vx2
ship between mass and energy is Newton’s Second Law, the ∫ P dt = ∫ Mvx dvx → ∫ d E = M ∫ vx dvx → E = M 2
relationship between a force acting on a mass and its accel-
eration. We will express the familiar The result is more easily recognized as kinetic energy, when
rearranged as
F = Ma (1.29)

 1
E Kinetic = Mvx2 (1.33)
as a differential equation 2
dv Kinetic energy will be given the subscript M to identify it as
F=M
(1.30)
dt the energy of a mass. Although gravitational potential energy
is also stored in mass, we will find it convenient to express
in order to write the equation in terms of velocity rather than that form of energy as work done against the gravitational
acceleration. Force and velocity are known as the power force, for reasons which we will discuss later.
variables of a translational mechanical system, since their
1
product is mechanical power, Eq. 1.17. (1.34)
E Kinetic ≡ E M = Mv 2
Energy is a quantity that moves or, more familiarly, flows. 2
The flow rate of energy is power. Although it is common to There is no such thing as negative mass, only accounting
speak of a “power flow,” power is a flow rate. Power is the deficits of mass or mass flow which yield negative signs.
energy flow rate into or out of an object, across a boundary, We also generally think of energy as a positive quantity, but
or through a conductor or transmission element during a unit here signs become a problem. Kinetic energy must be posi-
of time. We will use both terms, power and power flow, since tive due to the square of velocity. However, the electrostatic
the redundancy serves to remind us that power is a flow rate. potential energy between two charges Q1 and Q2 separated
We will now investigate the relationship between New- by a distance d can be either positive or negative, depending
ton’s Second Law and kinetic energy. If we restrict the mo- on the signs of the charges. If both charges have the same
tion of a mass M to a horizontal plane, so that it cannot gain sign, then the force between them is repulsive, and the po-
elevation against gravity, then the only type of energy the tential energy is defined to be positive. Conversely, if the
mass can store is kinetic energy. We can eliminate the vector two charges are of opposite sign, then the force between
notation, by assuming that the force acting on the mass and them is attractive, and the potential energy is defined to be
the velocity of the mass are in the x-direction. negative. The signs of potential energy required the same
16 1  Introduction to System Dynamics

arbitrariness of choice, as Benjamin Franklin made assigning 1.3.7 Power Sources


signs to charges. It will likely surprise you to learn that the
potential energy of a mass in a gravitation field is negative, A common application of system dynamics in machine de-
because gravity is an attractive force between masses. We sign is to size the motor needed to drive a system through a
routinely ignore the negative sign, when we calculate what specified motion. To do this properly, we must derive and
we refer to as the potential energy of a mass raised from the solve the differential system equation, which describes the
surface of the earth. However, what we actually calculate is motion of the machine. We cannot arbitrarily specify both
the change in the gravitational potential energy, as the mass the magnitude of the force acting to accelerate the mass and
is moved further from the center of the earth, which is a posi- the velocity of its point of application, without ­violating
tive change. As it happens, we are generally ignorant of the F = Ma. We will find a similar constraint in every type of
total amount of energy in an object, and are only interested energetic system we investigate.
in the change in the amount. If the sign of change is correct, Power is provided to energetic systems by power sources.
the sign of the potential energy is unimportant. A power source is defined by the power variable, which is
either known or, sometimes, under our control. The most
common examples of a power source are voltage sources,
1.3.6 Power Flows and Signs such as a chemical battery and an electric wall socket. A 12
VDC battery provides a nominal 12 volts at its terminals. An
The sign convention for energy transfers across system electric wall socket provides a nominal 120 volts AC or 240
boundaries varies with engineering discipline and subdis- volts AC, depending on location. Note that in both examples,
cipline. Engineering sign conventions typically define the current is not indicated, only the voltage. The device con-
expected, normal, or preferred quantity as positive and the nected to voltage draws current from the source at a varying
reverse as negative. For example, thermodynamics uses the rate, which is dependent on the nature of the specific device.
sign convention, where the positive direction of the power A battery and a wall socket are “voltage” sources, because
flow depends on the type of energy moving across a system the voltage is known, not because they only provide voltage.
boundary. For a heat engine, heat flow in is positive and work Voltage by itself is not electric power. DC electric power is
flow out is positive. Heat engines cannot be run in reverse. the product of voltage and current.
Turning the drive shaft of an automobile does not produce

fuel. The thermodynamic sign convention of heat in and PDC = vi (1.35)
work out as positive are the expected flow directions. Sys-
tem dynamics uses a general sign convention that inevitably AC electric power is calculated differently, because voltage
conflicts with the sign conventions used in one or more of and current, which vary sinusoidally, peak at different points
the subsystems present in a “hybrid” system, such as an elec- in the oscillation, as we shall see in Chap. 11.
tric motor. Further, unlike heat engines, ideal linear systems We can only specify the value of one of the two power
are reversible. An electric motor can function as generator. variables associated with any power source. If we attempt
Power sources can both provide and accept power. Oscilla- to specify both variables, then we are attempting to dictate
tions can occur, in which power flows periodically reverse the behavior of the energetic elements in the system. The
direction in motors. Although some elements in an energetic behavior of energetic elements is governed by physics, not
system have a conventional positive direction, most do not, by the engineer. System dynamics allows us to predict the
and, in any case, power can flow either way through all of behavior of a dynamic system to a specified input. If we wish
the elements. to change the behavior of the system to that input, we must
The generalized sign convention used in system dynamics change the system. The system will draw power from the
varies with the modeling and analysis method. Sign conven- source at a rate determined by the magnitude of the power
tions must be defined graphically for the results of the analy- variable we control and by the nature of the system.
sis to be interpreted correctly. Once a sign convention is cho- If we measure both power variables during the response
sen, consistency is essential. The sign convention chosen is of the system to the power source, we can then calculate the
applied to all energy transfers, both across a system bound- power drawn from the source by the system. Alternatively,
ary and into and out of an individual energetic element. This during machine design, we may pose the hypothetical ques-
text uses the linear graph method and its sign convention, tion, “What size (power) force source do we need to move
flow in is positive, flow out is negative. This is the same sign the point of application of the force with velocity, v( t), if the
convention applied to bank accounts. We shall see that sign force source provides force, F0?”
conventions within mechanical engineering are essentially We will illustrate power flow calculation and power source
arbitrary. The signs themselves are unimportant, as long as “sizing” with the following example. Again, we cannot dic-
they are used consistently, so that the accounting of energy tate both power variables of a power source without rewriting
transfer is correct. physics. Figure 1.18 shows a hypothetical case, in which we
1.3  Energy, Mechanical Power, and Coenergy 17

Fig. 1.18  a Desired velocity v( t)


of the point of application of the a b 20

force source. b Hypothesized 15


2.0
force F( t) needed to yield v( t) 10
v(t) 1.5 F(t)
kN 5
m
___ 1.0
sec 0
0 1 2 3 4
0.5 -5 t, sec
0 -10
0 1 2 3 4
t, sec

Fig. 1.19  a Segments for veloc-


ity function, v( t). b Segments of a b 20

input force function, F( t) 2.0


15 III
10
v(t) 1.5 F(t) I IV
m kN 5 II
___ 1.0
sec 0
0 1 2 3 4
0.5
I II III IV V -5 t, sec V
0 -10
0 1 2 3 4
t, sec

estimate the force needed to impose a specified velocity at The energy provided by the power source is the integral
the point of application of the force, and then calculate the of power over time.
power the source must provide and the energy required.
Mechanical power is P = F · v, Eq. 1.17. Although it is not t t

stated that the force and velocity of the point of application, E = ∫ P dt → E = ∫ F (t ) v (t )dt
(1.36)
Fig. 1.18, are colinear, we will assume that they are, and add 0 0

that constraint, when we present the results. When the force


and velocity of application of the force are colinear, we can Unlike the power function, P (t ), the energy function, E (t ),
dispense with vector notation, and express mechanical power will be a continuous plot. Power, being the flow rate of en-
as the product of the scalar variables, P = Fv, Eq. 1.31. No- ergy, can change magnitude or sign instantaneously. The
tice that the desired velocity v( t) and the hypothetical force amount of stored energy cannot change instantaneously, be-
F( t) are discontinuous functions, consisting of straight line cause it is a flow quantity, which accumulates in storage. The
segments. The line segments, v( t) and F( t), have the same du- amount of any flow quantity stored in a system or an element
rations. In other words, the discontinuities of the two functions cannot change instantaneously, without an infinite flow rate.
occur at the same times, Fig. 1.19. Physically, the changes in Integration of the power plot can be performed analyti-
velocity are responses to the changes in force. Notice also that cally or graphically. You most likely have performed graphi-
the input force, F( t), consists of horizontal segments of vary- cal integration when creating shear and bending moment
ing widths, corresponding to constant force magnitudes of diagrams. Recall that rectangular areas integrate to ramps or
varying durations. triangles, while ramps or triangles integrate to parabolas.
If the functions for both F( t) and v( t) are known, multi-
plication can be performed to yield the power provided by 1
∫ adt = at + C ∫ at dt = 2 at
2
the source, P = Fv. We shall see in Chap. 3 that discontinu- 1 and + C2
ous functions are constructed by summing scaled and time-
shifted Heaviside unit step functions. The constants of integration represent the amount of energy
For the time being, we will evaluate the product graphi- stored in the system at the beginning of the time interval. If
cally, segment by segment, noting that the result of multiply- we assume that the system is de-energized before the input,
ing a velocity segment by a constant force is a segment of F( t), acts on the system, we can neglect the constants of in-
the power function with the same shape as the velocity plot, tegration.
except when the product is inverted by the negative force Graphical integration is performed by calculating the area
between three and four sec. of each segment of the power plot, summing those areas to
18 1  Introduction to System Dynamics

40 25
35
20
30
(t) 15
25
kW sec
20 10

(t) 15 III
5
III IV V
kW 10 I II
0
5 I II IV 0 1 2 3 4
4 t, sec
0
0 1 2 3
Fig. 1.21   Plot of energy, E( t), provided by the force source, F( t), in
-5 t, sec V order to move the point of application of the source at velocity, v( t).
The units of energy in kW sec can be easily converted to joules,
-10 1 joule = 1 N m = 1W sec
-15
30
-20

Fig. 1.20   Plot of power, P( t), provided by the force source, F( t), in
order to move the point of application of the source at velocity, v( t). The
motor must be sized to provide the maximum power, 40 kW. One horse- 20

power equals 746 W. Converting from kW to hp, 40 kW


1 hp
= 53.6 hp (t)
746 W kWsec
10
calculate the amount of energy in the system at the end of
each time interval, and then connecting the end points of
each interval with the correct line shape for the power seg-
ment integrated. Graphical integration is simple but has the 0
disadvantage that it only yields values at the ends of the time 0 1 2 3 4
segments. Create a table similar to the following: t, sec
Fig. 1.22   Plot of the energy function E( t), produced by analytical inte-
Interval Time-end Shape Area Running Integrated gration of the power function P( t)
points kW sec total shape
I 0 1 Triangle   2.5  2.5 Parabola
t =1
II 1 2 Rectangle   2.5  5 Triangle  kW 
+ ( 2.5 kW ) t =1
t=2
P (t ) =  5 t
III 2 2.5 Rectangle +   15 20 Parabola  sec 
Triangle t =0
t = 2.5
IV 2.5 3 Rectangle   5 25 Triangle  kW
(t − 2 sec) + (10 kW ) t = 2.5
t =3
V 3 4 Triangle −10 15 Parabola +  20 kW + 40
 sec t=2
t=4
 kW
Plot the end points of each segment and “connect the dots” +  −20 kW + 20

(t − 3 sec)
sec t =3
with the correct shape curve. In the case of the parabolas, the
concavity and the rate of change at the two ends of the time
interval must correspond to the triangular area integrated, Notice the “time shifts” ( t − 2) and ( t − 3) created by sub-
Fig. 1.21. tracting the time of the beginning of the interval from the
To calculate the power provided by the source analyti- time variable t. We will introduce time shifts of input and
cally, we must first create the power function, P � (t ), which response functions in Chap. 3. Briefly, subtracting the begin-
we will then integrate to yield energy, Fig. 1.22. P (t ) is the ning time of an interval from the time variable t shifts time
sum of five separate functions, corresponding to the five t = 0 to the beginning of the interval, creating a “local” time
time intervals of Fig. 1.20. for just that integral.
1.4  Network Representation of Energy Flow 19

E (t ) = ∫ P (t ) dt
0
1 2
kW Branch
E (t ) = ∫ 5 t dt + ∫ 2.5 kW dt
0
sec 1
2.5
kW Node
+ ∫ 20 kW + 40 sec (t − 2 sec)dt
2
Fig. 1.23   Networks transfer a quantity between branches connected
3 4
kW at nodes
+ ∫ 10 kW dt + ∫ −20 kW + 20 (t − 3 sec)dt
2.5 3
sec
The energy function is now:
The integrals with time shifts require special handling. We 4
must define a variable equal to the shifted time. We define E (t ) = ∫ P (t ) dt
τ1 ≡ t − 2 sec for the integral, thus, 0

1 2 0.5
2.5
kW kW kW
∫ 20 kW + 40 (t − 2 sec) dt E (t ) = ∫ 5 t dt + ∫ 2.5 kW dt + ∫ 20 kW+40 sec τ 1 dτ1
2
sec 0
sec 1 0
3 1
kW
Differentiating τ1: + ∫ 10 kW dt + ∫ −20 kW+20 τ 2 dτ 2
2.5 0
sec
d τ1 = d (t − 2sec ) → d τ1 = dt − d ( 2sec ) → d τ1 = dt
0
The integration is performed piece-wise, evaluating each in-
Substituting τ1 into the integral and expressing the limits of terval separately.
integration in terms of τ1 yields
1 0.5
5 kW 2  40 kW 2 
E (t ) =
2
2.5
kW
0.5
kW t + 2.5 kW t +  20 kW τ1 + τ1
∫ 20 kW+40 (t − 2 sec) dt = ∫ 20 kW+40 τ1 dτ1 2 sec 0 1
 2 sec  0
2
sec 0
sec 1
 20 kW 2 
+ 10 kW (t − 2.5) 2.5 +  −20 kW τ 2 +
3
τ2 
Likewise, we define τ 2 ≡ t − 3 sec for the integral, thus  2 sec  0

4
kW Back-substitute for τ1 and τ2 and express the limits of those
∫ −20 kW+20 sec (t − 3 sec) dt
3
intervals, in terms of time t:

E (t ) =
1 2.5 1
5 kW 2  40 kW  20 kW
t + 2.5 kW t 1 +  20 kW (t − 2) + (t − 2)2  + 10 kW (t − 2.5) 2.5 +  −20 kW (t − 3) + (t − 3)2 
2 3

2 sec 0  2 sec 2 2 sec 0

Differentiating τ2:

d τ 2 = d (t − 3 sec ) → d τ 2 = dt 1.4 Network Representation of Energy Flow

substituting into the integral and expressing the limits of in- A network is a system, in which elements interact with
tegration in terms of τ2: one another by transferring a quantity via defined paths or
“branches” which join at connections or “nodes,” Fig. 1.23.
4
kW
1
kW Depending on the type of network, either the branches or
∫ −20 kW+20
3
sec
(t − 3 sec) dt = ∫ −20 kW+20
0
sec
τ 2 dτ 2 the nodes may change or store the quantity moving through
the network. Two examples illustrate the range in type and
scope of networks. The Internet is a worldwide network with
20 1  Introduction to System Dynamics

billions of devices which produce or process information. In-


formation flows between devices through various path types.
Information flows are routed to minimize transmission time
and cost at the nodes. A home plumbing system is a network
of at most dozens of elements. Water flows through the sys- Pump
tem without loss, unless there are leaks, but is also stored by
the system. The water in the system may be changed chemi-
cally by the system, if the water is “softened,” or energeti-
cally, if the water is heated. Further, the water flow rate and
flow pressure vary dynamically throughout the system, when Fig. 1.24   A water-filled piping system. The system is horizontal. The
the system is in use. pump increases pressure in the direction of the water flow. The pipes
have fluid resistance. Pressure drops in the direction of water flow
Information networks process information at nodes and through the pipes
transmit information without change through the branches,
which connect the nodes. The nodes of an information net-
work act on the quantity flowing through the network. The of the real system contains two energetic properties such as
branches of an information network conduct (transmit) the kinetic energy storage and strain energy storage, such that
information without change. Conversely, the branches of a both are significant enough to be included in the energetic
plumbing system act on the water by affecting its pressure network, then we will include two branches, with one ele-
and flow rate. Nodes in a plumbing system are locations se- ment for each energetic property. Our network model of an
lected to measure the changes in pressure throughout the sys- energetic system simplifies the representation of the system,
tem and to apply mass and energy conservation. The nodes of while retaining its relevant physical attributes. We will re-
a plumbing system are modeled as conducting water, without turn to the question of how this is done, when we investigate
storing it or changing its pressure. From fluid mechanics, we the modeling of the energetic properties of different types of
know that any redirection of fluid flow imparts a change in components.
pressure. Consequently, in a real plumbing system, there is a
pressure drop, when water flows through connections. Ideal
nodes have no physical properties. They are merely locations 1.4.1 Compatibility and Continuity Equations
in the system. The process of modeling a plumbing system
assigns the physical properties of the pipes and connections Fluid systems are mechanical systems, since they are de-
to branches. The flow of water divides at an ideal node with- scribed in terms of Newton’s laws. Because fluids deform
out a pressure change. The branches between the nodes pos- to fill a volume, Newton’s laws are expressed in terms of
sess physical properties, which affect the water’s pressure pressure rather than force. Consider a horizontal set of inter-
and flow rate. connected pipes with a pump to circulate water through the
The state of energetic systems changes as energy moves system, Fig. 1.24.
through the system. We will refer to the movement of en- If the two following facts are true:
ergy as its flow. In fluid systems, energy flows with mass. In • Water flows in the direction of a decrease in pressure,
electrical systems, energy flows with charge. In translational except through the pump which raises the pressure in the
and rotational mechanical systems made of solid compo- direction of the flow
nents, energy is the only quantity which flows through the • Unless the plumbing system leaks, the water flowing
system. We will extend the circuit representation of fluid and through the plumbing system is conserved
electrical systems to translational and rotational mechanical then the piping system can be described in general terms as
systems by focusing on power, the flow rate of energy, by a network with a potential driven flow of a quantity which
using the linear graph method. A linear graph is a network is conserved.
representation of an energetic system. A potential is the physical variable, which lessens or drops
The energetic networks will consist of energetic ele- in the direction of the flow which it drives. Pressure acts as
ments, represented by branches, which connect at nodes. the driving potential in a piping system. The mass of water
Each branch represents a different, single, lumped, and fun- is the quantity, which flows through the piping system and is
damental energetic property of the system. These idealized conserved. Pressure and mass flow rate are different types of
elements do not directly represent the object or system we variables. We can write two sets of summations, one in terms
are modeling. They represent a single energetic property or of mass flow rate, and the second in terms of pressure for
attribute. In reality, the energetic properties of a real sys- this piping system, knowing only how the pump and pipes
tem are distributed in space. The branches have no physi- are connected, but not knowing details, such as diameter and
cal dimensions, just the energetic property. If a component length of the pipes, or whether pressure is applied by the
1.4  Network Representation of Energy Flow 21

• Control Volume network, which has a potentially driven flow of a quan-


m in
tity which is conserved. Independent refers to mathemati-
cally independent; meaning one set of equations cannot be
derived from the other. Independence between continuity
and compatibility equations is assured, because both sets of
equations are summations written in terms of a single type
Pump of variable, either the flow variable or the potential variable.

m out These two sets of equations are easily written directly from
a network diagram. Both sets of equations apply under any
circumstance. They are always true. Neither set of equations
contains an approximation. Both sets of equations are exact
Fig. 1.25   A control volume cutting the piping system. Mass within the accuracy of the measurements.
conservation can be written in terms of mass flow rate for any control
volume drawn on the piping system

1.4.2 Compatibility Equations
pump. These two sets of equations must always be satisfied,
whether or not the pump is running. Pressure is a scalar variable, since it acts equally in all direc-
If we draw a control volume around any part of the sys- tions. Pressure is also a relative measure. The value of pres-
tem, such that we intersect one or more pipes, Fig. 1.25, then sure at a point has no meaning, until the reference pressure
we can write a statement of conservation of mass in terms measured was relative. Machine design usually uses pres-
of mass flow rate, by differentiating Eq. 1.27 with respect sure measured relative to atmospheric pressure, called gauge
to time: pressure, but there are occasions when absolute pressure
is used instead, where the reference pressure is a complete
d massIn d massOut d massStored vacuum.
(1.37) − =
dt dt dt To describe the behavior of an element in a fluid network,
we need the pressure difference across the element between
Equation 1.37 states that the mass flow rate of water into the nodes at either end. To describe the behavior of a fluid
the control volume, minus the mass flow rate of water out, system, the pressure differences across all of the elements in
equals the rate at which the mass of water, stored inside the the system must “add up.” Before we write any compatibility
control volume, changes. equations, we must indicate in the schematic of the system
In order to systematically extract all of the useful equa- the locations of those pressures of interest to us. The pump
tions from the network, we will find it convenient to draw raises pressure, as it pushes water into the system. There are
control volumes, so that they correspond to single nodes different pressures on either side of the pump. We will al-
where pipes connect. Because we can write mass conserva- ways locate a pressure node where the flow divides or con-
tion equations at individual nodes of the system, these “con- verges. Finally, let us say that the sections of pipe, which join
tinuity” equations are also called “node” equations. On the at the top right corner of the diagram, have different diame-
other hand, a pressure measurement at a single location in ters and, therefore, different fluid resistances. We will locate
the piping system, such as at a node, does not permit us to a pressure node at the location where the diameter changes.
write an equation. We must take pressure measurements at Finally, we will add a reference node for atmospheric pres-
three locations, if we are to write anything but a trivial equa- sure, Fig. 1.26.
tion. (A trivial equation is an identity, such as C = C, which The pressure difference between two points (or nodes) in
is true but of no value in algebra.) Useful equations either the piping system, from node 1 to node 2 can be measured di-
sum pressure changes around a loop in the network (the sum rectly, by connecting a manometer between those pressures,
must be zero); or equate the sum of pressure changes from but is more frequently calculated by measuring the pressure
one node to another node, following two different routes or at each point relative to atmospheric pressure and then sub-
paths through the network. Accordingly, equations written in tracting.
terms of pressure changes between nodes are either “loop”
equations or “path” equations. Equations written using the

( p1 − patm ) − ( p2 − patm ) = p1 − p2 ≡ p1,2 (1.38)
driving potential of an energetic system are called “compat-
ibility” equations, since, for the pressure changes between Notice that the reference pressure patm drops out, since it ap-
nodes to sum properly, they must be compatible. pears in the pressure measurements at both nodes. We will
As it happens, there are always two independent sets use p1 as a shorthand for p1 − patm , since we usually work
of algebraic equations, which can be extracted from any
22 1  Introduction to System Dynamics

2 3 p1,2 2

1
4 1
Pump
patmosphere Pump
5
patmosphere

Fig. 1.26   Piping system with pressure nodes at either end of pipe seg-
ments and on either side of the pump Fig. 1.27   Pressure difference between nodes 1 and 2 is designated
as p1,2 . The two nodes of the pressure difference are identified in the
subscript. The order of the node numbers in the subscript is the order of
with gauge pressure. If no reference pressure is specified, the terms in the pressure difference, i.e., p1 − p2 ≡ p1,2
then atmospheric pressure is implied.
Note that although pressure is scalar, a pressure difference
is directional. For the difference to be meaningful, we must If we measure the difference in pressure between all of
know the direction of the pressure difference, or “drop.” The the nodes, which form a complete loop at any instant in time,
pressure drop is positive, when the pressure of the first node the pressure differences must sum to zero, since we mea-
is greater than the second node (e.g., p1 > p2 in Fig. 1.27). sured our way around a loop back to where we started. If
The pressure difference from node 2 to node 1, Fig. 1.28, the sum does not equal zero, then we have a measurement
equals the opposite of the pressure difference from node 1 error. Summations of the drops in the driving potential of a
to node 2. system around a complete loop are called “loop” equations,
Eq. 1.39.
p2 − patm − ( p1 − patm ) = p2 − p1 ≡ p2,1
p2,1 = − ( p1 − p2 ) = − p1,2
(1.39) p1,2 + p 2,3 + p 3,4 + p 4,5 + p 5,1 = 0
(1.40)

The sign inversion, created by reversing the node subscripts Note that we do not need to know the direction of fluid
on the pressure difference, creates sign errors which prove flow through a network to state that the pressure differences
very difficult to find. We shall see that there must be a posi- around a loop sum to zero. The sum of pressure differences
tive direction defined for the flow through each branch of a around a loop in one direction (clockwise or counterclock-
network. We will customarily use the positive direction of wise) must equal zero, because the pressures at each node
the flow through the branch as the direction of the pressure are measured relative to the same reference, atmospheric
drop. If we need to express the pressure drop in the reverse pressure. Each pressure appears twice, once with a sign
direction, we will indicate it, by placing a negative sign on inversion. Expanding the sum of pressure differences:
the positive direction’s pressure drop.

( p1 − patm ) − ( p2 − patm ) + ( p2 − patm ) − ( p3 − patm ) + ( p3 − patm ) − ( p4 − patm ) + ( p4 − patm ) − ( p5 − patm )
+ ( p5 − patm ) − ( p1 − patm ) = 0
( p1 − p2 ) + ( patm − patm )  + ( p2 − p3 ) + ( patm − patm )  + ( p3 − p4 ) + ( patm − patm )  + ( p4 − p5 ) + ( patm − patm ) 
       
+ ( p5 − p1 ) + ( patm − patm ) = 0
 
 
( p1 − p2 ) + ( p2 − p3 ) + ( p3 − p4 ) + ( p4 − p5 ) + ( p5 − p1 ) = 0
( p1 − p1 ) + ( p2 − p2 ) + ( p3 − p3 ) + ( p4 − p4 ) + ( p5 − p5 ) = 0 → 0=0
1.4  Network Representation of Energy Flow 23

p2,1 2 p1,2 2 p2,3


3

p
3,4
1
1
Pump 4
p Pump
patmosphere 1,5

5 p
atmosphere

Fig. 1.28   Reversing the order of the pressure nodes inverts the sign
of the pressure difference between those nodes p2 − p1 ≡ p 2,1 = − p1,2
p4,5
p1,2 2 p2,3 Fig. 1.30   Arrowheads indicate the positive direction of the pressure
3 drops. The positive direction of the pressure drop from node 1 to node 5
is used in the path equation, Eq. 1.40. Note that the direction of the ar-
rowhead is reversed only for the pressure drop across the pump. Power
p sources, such as pumps, raise the driving potential in the direction of
3,4 the flow through them
1
4 p1, 2 + p 2,3 + p 3, 4 + p 4,5 + p 5,1 = 0
p Pump
5,1
p1, 2 + p 2,3 + p 3, 4 + p 4,5 + p 5,1 − p 5,1 = − p 5,1
5 p
atmosphere
p1, 2 + p 2,3 + p 3, 4 + p 4,5 = − p 5,1

p1, 2 + p 2,3 + p 3, 4 + p 4,5 = p1,5

p4,5 We will now generalize. The driving potentials in fluid sys-


tems (pressure), electric circuits (voltage), and thermal sys-
Fig.  1.29   Positive directions of flow through the pump and piping. tems (temperature) are scalar variables. Summing the dif-
The positive direction of the flow through the elements is used as the
order of the nodes for the pressure drops summed in the loop equation, ferences of the scalar potential variable around a loop in a
Eq. 1.39 system creates a “loop” equation. Alternatively, equating the
sums of the differences in the scalar potential variable be-
tween two nodes along different paths creates a “path” equa-
Pressure drops in the direction of steady fluid flow, except tion. Both loop and path equations are termed “compatibili-
for the flow through pumps. Pumps energize a fluid system ty” equations. The term comes from geometric compatibility
by raising pressure in the direction in which the pump pushes mechanical design. The linear graph method will allow us to
the flow. The positive direction for the flow through the pump represent energetic mechanical systems are networks. When
is shown in Fig. 1.29. However, the pressure drop through we do, we shall see that geometric compatibility provides a
the pump occurs in the opposite direction, from node 1 to set of equations analogous to the compatibility equations of
node 5, as shown in Fig. 1.30. The alternative to a loop equa- fluid, electrical, and thermal systems.
tion is to sum the pressure drops between two nodes along The reduction of individual equations which describe in-
the different paths between them in a “path” equation. The dependent aspects of an energetic network to a differential
path equation for the pressure drop from node 1 to node 5 is system equation which describes the dynamic response of
the system is largely a process of successive elimination by
p1,5 = p1,2 + p 2,3 + p 3,4 + p 4,5
(1.41) substitution, or, in other words, algebra. We will find it very
helpful to eliminate any redundant or dependent equations
from the list of equations we will create, before we begin the
Because reversing the order of the node subscripts inverts reduction. If we use dependent equations in our algebraic re-
the sign of a pressure difference, we can write the loop equa- duction, we will be prone to derive the true but trivial result
tion Eq. 1.39 as the path equation Eq. 1.40, by subtracting 0 = 0, instead of the differential equation we need. It can be
the pressure p5,1 from both sides of Eq. 1.39: difficult to spot redundant or dependent equations, particu-
larly when we work with larger systems. Consequently, we
24 1  Introduction to System Dynamics

2 3
second requirement ensures that the network is a circuit. It
will be more important when we draw network representa-
tions of translational and rotational systems.
There are aspects of a networks topology which are
1 “invariant” and do not vary or change, unless the network
4 Holes is changed. The invariants are the number of nodes and
Pump patmosphere
“holes” in a network, Fig. 1.31. A hole in a network is simi-
5 lar to a hole in a fishing net or window screen. It is a re-
gion surrounded by branches, which cannot be eliminated
by stretching or otherwise deforming the network. Further,
a hole cannot be eliminated by redrawing the network. The
Fig. 1.31   The “topology” of a network are aspects which do not example network has two “holes.”
change, when the network is stretched or deformed, as long as the ele-
ments remain connected. One aspect of the topology is the number of The number of independent compatibility equations that
holes in the network can be written for a network is less than or equal to the num-
ber of “holes” in the network. Often, only a few of the pos-
sible equations are useful, while the remaining are either
will establish a few simple rules to prevent the inclusion of redundant or trivial. The example network has two holes
redundant or dependent equations in our description of the and, therefore, a maximum of two independent compatibility
physics of a network. equations. There are three compatibility equations which can
The “topology” of a network refers to how the branches be written, Fig. 1.32. Only two of the three can be linearly
are connected. We will place only two requirements on how independent.
branches are connected to ensure that our networks will be If the piping system is simplified slightly by removing
proper energetic networks. node three, Fig. 1.33, the number of holes in the network
1. Branches have only two ends, each of which must be con- remains unchanged. However, only one useful compatibility
nected to a node. equation can be written. One of the remaining compatibility
2. There are no dead ends. It must be possible to travel equations is redundant, and the other is trivial.
through every branch in the network without reversing
direction.
The first requirement implies that the flow through the sys- 1.4.3 Continuity Equations
tem can only divide at nodes. Branches do not divide the
flow through by being wye or tee shaped. In other words, Continuity equations are written by applying conserva-
branches do not branch. Branching occurs at nodes. The tion of mass to the flow into and out of the nodes. We will

Fig. 1.32   Only two of the three 2 3 2 3 2 3


paths are linearly independent

1 1 1
4 4 4
Pump Pump Pump
patmosphere patmosphere patmosphere
5 5 5

Path 1-5 = 1-2-3-4-5 Path 1-5 = 1-2-4-5 Path 2-4 = 2-3-4

Fig. 1.33   Three possible path 2 2 2


equations are presented, but only
one is useful. The other two are
redundant or trivial 1 1 1
4 4 4
Pump Pump Pump
patmosphere patmosphere patmosphere
5 5 5

Path 1-5 = 1-2-4-5 Path 1-5 = 1-2-4-5 Path 2-4 = 2-4


1.4  Network Representation of Energy Flow 25

Fig. 1.34   a Branches need


unique identifications. Normally,
a Branch A 2 Branch B b Branch 1,2 2 Branch 2,3
they are named using a param-
eter of the branch. b Identifying
branches by the nodes they are
connected to does not work,
1 1
since parallel branches receive 3 3
the same identification. This
Pump Branch C Pump Branch 2,3
network has two branches named
Branch 2,3 patmosphere patmosphere
4 4

Branch D Branch 3,4

Fig. 1.35   a Each branch in the


system is given unique identifica-
a Branch A 2 Branch B
b •
mA

mB
2
tion. The positive direction for 3 3
each branch is indicated on the •
schematic. b Mass flow rates mD
Branch D
used in continuity equations 1 1
4 •
4

Pump Branch C mPump Pump mC
patmosphere patmosphere
5 5

Branch E •
mE

later find it convenient to model the water as incompress- termed “orienting” the branch, Fig. 1.35. The direction we
ible, and express the equations in terms of volume rather choose as positive for the flow in a branch is arbitrary for
than mass. Continuity equations are written at nodes and most elements. The mathematics will yield the correct flow
are independent of the energetic properties of the branches. direction by inverting the sign of a flow, if necessary; similar
Although there are many types of networks in which nodes to how arbitrarily assigned directions of forces in the mem-
have properties, such as transportation and information net- bers of a truss are corrected by the truss analysis. Power
works, we have defined a node as a location without physi- sources are an exception. The positive direction of the flow
cal properties. In the energetic networks we will work with, is not arbitrary. The positive direction of the flow is in the
the branches represent the different energetic properties of direction the driving potential increases. Motors and power
the system. The nodes exist to connect the branches to one transmission elements are the only other energetic elements
another. Nodes cannot store anything. that have defined positive flow directions, as we shall see in
In order to write continuity equations, we must: Chap. 6.
1. Identify the branches. Recall that we have adopted the sign convention that flow
2. Assign a positive direction to flow in each branch. into a node is positive, and flow out of a node is negative.
We need an identity for each branch, in order to identify the Just as compatibility equations can be written as either loop
flow through the branch. We cannot identify the flow through or path equations, continuity equations can be written in two
a branch by the nodes at either end, because parallel branch- ways: (1) equating the sum of the flows into and out of the
es which are connected between the same nodes have differ- node with zero; or (2) equating the sum of the flows into the
ent flows through them, Fig. 1.34. Normally, we will identify node, with the sum of the flows out. There are five nodes in
a branch with the parameter from the energetic equation for this system. Therefore, we can write five continuity equa-
that branch. For example, if the branch represented a spring tions. We shall see that only four of the five continuity equa-
with spring constant, K, we would identify the branch with a tions are independent. Any one of the five continuity equa-
K. This is the same convention used to identify elements in tions can be derived from the remaining four. Writing the
an electrical schematic. For this example, however, we will mass conservations equations in the form, Σ flow in =  Σ flow
simply identify the branches alphabetically. out, for all five nodes, starting with the flow out of the pump,
Defining the positive direction for flow through a branch, leads to the following:
by drawing an arrow pointed in the positive direction, is
26 1  Introduction to System Dynamics

Fig. 1.36   Block diagram of an Engineering Design Engineering Numerical Design


Objective Identify a Question Create an Model
engineering design decision Analyze Result Interpret Decision
Design Engineering
Model Result
Decision Model

{
Focus of Undergraduate
Engineering Education

m Pump = m A and energy which flows into a node must immediately


m A = m B + m C flow out.
m B = m D The number of useful compatibility equations which can
be extracted from an energetic network is unknown, but
m C + m D = m E
its maximum value is limited to the number of holes in the
m E = m Pump network. The number of independent continuity equations
which can be extracted from an energetic network is one
We can derive any one of these continuity equations from fewer than the number of nodes in the network.
the remaining four. We will demonstrate their dependence
by deriving m E = m Pump from the first four continuity equa-
tions, using elimination by substitution, and starting with the 1.5 Overview of Engineering Modeling
continuity equation, m Pump = m A . and Analyses

m Pump = m A → m Pump = m B + m C The focus of a modern engineer’s education is on developing


the mathematical skills and understanding needed to derive
m Pump = m D + m C → m Pump = m E and evaluate equations based on physical theory. There are
very good reasons to structure an undergraduate engineering
We do not want to include dependent equations in the set of curriculum this way, but the student is often unaware that the
equations we will develop to describe energetic networks, derivation and evaluation of an equation from an existing
because it can greatly increase the effort of reducing the set engineering model are only intermediate steps in the process
to an equation which describes the dynamics of the system. of making an engineering design decision, Fig. 1.36.
Consequently, our rule will be to always write one fewer We use the term, “analysis,” loosely to refer to engineer-
continuity equation than there are nodes in the network. It ing computation. However, engineering computation is not
will be convenient to adopt the habit of omitting the continu- analysis. Analysis is the opposite of “synthesis,” which
ity equation for the lowest potential node. means “to assemble.” Analysis means “disassemble” or,
in an engineering context, “to take a problem or a system
apart,” which is a key step in developing an engineering
1.4.4 Summary of Compatibility and Continuity model. Textbook problems do not teach engineering analysis
Equations since the system has already been taken apart and modeled.
An unfortunate but common result of the introductory
Compatibility equations are mathematical statements that courses of an engineering education is that engineering stu-
the differences between a scalar variable, such as pressure, dents develop unrealistic ideas regarding:
voltage, and temperature, measured between different loca- • Why engineering calculations are performed,
tions in an energetic system must sum to zero, when we add • The accuracy of the information used to perform an engi-
the differences around a loop and return to our starting point. neering calculation,
Likewise, if we measure the differences between the scalar • The uniqueness of the model the calculations are based
variables at different locations along two different paths be- on,
tween the same two starting and ending locations, the two • The confidence which can be placed in the results of engi-
sums must be equal. neering calculations, and
Continuity equations in fluid, electrical, and thermal • The manner in which they should approach a novel or
systems are statements of mass conservation, charge con- unfamiliar problem.
servation, and energy conservation, respectively. Nodes are These engineering student “myths” are eventually dispelled
the locations, where energetic elements connect in ener- by experience after the engineering graduate begins his or
getic networks. Nodes have no energetic properties. Since her professional practice, but it is far better to spare the stu-
they cannot store mass, charge or energy, all mass, charge, dent the pain of learning through error or embarrassment and
1.5  Overview of Engineering Modeling and Analyses 27

expose the myths now. In a nut shell, the study of engineer- successful designs and failures. Most often, the information
ing is quite different from the practice of engineering. regarding failures is not available, either because it does not
Engineering analyses are tools used by engineers to aid exist, since the failure loading was unknown, or it is “pro-
design decisions. The results of an engineering computation prietary,” meaning it is a trade secret. The limiting aspect
allow an engineer to make a rational decision between alter- of empiricism is obvious. Empiricism cannot be applied to
natives, or to validate a decision made using “engineering anything novel, since there is no experience base to build on.
judgment.” Engineering calculations are performed to pro-
vide a basis for engineering design decisions. Any type of
design, whether engineering design, graphic design, or land- 1.5.1 Engineering Modeling and Analysis
scape design, is a sequence of decisions. Design decisions Process
are constrained by economic, physical, and practical factors.
For example, a graphic designer may wish to use a size and The common aspects of the various engineering analyses
quality of paper for a project but cannot without exceeding you have studied was present in Sect. 1.0.6 as a generalized
the project’s budget; a floral designer may wish to use tropi- process of engineering modeling and analysis divided into
cal plants for an installation but cannot, because they would four steps:
not withstand a northern climate; and a mechanical engineer Step 1. Draw a picture(s) of what you have chosen to ana-
may wish to use a manufacturing process that the client does lyze, to aid your engineering decision making. This
not possess. is your model. Annotate the picture(s) to define all of
Additional constraints are imposed by design practice. the variables and parameters in your model.
Designers and engineers of today use design elements which Step 2. Write mathematical statements of relevant physical
may data back years, centuries, or millennia. The design truths.
question, “What diameter should the shaft be?” was asked by Step 3. Reduce the mathematical statements, eliminating
ancient engineers. The shafts we design today usually look all unknown variables except the input and output
like ancient shafts, since they typically have a circular or, variables, and solve the equation (or system of equa-
if not circular, an axially symmetric cross-section such as a tions) for the output variable(s).
square or hexagon. The modern engineer has more materi- Step 4. Evaluate the solution. Iterate if necessary.
als and more methods of shaping or processing materials to We will now elaborate and describe how this process will be
choose from than his or her ancient predecessors. Also, mod- applied to model and analyze dynamic systems.
ern engineers can use physical theory expressed in mathe-
matics to describe materials, and how loads applied to a shaft Step 1. Define the model by drawing a picture  Graphical
of a given geometry affect the materials to aid their design representation is the first step in engineering modeling. If you
decisions, which the ancients did not possess. However, the look back at your courses to date, every engineering analysis
objective remains unchanged: make justifiable decisions as started with a picture. In statics problems, your instructors
to the material and dimensions of the shaft, so that it will stressed the need to draw pictures (i.e., free body diagrams),
serve its purpose at an acceptable cost for a reasonable life by penalizing you points, even if you reached the correct
time. solution. Why such a hard line? Engineering graphics are
Although ancient engineers computed some quantities, not just an embellishment to make engineering computations
such as volumes and weights, their lack of physical theo- more attractive. Engineering graphics establish the model
ry and modern mathematics forced their design decisions and define the parameters and variables. Modeling is the
to be based on empiricism (i.e., use the same diameter as most important step in engineering analysis. A very approxi-
the last shaft with similar loading which worked). Modern mate solution to a reasonable model is helpful, whereas a
engineers still use empiricism. Any change in a successful very precise solution to the wrong model is dangerous.
design requires rational justification. In some cases, empiri- The relevant attributes needed to model the dynamic
cism is used because there is a substantial lag between the characteristics of a system are these: how is energy supplied,
development of a new material or process and development transformed, stored, and dissipated. Most dynamic models
of mathematical theory to describe it. In other cases, empiri- use ideal, linear element equations which assume that the
cism is used because the data needed to apply mathematical way energy is transformed, stored, or dissipated can be de-
theory are unavailable, or so scattered or variable over space scribed by a constant parameter. This is never really true,
or time, that the mathematical theory yields contradictory re- but it is generally reasonably accurate. We typically must
sults. However, reliance on empiricism can be expensive and include more ideal elements in our model than we see in the
limiting. It can be expensive, because the only way to opti- physical object, since our model represents each attribute in-
mize an empirical design, i.e., to create the least expensive dividually; whereas, in reality, all of the attributes are com-
design which serves its purpose, is by comparison between bined in the physical object.
28 1  Introduction to System Dynamics

Modeling is the most interesting step of engineering Engineers do not solve unfamiliar problems with either
analysis because it requires the most judgment. In systems the efficiency or the structure of a textbook analysis. The
dynamics, we will use “lumped parameter models. You have, first step an engineer takes is to ask him or herself, “What
in fact, already drawn many lumped parameter models. Free do I know about this model?” This step typically involves
body diagrams are examples of lumped parameter models, rough sketches and lists of unrelated equations, some general
since physical properties that are actually distributed, such principles, others specific to problem at hand. It is only after
as mass, are represented by a single parameter. Electrical producing these problem-solving resources that the engineer
schematics are also lumped parameter models. For example, attempts to impose some order and logic. It is seldom suc-
electrical resistance is present in all conductors at room tem- cessful, or even helpful, to begin a mathematical derivation
perature. However, in electrical schematics, resistance is immediately.
only indicated for discrete components named resistors in The general method of engineering analysis is to separate
order to keep the lumped parameter model simple enough Step 2, the process of writing equations that represent the
to reduce it mathematically. The wires connecting compo- various physical truths of the system, from Step 3, the re-
nents have resistance also, but we only include “significant” duction of those equations to the input–output relationship.
amounts of resistance in the model. Separating these processes into two distinct steps helps to
How does one judge “significance”? Sometimes, it is ob- both identify all of the relevant physical truths, and ease the
vious because there can be differences by orders of magni- mathematical reduction, as will be demonstrated in the ex-
tude between the amounts of a given physical property in amples that follow.
two elements of a system. At other times, it may not be at What we is need a complete set of independent equations
all obvious, and may require iterating the model. We will which describe the system. In most cases, the statements of
use a rule of thumb for the first model, such that an element physical truth in an engineering problem can be grouped into
which stores or dissipates less than 10 % as much energy/ five independent categories, the order of which is irrelevant.
power as the next smallest similar element in the system can Although the flow or creation of entropy, the fifth category
be neglected. below, is important in reality, our models will not include its
Once we have established a model, the reduction and so- effects, other than viewing energy dissipation in the form of
lution is algorithmic, and in engineering practice is done by heat as irreversible.
a computer. Consequently, modeling is the key step in the
analysis. If the model (whether it is a free body diagram, Independent Physical Truths:
electrical schematic, control volume, or lumped parameter 1. Compatibility of displacements, velocities, voltages tem-
model) is wrong, the entire analysis is at best useless, re- peratures, and pressures. (Loop or Path Equations)
gardless of the sophistication of the reduction and solution. 2. Continuity of current, Conservation of mass, or Equilib-
If the model is correct, one can often make the necessary rium of forces and torques. (Node Equations)
engineering decisions simply by looking at the picture. In 3. Material or Elemental Model. (Constitutive Equations)
more complex situations, you must proceed from a graphical 4. Energy Conservation and Power.
model to a mathematical model. 5. Entropy Creation and Flow.
The graphical model (picture) must be annotated with all
relevant parameters and variables before proceeding to the 1. Compatibility: A fundamental physical truth is that, if you
second step of writing equations. If in Step 2 you find you go around a loop, you get back to where you started. If you
have forgotten a parameter or a variable, do not neglect to measure your progress around a loop in steps, the steps must
add it to the picture. Failure to show all parameters and vari- be compatible, in the sense that the sum of the steps must be
ables graphically results in errors, particularly sign errors, equal to the whole. Fluid systems must have pressure drops
and miscommunication. across components that add to zero, when you go around a
loop. Electrical systems must have voltage (potential) drops
Step 2. Write mathematical statements of relevant physi- across components that add to zero around loops. Thermal
cal truths Many engineering students mistakenly believe systems must have temperature drops that add to zero around
that engineers faced with an unfamiliar problem can think as loops. Mechanical systems must have geometric compatibil-
clearly and efficiently as analyses presented in engineering ity of displacement, velocity, and acceleration. The deforma-
text books. If you have ever asked yourself, “How did the tions of the individual parts must be compatible, in that they
author know to introduce that equation at this point in the must add up to the deformation of the whole object.
analysis?” it may well be that the author didn’t the first time
he or she solved the problem. Engineering texts are revised 2. Continuity, Conservation, or Equilibrium: Another cat-
through many drafts to be as clear and as efficient as pos- egory of fundamental physical truth is that matter and energy
sible.
1.5  Overview of Engineering Modeling and Analyses 29

can either stay put or go somewhere, but they cannot just ed as if they occurred in one dimension. The restriction to
disappear. They are conserved. The conservation principle one dimension allows us to work with scalar equations. The
can be applied by measuring flow rates. In the special case elemental equation for a mass is Eq. 1.29:
of an element that cannot store matter or energy, what flows
in must flow out. In fluid systems, if an incompressible fluid F = Ma
flows in one end of a rigid pipe, it must flow out the other.
In electrical systems, if current flows in one end of a wire, it Rewriting acceleration in terms of the power variable ve-
must flow out the other. In thermal systems, if heat flows in locity yields the differential relationship between the power
one end of a metal bar, it must flow out the other. variables of a mechanical system, F and v, per Eq. 1.30:
Statements of equilibrium in mechanical systems are
analogous to continuity since they involve quantities, forces dv
F=M
or torques, which sum to zero at a point or on a body. We dt
will use “dynamic equilibrium” which includes as a force
the net force which accelerates mass. The net force acting to Note that the time derivative is on the power variable veloc-
accelerate a mass is the opposite of the D’Alembert’s inertial ity which determines the kinetic energy stored in the mass.
force. The inertial force is the force that “pushes back,” when It takes time to transfer energy into or out of a mass, hence
a mass is accelerated. When the net force acting to accelerate the derivative on the energy storage variable for the mass,
a mass, which we will call FM, is included in an equilibrium velocity.
statement, the force summation equals zero. This dynamic Hooke’s law, Eq. 1.6, is the elemental equation for a
equilibrium is a method of accounting for all forces acting at spring,
any point in a system.
F = Kx
3. Material or Elemental Model: Compatibility and Con-
tinuity equations are summations which equal zero. There is Rewriting displacement in terms of the power variable, ve-
only one type of variable in each summation. Force summa- locity, requires differentiating both sides of the equation with
tions have only forces. Pressure summations have only pres- respect to time:
sures. The material or elemental model equations relate two dF dx
different variables by means of a parameter. The equation =K
dt dt
which describes a spring element is familiar F = Kx, where
F is the force in the spring, K is the spring constant (also dx
called the spring rate), and x is the deformation of the spring. and then substituting for = v, yielding a differential equa-
dt
This equation relates two different types of variables, force tion for elastic strain energy storage:
and displacement. The spring constant is the parameter. The
equation for a resistor is another such elemental equation  dF
v = iR, which relates two different types of variables, i.e., = Kv (1.42)
dt
voltage and current. Resistance is the parameter. The con-
cept of a parameter carrying units in order to relate dissimilar Writing the equations for mechanical energy storage ele-
variables is attributed to Fourier. Previously mathematical ments in terms of the mechanical power variables, F and v,
statements were proportionalities, not equations. allows us to develop mathematical analogies between differ-
We will work with elemental equations expressed in ent types of physical systems, because the equations have the
terms of power variables of the system. The product of volt- same form. For example, energy is stored in an electric cir-
age times current is electric power. Hence, voltage and cur- cuit in either the electric field of a capacitor or the magnetic
rent are the power variables for an electrical system. The field of an inductor, Eqs. 1.43 and 1.44,
familiar equations for electrical system elemental equations
are written in terms of electrical power variables. However, dv
(1.43)
i=C
the familiar equations for mechanical systems are not. dt
In system dynamics, we view masses and springs as me-
chanical energy storage elements. We will use unfamiliar di
(1.44)
v=L
forms of the elemental equations, F = Kx and F = Ma, in dt
order to express the equations in terms of the power variables
of mechanical systems. We will also restrict each element to where C is the capacitance in farads, and L is the inductance
motion or deformation in one dimension. If deformations in in henrys. Notice that the equations for the electrical energy
two or more dimensions are coupled, they will be represent- storage elements are also differential equations.
30 1  Introduction to System Dynamics

Compare the elemental equations for a capacitor and an We also see similarity in the mathematical form for these
inductor with those of a mass and a spring: physically different energy storage elements, as seen in their
element equations.
dv dv
Mass: F = M Capacitor: i = C
dt dt 1 1 2
Mass: E M = Mv 2 Capacitor: EC = Cv
1 dF di 2 2
Spring: v = Inductor: v = L
K dt dt 1 1 2 1
Spring: E K = F Inductor: E L = LiL2
2K 2
Even though mechanical and electrical systems are very dif-
ferent, the equations which describe them are not. Energy The linear graph method uses physical analogies to classify
storage elements have a differential relationship between power variables. There are two possible cases. The power
the power variables of the system. The time derivative is on variables are the same at either end of an energetic element,
the variable, which determines the amount of energy stored such as the flow through a pipe, or they are different at ei-
in the element. The coefficient carries the units needed to ther end of an element, such as the pressure drop across a
equate two physically different variables and scales them to pipe. When a power variable is the same at both ends of an
the set of units, SI or US customary, being used. element, it either acts or flows through the element. It is a
The linear graph method uses these analogies between “through” variable. When the power variable can be dif-
different types of physical systems, which will allow us to ferent at the two ends of an element, or changes across an
create a single model to represent devices which contain dif- element, it is an “across” variable. An equally useful set
ferent subsystems, such as an electric motor which has both of analogies groups variables by those which represent an
electrical and mechanical elements. “effort,” and those which represent either “flow” or motion.
This set is used by a different modeling method known as
4. Energy Conservation and Power: We can write math- Bond Graphs.
ematical statements which express the amount of energy Analogies are valuable, because they help leverage our
stored in an element or the rate at which power is dissipated knowledge, so that we may use what we understand, in order
by an element. Energy storage equations always contain to gain insight into what we do not understand. Analogies
an elemental parameter, and all but thermal capacitance also simplify mathematical analysis. Analogies between
and gravitational potential energy have a squared term. For different energetic systems have limits, since they compare
example, the amount of kinetic energy stored in a mass is dissimilar systems. Thermal systems are energetic systems
Eq. 1.34: which have a driving potential or across variable, tempera-
ture, and a through variable, heat flow rate. Although we will
1 be able to represent thermal system using the linear graph
EM = Mv 2
2 method, some analogies and calculations will not be appli-
cable. In all systems except thermal systems, the product of
The amount of energy stored in a spring can be expressed in the through variable of an element and the difference of the
terms of either the applied force or the resulting deformation, across variable at either end is the power flow into or out of
Eqs. 1.11 and 1.12: the element. In thermal systems, power is the through vari-
able, heat flow rate. Thermal systems have only one type of
1 2 F2 energy storage, thermal capacitance, energy stored in mass
EK = Kx =
2 2K as heat. The amount of thermal energy stored in a mass ap-
pears below,
The amount of energy stored in the electric field of a capaci-
tor is the following
 ET = C p MT (1.47)
1
EC = Cv 2
(1.45) in which Cp is heat capacity. Note that temperature T is not
2 squared. We shall see in Chap. 5, that the energetic equations
where C is electrical capacitance. The amount of energy for thermal systems do not follow the same mathematical
stored in the magnetic field of an inductor is the following patterns as those for mechanical, fluid, and electrical sys-
tems.
1
E L = LiL2
(1.46)
2 Step 3. Reduce the mathematical statements,  eliminating
all unknown variables except the input and output variables,
1.5  Overview of Engineering Modeling and Analyses 31

and solve the equation (or system of equations) for the out- Rigid end cap
put variable(s). Bar 1, Area A 1 , Modulus E 1 x
By separating Steps Two and Three, the derivation of the
system equation (i.e., the differential equation that repre-
sents the dynamics of a system) becomes a straightforward, Bar 2, Area A 2 , Modulus E2
F
although sometimes laborious, process of reducing the math-
ematical statements of physical truth to a single equation or a
system of equations. The reduction is an exercise in making
substitutions to eliminate unwanted variables, possibly dif- L
ferentiating with respect to time, but never integrating. It is
literally possible to start the reduction by randomly choosing Fig. 1.37   A mechanical system consisting of two dissimilar elastic
any compatibility, continuity, or elemental equation, with the bars of unloaded length L connected to a rigid end cap. Force F acts
exception of “trivial” equations of the form, x = x. The energy on the end cap
equations are not used to derive the system equation. They
are used to provide the initial conditions to solve the differ-
ential equation, as we shall see. Engineering students have little experience so, at first, they
Note the use of the verb, “solve,” not the verb, “inte- should only expect themselves to identify and reject extreme
grate.” Our purpose is to derive a mathematical model that or absurd results. With time, however, the calculations per-
allows us to predict how a physical system will change with formed as an engineering student serve as the beginning of
time. Our purpose is not to investigate various techniques of professional experience, if the student takes the time to think
integration. We will be able to represent high-order dynamic about the result instead of hurrying onto the next problem.
systems as the superposition of first and second-order re-
sponses. Consequently, we will only need to solve first- and
second-order ordinary differential equations with constant 1.5.2 Engineering Modeling and Analysis
coefficients. The verb, “solve,” means to find the solution. Examples
Because we are dealing with physical systems, we will al-
ready know the forms of the possible solutions. In the case of The general process of engineering modeling and analy-
first-order differential equations, there are only two possible sis is illustrated with two examples. The first example is
responses, and “solving” will be simply identifying which from introductory engineering mechanics, and the second is
response is reasonable, and calculating the appropriate coef- from electric circuits. The systems are clearly very differ-
ficients to use in the standard form. Second-order dynamic ent. However, the general process of modeling and analy-
systems have a much broader range of possible responses. sis is the same. The only difference you may see between
Their solution is, consequently, more involved and will re- how these examples are worked, and textbook examples
quire the use of complex variables. However, we will again you have studied in the past, is the distinct separation of
know the standard form with which we are working, and the the steps of: (1) annotating the engineering graphic of the
solution will involve finding undetermined coefficients or physical model; (2) extracting mathematical statements of
manipulating an equation to match a known Laplace trans- physical truths; and (3) reducing the set of equations to an
formation. input–output relationship. Your first impression may be
It is essential to explore the analytical solution of sec- that this general method is inefficient, which is true in most
ond-order systems to understand the response of dynamic cases. There is nothing more efficient than having a formula
systems. However, once that understanding is gained, it is to yield a needed result. However, working a textbook prob-
much more productive to use numerical solutions. We will lem designed to illustrate the use of a formula and address-
use computational software, either Mathcad or MATLAB, ing a novel or unfamiliar design question are very different
to solve and plot the response of our system equations to a experiences. It is in the latter case, that a general method is
variety of inputs. Mathcad or MATLAB will also allow us to the most efficient.
model higher-order dynamic systems as a set of simultane-
ous differential equations. 1.5.2.1 Example 1:  Mechanical Design Analysis
Design Question. Two dissimilar metals bars with Young’s
Step 4. Interpret the results of the calculation  Every engi- moduli E1 and E2 have the same length L but different cross-
neering calculation requires interpretation. The first question sectional areas A1 and A2, as shown in Fig. 1.37. A force F is
the engineer must ask him or herself is “Do I believe these applied by an ideally rigid member such that the bars elon-
results?” Do they seem reasonable within your experience? gate without rotation. Find the stress in bar one.
32 1  Introduction to System Dynamics

x1 x of physical truths helps organize both the thought and the


presentation of one’s work.
F1 F1

x2 F 1. Compatibility: Geometric compatibility tells us that if the


F2 F2 bars elongate without rotation, then the displacements x1 and
x2 are equal.
L x1 = x2 = x

Fig. 1.38   Free body diagrams of the forces acting on the bars and on 2. Continuity, Conservation, or Equilibrium: The free
the rigid end cap, as well as displacements of the ends of the two bars
and the rigid end cap
body diagram allows us to write an equilibrium equation.

F = F1 + F2
Step 1. Draw and annotate a picture (or pictures)  The
geometry of the structure is shown Fig. 1.37. We need to 3. Material or Elemental Models: We can relate stress and
draw free body diagrams of the bars and the end cap, in order strain in both materials.
to define the axial forces in each bar, establish an equilib-
rium relationship between the external force F applied and σ 1 = E1ε1 σ 2 = E2 ε 2
the forces applied by the bars to the end cap, and to establish
the relationship between displacement of the end cap and the We can also write equations that relate displacements to
displacements of the two bars. We would also draw free body strains and forces to stresses
diagrams of the individual bars, if we needed to determine
the reaction forces, Fig. 1.38. x1 x2
ε1 = ε2 =
It is essential to recognize that engineering models are L L
conveyed by the annotated engineering drawings we begin an F1 F
σ1 = σ2 = 2
analysis with. Figure 1.37 implies two important constraints, A1 A2
which greatly simplifies the analysis. The constraints are not
stated. They are implied by omission. and write equations for the spring constants (or spring rates)
1. All of the vertical dimensions needed to calculate the of the bars.
moment on the end cap due to the applied forces are omit-
F1 F2
ted. If this were a real system, the end cap could rotate, K1 = K2 =
unless external force F was applied at a point, such that x1 x2
the moments created by the bar forces F1 and F2 were
equal and opposite. Moment equilibrium for the end cap 4. Energy and Power: The amount of strain energy stored
would be provided by the rigid connections between the each bar is
bars and end cap and the bars and the support. We will
accept the constraint of purely horizontal displacement. 1 1
E1 = K1 x12 E2 = K 2 x22
2. The external force F is not defined. The mechanical sys- 2 2
tem is an elastic structure which stores strain energy. We
concluded in Sect. 1.2, that it is impossible to instanta- The total amount of energy stored in the system is the sum of
neously impose a finite force on an elastic object, with- the strain energy in both bars.
out providing infinite power. Consequently, there is an
implied dynamic constraint that, either the external force E = E1 + E 2
F( t) is applied gradually, i.e., ramps up from zero, to limit
the power needed or the analysis is only valid in “steady- Because the system is elastic, the total energy in the system
state,” after all of the transient fluctuations in force and can be written in terms of the applied force F and the dis-
displacement have ended. placement x:

Step 2. Write mathematical statements of physical 1


E= Fx
truth  A systematic approach to identifying physical truths 2
which can be expressed as mathematical statements makes
this step much easier. Typically, when students are stuck on a Note that the elastic energy E ≠ Fx because F is not con-
problem, it is because they have omitted one or more physi- stant. F increases linearly with displacement.
cal truths. A mental checklist of general categories or types
1.5  Overview of Engineering Modeling and Analyses 33

Step 3. Reduction and Solution  Having written mathemat- tion in your list with F2 in it, and use it to eliminate F2. We
ical statements for all physical truths we can from the engi- will use
neering model, the remaining effort is one of pure algebra. F2
We will “reduce” the set of equations we created, using elim- σ2 = → F2 = σ 2 A2
A2
ination by substitution. It is helpful to state the input variable
acting on the system and the variable desired as the result substituting
or output of the analysis. The input variable acting on the
system is the external force F. The desired output variable is F − σ 2 A2
σ1 =
the stress in bar number one, σ1. Our objective is to produce A1
an equation with only the variables, F and σ1. Parameters can
remain, since, in fact, they are necessary, but no variables, A2 is a parameter. It remains in the equation. We now substi-
other than the input F and the output σ1, are permitted in the tute to eliminate σ 2 using
final equation.
Start the reduction by choosing any equation in the list, σ 2 = E2 ε 2
other than a trivial equation. Trivial equations are algebraic
dead ends which lead to the result, 0 = 0. There is no “right” which yields
equation to start with. If there is an equation with the input
F − E 2 ε 2 A2
and output variable in it, that would be a good choice. Oth- σ1 =
erwise, start with an equation with either the input or output A1
variable. We will reduce the equation list twice, starting with
two different equations to demonstrate that it does not matter We have introduced a parameter, E2, and a variable, ε 2 . Elim-
which non-trivial equation we start with. inate ε 2 by substitution. Use the definition of strain ε2
x2
Reduction One: Input F, Output σ1. Let’s start with this ε2 =
L
equation:
to yield
F
σ1 = 1 x2
A1 F − E2 A
σ1 = L 2
The only variables allowed in our result are the input F and A1
the output σ1. A1 is a parameter. It stays in the equation. F1
is an unwanted variable. We will substitute to eliminate it. We have again introduced a parameter, L, and a variable, x2.
Look in list of equations for one with F1 in it and use it. It Eliminate x2 by substitution. Use of
doesn’t matter which equation with F1 we use. We can use
the same equation twice if needed, but it we immediately use x1 = x2
the same equation, we undo the previous substitution. There
are three equations in the list which contain F1: yields
x1
F F F − E2 A
F = F1 + F2 σ1 = 1 K1 = 1 σ1 = L 2
A1 x1 A1

We will use Eliminate the displacement x1. Using the definition of strain
ε1
F = F1 + F2 → F1 = F − F2 x
ε1 = 1 → x1 = ε1 L
L
which yields
yields:
F − F2
σ1 = ε1 L
A1 F − E2 A2
L F − E 2 ε 1 A2
σ1 = =
We have introduced the input F and the bar force F2. The A1 A1
input F stays in the equation. F2 is unwanted. Find an equa-
Eliminate the strain ε1 : Use Hooke’s law
34 1  Introduction to System Dynamics

σ1 Was this the most efficient solution? Probably not, but


σ 1 = E1ε1 → ε1 =
E1 that is irrelevant. The point is that once you have written a
complete set of independent mathematical statements for all
to yield relevant physical truths for a system, then you can solve for
any output variable.
σ1
F − E2 A2
E1 Reduction Two: Input F, Output σ1. This time we will start
σ1 =
A1 with

We now have an expression that has only the input variable F = F1 + F2


F, the output variable σ 1 , and parameters. We are done sub-
stituting. We are not yet in a useful standard form. We rear- Eliminate F1 and F2, using the definition of stress
range this equation to isolate the output variable on one side.
This process of rearranging is known as “solving” for a vari- F1 F2
σ1 = → F1 = σ 1 A1 and σ 2 = → F2 = σ 2 A2
able, since it yields an equation for that variable. A1 A2

E2 yielding
F− A2σ 1
E1 F E 2 A2
σ1 = → σ1 = − σ1 F = σ 1 A1 + σ 2 A2
A1 A1 E1 A1
E 2 A2 F  E 2 A2  F Eliminate σ2, using Hooke’s law σ 2 = E2 ε 2 .
σ1 + σ1 = → 1 +  σ1 =
E1 A1 A1  E 1 A1 A1
F = σ 1 A1 + E2 ε 2 A2
x2
We wish to present our final result as a proper ratio, not as a Eliminate ε2, using the definition of strain ε 2 = .
ratio of ratios. If we divide both sides to clear a factor from L
one side, we will then have ratios to clear later. Rather than x2
divide, we will save effort and reduce errors, if we place the F = σ 1 A1 + E2 A2
L
left side term over a common denominator, so that we can
clear it from the output variable σ 1 by multiplying both sides Eliminate x2, using x1 = x2 .
by its inverse.
x1
 E 1 A1 E 2 A 2  F  E 1 A1 + E 2 A 2  F F = σ 1 A1 + E2 A2
L
 +  σ1 = →   σ1 =
 E1 A1 E 1 A1  A1  E 1 A1  A1
x1
Eliminate x1, using the definition of strain ε1 = .
 E 1 A1   E 1 A1 + E 2 A 2   E 1 A1  F L
    σ 1 =  
 E 1 A1 + E 2 A 2   E 1 A1   E 1 A1 + E 2 A 2  A1 F = σ 1 A1 + E2 A2 ε1

 E1  Eliminate ε1, using Hooke’s law σ 1 = E1ε1 .


σ1 =  F
 E 1 A1 + E 2 A 2 
σ1
σ 1 = E1ε1 → ε1 =
Now, check the equation. Always check units. If the units E1
of any term in the equation are wrong, then the equation is σ1
F = σ 1 A1 + E2 A2
wrong. These units check, since the unit of stress is force/ E1
area. Now check for reasonableness, since the equation can
be wrong, even if the units are correct. This equation tells us The only variables are the input F and the output σ1. We are
that the stress in bar one increases, if either the area of the finished substituting. Collect the output variable σ1, and ex-
bar decreases, or the modulus of the bar increases. Is this press the result in a useful form:
reasonable? Yes.
1.5  Overview of Engineering Modeling and Analyses 35

open other electrical circuit components are identified by their


switch R L dominant electrical property. The elements in an electrical
schematic are connected by conductors which are modeled
+ as ideal conductors, which carry the current needed without
loss of power. Actual components are connected by either
C copper wiring or copper circuit board “traces,” both of which
are highly but not ideally conductive. They must be sized in
- cross-section to carry the current needed without excessive
resistance.
Fig. 1.39   An electric circuit consisting of resistor R, inductor L,
capacitor C, a battery, and a switch Step 1: Draw and annotate a picture  Electrical schemat-
ics are lumped parameter models of electric circuits, so we

σ1  E2 A 2   A1 E1 E2 A 2 
F = σ 1 A1 + E2 A 2 → F =  A1 +  σ1 → F =  + σ1
E1  E1   E1 E1 

 A1 E1 + E2 A 2   E1   E1   A1 E1 + E2 A 2 
F=  σ1 →  F=     σ 1
 E1   A1 E1 + E2 A 2   A1 E1 + E2 A 2   E1

E1
σ1 = F
A1 E1 + E2 A 2

1.5.2.2 Example 2: Electric Circuit, a Dynamic do not have to draw a picture. However, we do have to anno-
System tate the picture to show all of the variables and parameters
An electric circuit consisting of a battery, a switch, a resis- used in our equations. We must identify the nodes between
tor with resistance R, an inductor with inductance L, and the elements. We must also show an assumed positive direc-
a capacitor with capacitance C is shown in the schematic, tion for the current flow through each of the elements. The
Fig.  1.39. Derive a differential equation which relates the positive direction is arbitrary, except for the battery which
battery voltage to the voltage across the capacitor. is the power source. The negative and positive terminals are
We will investigate the electrical phenomena and elec- identified by the short and long bars of the “pile” symbol in
tric circuits in Chap. 5. For the time being, we will work addition to the plus and minus signs. We will identify the
with them as energetic systems presented as a network (i.e., current through each individual element. Individual currents
in a circuit diagram). Current i is the quantity, which flows are not necessary in this simple circuit, because all of the
through the network elements, and voltage v is the driving elements have the same current flowing through them, since
potential, which drops in the direction of the flow. The ter- there is only one loop. Currents flowing through an element
minology, “driving potential,” comes from electric circuits. will be identified by the element’s parameter as a subscript.
Electrical schematics are lumped parameter models of For example, the current through resistor R will be identified
electric circuits. The elements are identified by their param- as iR. The exception will be the current flowing through a
eter and by a graphic symbol. Components are modeled as source which does not have a parameter.
having a single electrical property. This is only a model. All There are two ways to denote the potential or voltage
components possess the ability to dissipate and store energy. drop between nodes in an electric circuit. One can either
Both electric and magnetic fields store energy. Electric fields use a subscript which indicates the element across which
are created by electric charge. Magnetic fields are created the voltage drops, e.g., vR, to indicate the voltage dropping
by electric charge in motion. Consequently, some energy is across resistor R, or, alternatively, two subscripts to indicate
stored in any an electric current. There is also some energy the nodes of distinct voltages between which the element
lost with the flow of electric charge, which manifests as a is connected. We will use the node notation that we intro-
voltage drop in the direction of the current. duced in Sect. 1.4.2 using the piping system. Although the
A component is modeled as a resistor, if its electrical node subscripts are more complicated than parameter sub-
resistance is its dominant electrical property. Likewise, the scripts, the node notation yields benefits when working with
36 1  Introduction to System Dynamics

closed compatibility equations we write, because we only want in-


switch R 2 L 3 dependent equations. The maximum number of useful com-
1 patibility equations equals the number of “internal” loops in
+ the circuit. Internal loops are the smallest loops; the holes in
iR iL the network. If we write more compatibility equations than
i iC C we have internal loops, we generate dependent equations.
Dependent equations are equations that can be derived from
- equations we already have. Dependent equations give you
the familiar but useless result of 0 = 0.
g g
There is one internal loop in this circuit. Using nodal sub-
Fig. 1.40   The electric circuit of Fig. 1.39 annotated with voltage nodes scripts, the compatibility equation is
between the elements, and the positive direction of current flowing
through the elements v1g = v12 + v23 + v3g

2. Continuity Equations (Node Equations): These equa-


complicated systems. Specifically, node subscripts elimi- tions express the fact that, because nodes are geometric lo-
nate variables and equations, when elements are connected cations and not physical elements hence have no physical
in parallel. Recall the convention introduced in Sect. 1.4.2, properties, nodes cannot store electric charge. Current that
wherein the order of the subscripts is the order of the subtrac- flows into a node must flow out. Once again, we use the sign
tion to calculate the difference in the driving potential. e.g., convention, and define flow into a node to be positive and
flow out of a node to be negative. We need one fewer node
v1 − v2 ≡ v12 equation than there are nodes in the system, to avoid creat-
ing a dependent equation. By convention, we will not write
Also recall that reversing the order of the subscripts inverts a node equation for the ground node. The node equations for
the sign of the potential or voltage drop. this simple circuit do not convey much information, since
the current is the same through each element. In the general
v12 = −v21 since v1 − v2 = − ( v2 − v1 ) case, however, the current though each element is unique.

The annotated schematic showing the nodes of distinct val- Node 1: iSource = iR Node 2: iR = iL Node 3: iL = iC
ues of potential and the assumed positive direction for cur-
rent through each element is Fig. 1.40. 3. Elemental Equations: The elemental equation must re-
The two nodes identified as g for “ground” have the same spect the assumed positive direction of the current in the ele-
voltage because the conductor between them is modeled as ment indicated on the schematic. The potential or voltage
ideal. There is no voltage drop in the direction of current drop in the elemental equation is in the direction of the as-
flow through an ideal conductor. If two nodes always have sumed current. First, write the expression for the type of ele-
the same voltage, then they are, in fact, the same node, as ment, and then edit it to describe the specific element in this
indicated by the dashed oval on the circuit diagram. Ground circuit, by adding the element’s parameter as the subscript to
voltage is an arbitrary reference voltage. It may or may not current, i, and the nodes the element is connected to as the
correspond to “earth ground,” which is the voltage of the subscript to voltage, v. The node subscripts on voltage must
earth at the location of the circuit. An alternative reference be ordered in the positive direction for the current through
voltage is “chassis ground,” where the “chassis” is, or used the element, which is also the positive direction for a voltage
to be, the mechanical support of the circuit components. drop across the element.
Chassis ground now commonly refers to the voltage refer-
ence of the power supply of the circuit. The ground reference Resistor: v = Ri → v12 = RiR
g in Fig. 1.40 is the negative terminal of the battery; it is the di diL
Inductor: v = L → v23 = L
chassis ground. dt dt
dv dv3 g
Step 2. Write mathematical statements of physical truth.  Capacitor: i = C → iC = C
dt dt
1. Compatibility Equations (Loop or Path Equations): 4. Energy and Power Equations: There are two energy stor-
These equations express the fact that voltage drops around age elements in this circuit, the capacitor and the inductor,
loops must sum to zero. It makes no difference where you and one dissipative element, the resistor. We can write the
start the summation. It does make a difference how many equation that the energy stored in the system is the sum of
1.5  Overview of Engineering Modeling and Analyses 37

the energy stored in the inductor and the energy stored in or the output variable. We will start with the compatibility
the capacitor. We can also write equations for the amount of equation:
energy stored in the inductor and in the capacitor. Finally, we
can write an equation for the power dissipated in the resistor. v1g = v12 + v23 + v3g
In Chap. 3, we will use energy equations to derive the initial
conditions we need to solve the differential system equation. We keep the input and output variables, v1g and v3g. We must
As we will explore in depth, the energy storage variables eliminate v12 and v23. We can substitute the resistor elemental
define the energetic state of an energetic system, and will be equation for v12.
called the “state” variables.
v1g = RiR + v23 + v3 g
Energy Stored in the System: E sys = E L + EC
Now we must eliminate iR. We use the continuity equation
1 2
Energy Stored in the Inductor: E L = LiL for node 2, iR = iL , to eliminate iR:
2
1
Energy Stored in the Capacitor: EC = Cv32g v1g = RiL + v23 + v3 g
2
Power Dissipated in the Resistor: PR = iR v12 and the continuity equation for node 3, iL = iC , to eliminate
iL:
5. Reduction and Solution. The only difference between the
reductions for the previous example of the elastic structure v1g = RiC + v23 + v3 g
and this circuit is that the elemental equations for the capaci-
tor and inductor are differential equations. Consequently, Introduce the derivative of the output variable v3g by substi-
the resulting system equation will be a differential equation. tuting in the capacitor’s elemental equation
The order of the differential system equation will equal the
dv3 g
number of independent energy storage elements; two in this iC = C
case. We will return to the question of what makes an energy dt
storage element in a system independent. For the moment, to eliminate iC.
when the two energy storage variables are different types of
variables, current iL and voltage v3g, as in this case, then the dv3 g
v1g = RC + v23 + v3 g
energy storage elements are independent. dt
There are three firm rules to follow during the reduction
of a set of equations a differential system equation: Voltage v23 is eliminated, using the inductor’s elemental
1. Do not use the energy equations to derive the system equation:
equation. Energy equations are integrals. We are deriving diL
a differential equation. The energy equations are used to v23 = L
dt
provide the initial conditions needed to solve the differen-
tial system equation, but not to derive the equation. yielding
2. Never integrate during the reduction, only differentiate.
The system equation is a differential equation, not an in- dv3 g diL
v1g = RC +L + v3 g
tegral equation. dt dt
3. If you reach the end of the reduction, and you have not
introduced all of the elemental parameters, then there is Current iL is eliminated, by again using the continuity equa-
an error. tion for node 3, iL = iC ,
The reduction technique is elimination by substitution. The
only difference we will see from the previous mechanical dv3 g diC
v1g = RC +L + v3 g
example is that we may have to differentiate both sides of the dt dt
equation with respect to time, in order to make a substitution.
We can randomly pick any equation from our set of Current iC is eliminated, by again using the element equation
equations, except a trivial equation, and start substituting to dv3 g
eliminate all unwanted variables. Our input variable is the for the capacitor, iC = C ,
voltage across the battery, v1g, and the output variable of in- dt
terest is the voltage across the capacitor, v3g. It is generally dv3 g d  dv3 g 
v1g = RC +L C + v3 g
easiest to start with an equation that contains either the input dt dt  dt 
38 1  Introduction to System Dynamics

We have finished substituting. We now apply the differential Distribute the units of operator onto the terms of the summa-
tion on the right side.
d  dv3 g 
operator to  C , yielding
dt  dt   1   d v3 g   R dv3 g   1
2

dv3 g d 2 v3 g  LC 1g   dt 2  +  L dt  +  LC v3 g 
v =
   
v1g = RC + LC + v3 g
dt dt 2
All of these terms must have the same units as the single
This is a second-order differential system equation. Put it term on the left-hand side. Second-grade math applies to
into standard form for higher-order differential equations, by differential equations. You cannot add apples and oranges.
arranging the terms in order of decreasing differentiation and First, simplify the expressions, by dropping the subscripts on
clearing the coefficient of the highest-order derivative. the variables, and by evaluating the units of the derivatives.

d 2 v3 g dv3 g  1   v  R v  1 
v1g = LC 2
+ RC + v3 g  LC v  =  t 2  +  L t  +  LC v 
dt dt

1 d 2 v3 g R C dv3 g 1 What remains is an expression consisting of the variables,


v1g = + + v3 g
LC dt 2
L C dt LC v and t,and the elemental parameters. If our purpose was a
dimensional analysis, then we would express the parameters,
1 d 2 v3 g R dv3 g 1 R, L, and C, in terms of fundamental units. However, it is
v1g = + + v3 g
LC dt 2 L dt LC not our purpose to perform a dimensional analysis. We sim-
ply wish to check the consistency of the units of each term
We will review the method of undetermined coefficients in the equation. We will find that performing a dimensional
and the use of Laplace transformations for solving ordinary analysis becomes a challenge, when we work with motors
differential equations with constant coefficients in Chap. 2. and other “hybrid” systems which contain two or more dif-
However, before we invest time in solving an equation, we ferent types of energetic subsystems.
will check the consistency of its units. If the units are in- The easy way to check the consistency of the units of the
consistent, then the equation is erroneous. Remember that system equation is to express the units of the elements pa-
checking for unit consistency only reveals some errors. The rameters in terms of the power variables of the system and
equation can still be wrong, even if the units of each term are time. By rearranging the element equation for the resistor,
consistent. inductor, and capacitor to solve for the parameter, and then
d
What are the units of the derivative operator, ? Recall applying the units of operator [ ]:
dt
when the derivative was introduced in Calculus I. The op-
erator d represents an infinitesimally small change. It has no Resistor:
d 1
units. Consequently, the units of are . Square brackets v v
dt t v = Ri
(1.50)→ R=
i
→ [ R ] =  i 
are used to denote “the units of,” so the previous sentence  
can be expressed symbolically as
Inductor:
 d  1 
(1.48)
 dt  =  t  di dt  vt 
(1.51)
v=L → L=v → [ L] =  
dt di i 
Similarly, the units of the second derivative with respect to
time are Capacitor:
d  1 2
(1.49) dv dt it
 2= 2
 dt   t 
i=C
(1.52)
dt
→ C=i
dv
→ [C ] =  v 
 

The units of the differential system equation are checked by We have expressed the units of the element parameters in
applying the units of operator [ ] to both sides of the equa- terms of the electrical system’s power variables, i and v, and
tion. the independent variable time t. Substituting the units of the
parameters into the differential equation:
  d v3 g R dv3 g 
2
 1 1
v
 LC 
1 g =  2
+ + v3 g   1   v  R v  1 
 dt L dt LC   LC v  =  t 2  +  L t  +  LC v 
Summary 39

yields • The accuracy or reliability of the model or the data the


model uses has reached its useful limit.
v  Engineers are typically more visual than verbal people. Most
 i v  v  i v  i v  engineering design is fundamentally graphical. Although
 v =  2  +  + v
 vt it   t   vt t   vt it  practicing engineers need to understand calculus and dif-
 i  ferential equations to understand the physical processes and
design methods we use, it is more important to be able to
Cancel terms, but be careful with improper ratios. visualize the mathematical relationships than it is to remem-
ber formulas. The vast majority of the mathematics done by
  practicing engineers is arithmetic with a small amount of
 v  i 
 i v  v    algebra. The primary challenge of engineering mathematics
v =  2  +  i v  v t   +  i v v
   is organizational not conceptual. If you have been relying
 v t it   t   v t t  i    v t it 
   upon example problems in texts rather than understanding
 i  v t   the sequence of steps needed in a design calculation, then
now is the time to develop a method to guide your thought.
The sequence of a lengthy series of calculations is often best
 i v   v  v v i   i v 
 v t it v  =  2  +  i t v t  +  v t it v  represented graphically. Engineers need pictures. Learn how
  t      to use them to your advantage.
The steps in engineering analysis are to:
 i v  v v v i   i v  1. Draw a reasonable model, annotated to show the relevant
 v =  2  +  + v variables and parameters. Never write an equation which
 v t it  t   i t v t   v t it  contains a variable or parameter that has not been defined
on a graphic.
v v v v 2. Write a complete set of independent mathematical state-
 t 2  =  t 2  +  t 2  +  t 2  ments of physical truth relevant to that model.
3. Reduce the mathematical statements to relationships
The units of each term are consistent. which relate what we know (the input) to what we want to
know (the output).
4. Evaluate the solution. Iterate if necessary.
Summary Begin your transition from engineering student to practicing
engineer now.
Chapter 1 introduced system dynamics as the study of en- • A model represents only one (or some) of the attributes of
ergetic systems; systems in which the flow of energy from the subject object, system, or process.
a power source into and out of energy storage and dissipa- • The model is the basis of the subsequent calculations.
tion elements produces a dynamic response of the power Therefore, it is of utmost importance to be confident that
variables of that system. Energy is a quantity which can the model represents the actual physical object, system,
move, be stored, change form, and be dissipated as heat. or process with sufficient accuracy for the intended use of
Mechanical energy storage elements, the ideal spring which the analytical results.
stores elastic strain energy, and the ideal mass were used to • Engineering models can be physical, analogous, or math-
introduce energy and coenergy. Ideal springs and masses ematical. It is important to include a graphical representa-
were also used to introduce the concept of energetic models tion of the model, and to update it as the model evolves.
in which an energetic element possesses a single property • Calculations yield results. There are no “answers” in the
or attribute. Modeling of physically real systems which are sophomoric sense, because the results vary as the model
made of components that possess a number of different en- is refined.
ergy properties is the subject of the remainder of this text. • Every engineering calculation requires interpretation, i.e.,
The process of engineering modeling and analysis was “What does this result mean?”
also discussed. It is an iterative process. It begins with a • Although analysis is challenging for an engineering stu-
simple model of the part, machine, or system. The model’s dent to learn, synthesis is more challenging for the prac-
complexity is increased, as insight is gained from the results ticing engineer because the choices are open ended. Anal-
of the previous iteration, until one of the following occurs: ysis is algorithmic. Design is synthesis.
• You have the information you need to make the engineer- • Increasing understanding, knowledge, and experience
ing decision, or increases the design options of the engineer.
• You have run out of time or money, or
40 1  Introduction to System Dynamics

Fig. P1.1  a Force F( t) acting on


a mechanical system. b Colinear
a 200 b 2.0
velocity v( t) 150 1.5
v(t)
F(t) 1.0
100 m
___
N sec
50 0.5
0 4 0
0 1 2 3 0 1 2 3 4
-50 t, sec t, sec
-100

Fig. P1.2  a DC current i( t) driv-


ing a motor. b Voltage v( t) of the a 20 b 200
power supply 15 150
i(t) 10 v(t) 100
A VDC
5 50
0 0
0 2 4 6 8 0 2 4 6 8
-5 t, sec -50 t, sec
-10 -100

Problems Branch A 2 Branch B


Problem 1.1 Translational mechanical power is the dot
product of force and the velocity of the point of application
of the force, P = F · v. When the force and velocity of the 3
1
point of application of the force are colinear, the scalar equa- 4
tion P = F v can be used. A mechanical system was de-ener- Pump Branch C Branch D
gized before the force shown in Fig. P1.1a with the colinear
velocity shown in Fig. P1.1b acted on it. patmosphere
5
1.1.a Plot the energy provided by the power source from 0
to 4 sec.
1.1.b What is the minimum horsepower motor which could
provide the power n­ eeded? Branch E
Fig. P1.3   Piping system schematic. The fluid is modeled as incom-
Problem 1.2 DC Electrical power is product of current pressible. There is fluid resistance in each branch, which decreases
times voltage, P = iv. A DC motor and the system it drives pressure in the direction of the fluid flow
was de-energized, before the DC motor’s power source pro-
vided it the current i at the voltage v, shown in Fig. P.1.2a
and b. 1.3.a O rient the flow in each branch. The positive direction
1.2.a P lot the energy provided by the power source from 0 for flow through pump is from node 5 to node 1.
to 8 sec. 1.3.b Write a complete set of independent compatibility
1.2.b What is the minimum horsepower motor which could equations in the form of path equations.
accept the power from source? 1.3.c Write a complete set of independent continuity equa-
tions for volume flow rate.
Problem 1.3 A piping system is shown in Fig. P1.3. The
fluid is modeled as incompressible. Consequently, the conti- Problem 1.4 A piping system is shown in Fig. P1.4. The
nuity equations can be written in terms of volume flow rate, fluid is modeled as incompressible. Consequently, the conti-
Q, rather than in terms of mass flow rate. The branches are nuity equations can be written in terms of volume flow rate,
identified by letter, and pressure nodes between the branches Q, rather than in terms of mass flow rate. The branches are
are numbered.
Problems 41

Branch A 2 Branch B 80,000

patmosphere 60,000

1 3 F, lb
4 40,000
Pump Branch C Branch D

5 20,000
Branch E
0
Branch F 0 1 2 3 4
x, in
Fig. P1.4   Piping system schematic. The fluid is modeled as incom-
pressible. There is fluid resistance in each branch, which decreases Fig. P1.6   Force in pounds vs. displacement in inches. The test results
pressure in the direction of the fluid flow  − 
x

( )
are approximated by F (t ) = 7 × 104 lb 1 − e in 
 

closed
switch R 2 L 3
σyield 1
+
iR iL
Stress, σ i source iC C
-
g g
Fig. P1.7   Electric circuit consisting of a battery, resistor, inductor,
ε yield Strain, ε
ε rupture and capacitor in series, annotated with voltage nodes and the positive
directions of current through the elements
Fig. P1.5   Stress–strain plot of a specimen tested to rupture. The speci-
men was 0.5 in. in diameter and 3 in. long. The yield stress, yield strain,
and rupture strain are σ yield = 90ksi ε yield = 0.18% ε rupture = 1.02%
Problem 1.6  The force vs. displacement plot of a load test
is shown in Fig. P1.6.
1.6.a C
 alculate the energy transferred to the specimen for
identified by letter, and pressure nodes between the branches the displacement, x = 4 in., in foot-pounds and joules.
are numbered. 1.6.b Calculate the coenergy for the displacement, x = 4 in.
1.4.a Orient the flow in each branch. The positive direction
for flow through pump is from node 5 to node 1. Problem 1.7 An electric circuit consisting of a voltage
1.4.b Write a complete set of independent compatibility source, represented as a battery, a switch, a resistor, an in-
equations in the form of path equations. ductor, and a capacitor, is shown in Fig. P1.7. The energetic
1.4.c Write a complete set of independent continuity equa- equations of this circuit are listed. Use elimination by substi-
tions in terms of volume flow rate, Q. tution to derive the system equation for the input voltage and
output variable indicated:
Problem 1.5  The stress vs. strain plot of a specimen tested i. Input Variable: v1g, Output Variable: v12
to rupture is approximated by the elastic–perfectly plastic ii. Input Variable: v1g, Output Variable: v23
model shown in Fig. P1.5. iii. Input Variable: v1g, Output Variable: v3g
1.5.a Calculate the energy density needed to break the spec- iv. Input Variable: v1g, Output Variable: iL
imen in US customary and SI units. and check its units in terms of the power variables, current,
1.5.b How much energy was needed to break the specimen? voltage and time.
42 1  Introduction to System Dynamics

1 R 2 1 L 2

iR iL
V iL L iC C V i source
iR R iC C
i source

g g

Fig. P1.8   Electric circuit consisting of a voltage source, resistor, Fig. P1.9   Electric circuit consisting of a battery, resistor, inductor, and
inductor, and capacitor in series, annotated with voltage nodes and the capacitor in series, annotated with voltage nodes and the positive direc-
positive directions of current through the elements tions of current through the elements

Energetic Equations: Energetic Equations:

Continuity (Conservation of Charge), Node Eqs: Continuity (Conservation of Charge), Node Eqs:

Node 1: isource = iR Node 1: isource = iR


Node 2: iR = iL Node 2: iR = iL + iC
Node 3: iL = iC
Compatibility of Voltage Drops, Path Eqs:
Compatibility of Voltage Drops, Path Eq:
v1g = v12 + v2 g v2 g = v2 g
v1g = v12 + v23 + v3g
Element Eqs:
Element Eqs:
Resistor: v12 = RiR
Resistor: v12 = RiR
diL
diL Inductor: v2 g = L
Inductor: v23 = L dt
dt
dv2 g
dv3 g Capacitor: iC = C
Capacitor: iC = C dt
dt

Energy Eqs: Energy Eqs:


System: E sys = E L + EC
System: E sys = E L + EC
1
Inductor: E L = LiL2 1 2
2 Inductor: E L = LiL
2
1
Capacitor: EC = Cv32g 1
2 Capacitor: EC = Cv22g
2
Problem 1.8 An electric circuit consisting of a voltage Problem 1.9 An electric circuit consisting of a voltage
source, a resistor, an inductor, and a capacitor, annotated source, a resistor, an inductor, and a capacitor is shown in
with nodes of distinct values of voltage and ­arrows indicat- the schematic Fig. P1.9. The energetic equations of this cir-
ing the positive direction of current through each element, cuit are listed. Use elimination by substitution to derive the
is shown in Fig. P1.8. The energetic equations of this circuit system equation for the input and output variable indicated:
are listed. Use elimination by substitution to derive the sys- i. Input Variable: v1g, Output Variable: v12
tem equation for the input and output variable indicated: ii. Input Variable: v1g, Output Variable: v2g
i. Input Variable: v1g, Output Variable: iR iii. Input Variable: v1g, Output Variable: iL
ii. Input Variable: v1g, Output Variable: v2g iv. Input Variable: v1g, Output Variable: iR
iii. Input Variable: v1g, Output Variable: iL v. Input Variable: v1g, Output Variable: iC
iv. Input Variable: v1g, Output Variable: v12 and check its units in terms of the power variables, current,
v. Input Variable: v1g, Output Variable: iC voltage and time.
and check its units in terms of the power variables and time.
Problems 43

Fig. P1.10  a Force F(t) acts


a b 1
on mass M which slides on a x,v
lubricating film with damping b
against spring K which is con-
nected to ground. b The linear
graph of the system F(t) K
M g F(t) b M K
g
1

g Lubricating fluid
Damping b g

Energetic Equations:
C
Continuity (Conservation of Charge), Node Eqs:
p(t) R1 I
3
Node 1: isource = iL Pump
1 2
Node 2 : iL = iR + iC
vent to
atmosphere
R2
patm
Compatibility of Voltage Drops, Path Eqs:
Fluid p
Reservoir atm

v1g = v12 + v2 g v2 g = v2 g
Fig. P1.11a   Fluid system modeled as a pressure source, two fluid resis-
tances, a fluid “inertance,” and a fluid capacitance
Element Eqs:

Resistor: v2 g = RiR Energetic Equations:


diL
Inductor: v12 = L Continuity (Force Equilibrium) Node Eqs:
dt
dv2 g
Capacitor: iC = C F (t ) = Fb + FM + FK
dt
Energy Eqs: Compatibility of Velocity, Path Eq: v1g = v1g
dv1g dFK
System: E sys = E L + EC Element Eqs: Fb = bv1g FM = M = Kv1g
dt dt
1
Inductor: E L = LiL2 Energy Eqs:
2
1
Capacitor: EC = Cv22g System: E sys = E M + E K
2
1
Problem 1.10 A translational mechanical system consisting Mass: E M = Mv12g
2
of a mass M sliding on a lubricating fluid film with damping
b, and a spring K attached between the mass and ground is F2
Spring: E K = K
shown in Fig. P1.10a. The linear graph of this energetic sys- 2K
tem, analogous to an electric circuit diagram is Fig. P1.10b.
The energetic equations are listed. Use elimination by substi- Problem 1.11 A schematic of a hydraulic system is shown
tution to derive the system equation for the input and output in Fig. P1.11a. The pump, modeled as a pressure source p( t),
variable indicated: discharges fluid into a hydraulic circuit consisting of two
i. Input Variable: F(t), Output Variable: Fb fluid resistances, R1 and R2, a fluid inertance I, which stores
ii. Input Variable: F(t), Output Variable: FM kinetic energy, and a fluid accumulator with capacitance C,
iii. Input Variable: F(t), Output Variable: FK which stores energy by compressing a spring or nitrogen-
iv. Input Variable: F(t), Output Variable: v1g filled bladder. Figure P1.11b is the linear graph of the sys-
and check the its units in terms of the power variables, force, tem, analogous to an electric circuit. The energetic equations
velocity and time.
44 1  Introduction to System Dynamics

R1 I Element Eqs:
1 2 3
Fluid Resistance R1: p12 = R1QR1
Fluid Resistance R2 : p3 g = R2 QR2

p(t) C R2 dQI
Fluid Inertance I : p23 = I
dt
dp3g
Fluid Capacitance C: QC = C
dt
patm patm
Energy Eqs:
Fig. P1.11b   Linear graph of the hydraulic system
System : E sys = E I + EC
are listed. Use elimination by substitution to derive the sys- 1 2
tem equation for the input and output variable indicated: Inertance: E I = IQ I
2
i. Input Variable: p( t), Output Variable: p12
1
ii. Input Variable: p( t), Output Variable: p23 Capacitance: EC = C p32g
2
iii. Input Variable: p( t), Output Variable: p3g
iv. Input Variable: p( t), Output Variable: QI
v. Input Variable: p( t), Output Variable: QC
vi. Input Variable: p( t), Output Variable: QR
2
References and Suggested Reading
and check its units in terms of the power variables, pressure,
volume flow rate Q, and time. Crandall SH et al (1982) Dynamics of mechanical and electromechani-
cal systems. Krieger, ­Malabar
Karnopp DC, Margolis DL, Rosenberg RC (2012) System dynamics:
Energetic Equations: modeling, simulation, and control of mechatronic systems, 5th edn.
Wiley, New York
Continuity (Conservation of Volume of Incompressible Kulakowski BT, Gardner J, Shearer L (1997) Dynamic modeling and
control of engineering systems, 3rd edn. Cambridge, Cambridge
Fluid), Node Eqs: Ogata K (2003) System dynamics, 4th edn. Prentice Hall, Englewood
Cliffs
Node 1: Q = QR1 Rowell D, Wormley DN (1997) System dynamics: An introduction.
Prentice Hall, Upper Saddle River
Node 2: QR1 = QI Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading
Node 3: QI = QR2 + QC

Compatibility of Pressure Drops, Path Equations:

p1g = p12 + p23 + p3g p3 g = p3 g


Differential Equations, Input Functions,
Complex Exponentials, and Transfer 2
Functions

Abstract 
Differential system equations describe the dynamic relationship between an input driving
the system, and one of the power variables within the energetic system. We simplify, or
linearize, the individual energetic element equations, in order to derive a system equation
which is an ordinary differential equation with constant coefficients, a form which we can
solve for the output or response function. The method of undetermined coefficients super-
poses or sums the response of a system into the natural or homogeneous response of the sys-
tem to a disturbance to its energetic equilibrium, and the steady-state or particular response
to each input driving the system. Systems with two or more independent energy storage
elements yield differential system equations which may describe oscillations or vibrations.
Complex numbers, complex exponentials, and Euler’s equations simplify the solution and
interpretation of the response of oscillatory systems. The Laplace transformation transforms
differential equations into algebraic equations, which can be expressed as multiplicative
dynamic operators called transfer functions. The chapter’s appendix introduces Mathcad
and MATLAB to plot the solutions or response functions.

2.1 Introduction differential system equation for an input is a prediction of the


response of the actual system to that input, Fig. 2.1.
This chapter presents most of the advanced mathematics used The most accurate description of dynamic systems re-
in the remainder of the text. The topics and methods intro- quires non-linear differential equations. However, most non-
duced here will be developed further, when they are applied linear differential equations cannot be solved manually using
in later chapters. Most engineering students do not under- classical techniques. We can solve non-linear differential
stand advanced mathematics, until applying it to physically system equations using the appropriate numerical method,
meaningful problems. This is particularly true for differen- such as Runge–Kutta method, if the solution exists, and we
tial equations. It is certainly true that mathematical ability will learn how to use the computational tools, Mathcad and
varies between individuals, and that understanding advanced MATLAB, to solve linear and non-linear differential equa-
mathematics takes some students longer than others. Howev- tions.
er, with persistence, it is possible for all engineering students However, engineers must also understand the analytical
who have passed their courses in calculus and differential solutions of ordinary differential equations, in order to under-
equations to gain an understanding of this mathematics. stand the physics of dynamic systems, and to effectively use
Any quantity which flows requires time to move or ac- design techniques developed from the analytical solutions.
cumulate. Energy flows across system boundaries and be- We shall apply an engineering perspective to differ-
tween elements within an energetic or dynamic system. The ential equations. The differential equations we will derive
time dependency of energy storage leads to the power flows are input–output relationships. These differential “system”
in dynamic systems described by differential equations. The equations will be viewed as operating on the input to the
interaction of the individual energetic elements which com- energetic system, an input that is under our control, to yield
prise a system with the power source that provides the input is an output variable, any power variable in the system, whose
described by differential system equations. The solution of a time response we wish to predict (i.e., calculate).

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_2, © Springer Science+Business Media New York 2014 45
46 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

even when that knowledge is far from complete, to make


Input Variable Dynamic Output Variable
rational decisions and recommendations.
Forcing Function System General Solution
We will limit ourselves to linear differential equations
and view them as dynamic “operators” which operate on an
Fig. 2.1   Differential system equations are input–output relationships
used to predict the response of energetic systems to different inputs
input to yield an output. Correspondingly, we will view the
dynamic system as a physical device which responds to or
operates on its input power. The responses of interest to us
Unit Step Response are time histories of the system’s power variables. An un-
F(t) Unit Step Input v(t)
1
derstanding of the mathematics in this chapter changes a
student’s perspective from “How do I solve this differential
0 time Differential 0 time equation?” to “What is the possible behavior in this system?”
System
Input Variable, F(t) Equation Output Variable, v(t)

Fig. 2.2   A differential system equation can be viewed as a mathemati- 2.2  Input Functions
cal operator. The differential system equation operates on an input func-
tion to yield the corresponding response function
The inputs to differential system equations are mathemati-
cal functions which approximate the time history of a power
We shall see that, if we restrict the models of the energy variable of an actual power source driving a physical sys-
storage and dissipation mechanisms, such that they (1) have tem. The most common input functions are “singularity”
constant properties, and (2) a linear relationship between functions and sinusoids. A “singularity” is a finite quan-
two power variables, or between a power variable and the tity divided by zero. The useful singularity input function
time derivative of a power variable, the resulting differential is the Heaviside unit step function. The singularity of the
system equation is linear. Linear differential system equa- Heaviside step function is its time derivative at the instant
tions allow the use of “superposition,” in which a solution is the function makes it “step change” from zero to one. The
constructed by breaking down an arbitrary input to a sum of step change is modeled as occurring instantaneously. Hence,
simpler inputs, solving for each input individually, and then there is a finite change in zero time. The two other singular-
summing the solutions. This is the basis of the Method of ity functions commonly used as input functions are the unit
Undetermined Coefficients. impulse function and the unit ramp function, which are the
We will extend our use of superposition to construct ar- time derivative and the time interval of the Heaviside step
bitrary inputs to our dynamic systems from scaled, time- function, respectively. Sinusoids are used to represent peri-
shifted, and superimposed Heaviside unit step functions. odic inputs. The Fourier series allows any periodic function
The response of a dynamic system to an arbitrary input is to be approximated by a sum of scaled (or “weighted”) sine
constructed by performing the same scaling, time shifting, and cosine terms of integer multiples of a base or “funda-
and superimposing of the response of the system to a single mental” frequency. The use of a single sinusoid as an input is
unit step input, as was performed on the input unit step func- presented in this chapter. The Fourier series and “frequency
tions. The use of “unit step response functions” will shift response,” which is the steady-state response of a system to
our focus from how to solve linear differential equations to a sinusoidal input, are discussed in Chap. 10.
how to use the solutions of these equations in engineering
applications, Fig. 2.2.
The carrot for the engineering student, who believes that 2.2.1  Power Sources
he or she never understood differential equations and can-
not bear to look at another differential equation, is straight- An input to an energetic system is one of the two power vari-
forward. Step response functions allow us to identify the ables of a power source connected to the system. We model
solution of differential system equation, rather than solve an ideal power source as a device which can maintain a value
the equation. There is a less obvious benefit also. Textbook of one of its two conjugate power variables by supplying as
problems provide sufficient information to apply rigorous much power as is drawn by the system. We can only control
mathematical methods. The nature of the design process is one of the two power variables of a power source. The power
the opposite. Although practicing engineers must anticipate drawn from the source by the system determines the value of
and accommodate the response of an energetic system to a the second power variable.
power input, they rarely have the luxury of formulating a There are no ideal power sources, but there are many that
differential equation. Typically, very little is “known,” since can be reasonably approximated as ideal. In general, the less
the design is still in flux. An engineer must use the available power drawn from a power source relative to its capacity,
analytical tools with what he or she knows about a system, the more ideally the power source behaves. For example, a
2.2  Input Functions 47

chemical battery, such as an automotive lead-acid storage us(t)


battery, will maintain a fixed voltage over a range of currents 1
for some duration. A battery will behave close to ideally, if
the current drawn by the system and the duration it is drawn
for are small, relative to the battery’s capacity. When the cur-
rent drawn is no longer relatively small, we generally need to
include an energy dissipater in the source model, so that the 0 time
output voltage decreases as the current flow increases. The
usefulness of any dynamic model is limited to an operating Fig. 2.3   The Heaviside unit step function. The step ­function “transi-
tions,” or changes its output value from zero to one, when its argument
range of the system’s variables. equals zero
An ideal linear power source must “source,” or provide
power, and “sink,” or accept power. A lead-acid storage bat-
tery provides power when it discharges and accepts power to note that the Heaviside step function is constant, except at
when it recharges. Many batteries can provide only power. the instant it transitions from zero to one. It has just one tran-
They cannot be recharged. We deal with a source that cannot sition. Consequently, the Heaviside unit step function cannot
accept power, by limiting the valid range of the results of the be “turned off.” It must be “negated” by adding or superpos-
dynamic model. The results are valid, until the time the model ing a negative step at the time we wish the sum to equal zero.
predicts the power flowing back into the source. If that duration We will address “time shifts” in Chap. 3.
of the response is too brief to reveal the aspect of the system There is neither a single definition nor standard notation
we are interested in, then we must program a “computational for the Heaviside unit step function. The Heaviside unit step
model” in order to impose conditional logic on the behavior of function is sometimes defined to have a value of 0.5 when
the source and solve it using a “numerical method.” Numerical the argument equals zero:
methods are computer-based solutions using finite differences
 us (t ) = 0 for t<0
to approximate derivatives. They are presented in Chap. 8.
us (t ) = 0.5 for t=0
us (t ) = 1 for t≥0 (2.2)
2.2.2  Heaviside Unit Step Function
The most common notation for Heaviside is H (t ). Anoth-
A typical differential equations problem reads as follows: er common notation is 1(t ), where “1” is the number one.
Mathcad uses Φ (t ) . MATLAB’s symbolic math toolbox uses
dv Heaviside( t).
6 + 3v = 8
dt We define time, t = 0. If we are applying only one step
input to a system, we will usually choose t = 0 to be the
which, if interpreted literally, makes no physical sense, be- instant the step input is applied. The initial conditions are
cause the input is not a function of time. It is common mathe- usually presented in textbook differential equation prob-
matical notation, however, because it is implied that the input lems at time, t = 0 . Since the transition of the Heaviside
is applied to the system at time, t = 0 , where time, t = 0, is step function from the value of zero to the value of one
defined as the beginning of the dynamic response. For our is instantaneous, we need to split hairs and define three
analyses to be useful, we need to be able to turn on and turn different time zeros. Time, t = 0, is the instant the step
off the input when we wish. The Heaviside unit step function, is applied. We will call the instant before the step is ap-
invented by the British engineer, Oliver Heaviside, serves plied, t = 0 −, and the instant after the step is applied, t = 0+ ,
that purpose, Fig. 2.3. Fig. 2.4.
The Heaviside unit step function has two possible values. The derivative of the Heaviside’s step function transition
It equals zero, when its argument is negative, and equals one, is a problem. Mathematically, the value of the derivative at
when its argument is positive. It is defined as: the instant of the transition is infinite, because the transition
from zero to one occurs instantaneously, that is, over zero
 us (t ) = 0 for t < 0 time. The derivative of the step function is zero for all other
us (t ) = 1 for t ≥ 0 (2.1) values of time, since the step function is constant, except at
the instant of its transition.
where “u” stands for “unit magnitude” and the subscript, “s”
stands for step. We will use this notation. The Heaviside step  dus (t )
= ∞ for t = 0 (2.3)
function’s change from zero to one is referred to as its “transi- dt
tion” or, informally, when the step “turns on.” It is important
48 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

us(t) t>0 ∞
us(t) = 1
1 δ(t)

t<0 t<0 t>0


us(t) = 0 δ(t) = 0 δ(t) = 0
time time

t = 0- t = 0+ t = 0- t = 0+
t=0 t=0
Fig. 2.4   The Heaviside unit step function, us (t ) , Eq. 2.1. The time, Fig. 2.5   The unit impulse function δ (t ), Eq. 2.5
t = 0− , is the instant immediately before the transition. Time, t = 0, is
the instant the step transitions. Time, t = 0+, is the instant immediately
after the step has transitioned
tial conditions needed to solve differential equations, after
we introduce “state” variables, and investigate the math-
and ematical implications they have as the energy storage vari-
ables of a system.
 dus (t )
= 0 for t ≠ 0 (2.4)
dt
2.2.3  Unit Impulse
If a Heaviside step function is differentiated as an input term
of a system, which value should we use, zero or infinity? The unit impulse function, Eq. 2.5, Fig. 2.5, is the derivative
When the Heaviside step input is applied at time, t = 0, we of the Heaviside unit step function. The amplitude of the unit
restrict the solution of the equation to time, t > 0, since the impulse function is infinite, but its duration is infinitesimal
input must precede the response to respect cause and ef-
fect, and use the value dus (t ) / dt = 0 . If the system is de- 
δ (t ) = ∞ for t = 0
energized before the step input is applied, and the input term (2.5)
of the system equation is differentiated, what drives the re- δ (t ) = 0 for t ≠ 0
sponse of the system? Although the derivative of the step
input is zero, the input to the system is not zero. The input The unit impulse function is the derivative of the Heaviside
to the system is a step. It is the input to the system, which step function. Inversely, the integral of the unit impulse is
determines the initial values of all the power variables in the the unit step function. Hence, the area under the unit im-
system and their derivatives. The name, “system equation,” pulse equals one, which led to the name “unit” impulse,
is somewhat misleading. A more descriptive but awkward when the impulse has infinite magnitude. A second com-
name would be “the differential equation for a power vari- mon name for the unit impulse is the Dirac delta function,
able in an element in the system.” There are two power vari- δ( t). Paul Dirac was a British physicist influential in the
ables in every element of a dynamic system. We can formu- development of quantum mechanics. We will use the com-
late a differential equation to describe the response of all but bination of the name unit impulse and the symbol, δ( t),
one of these variables to the input applied to the system, with as a reminder that the magnitude is not unity. Needless to
the input variable itself being the only exception. The input say, it is impossible to apply a unit impulse with infinite
term (or terms) applies to the differential system equation of amplitude, but no duration, to a physical system. The unit
a power variable of one energetic element of the system. The impulse’s use is purely mathematical. It represents the de-
differential system equation for the other power variable of rivative of the Heaviside step function. The energy which
the same energetic element will have the same input, but the drives the response is delivered to the system by the step
input term(s) will be different; they will have different coef- function, not by impulse.
ficients and differentiation.
We consider the Heaviside step function further in
Chap. 3, when we discuss modeling input steps, and using 2.2.4  Unit Ramp
superposition (summation) of step inputs to form pulse in-
puts. We will also establish how to determine the initial The integral of the Heaviside unit step is the unit ramp func-
values and derivatives of the power variables, i.e., the ini- tion, which has a slope or “ramp rate” of one. The unit ramp
2.3 Linearity 49

1.0
ur(t)
1 sin(ωt) us(t)
0.5

time
0 1 time
-0.5
Fig. 2.6   The unit ramp function ur (t ) = t us (t ), Eq. 2.6

-1.0
function is the product of the independent variable time, t,
Fig. 2.7   The product of a sine and the Heaviside unit step function
and the Heaviside step function, Eq. 2.5, Fig. 2.6.
f (t ) = sin (w t )us (t ), Eq. 2.7


ur (t ) = tus (t ) (2.6) closely the power variable of a source corresponds to an
ideal step depends on the physical device, and the amount
of power drawn from it, relative to its maximum limit and
2.2.5 Sinusoids the characteristic times of both the power source and the
system. When the functional form of the step input from an
The other important type of input is a sinusoid. Sinusoidal actual source deviates from the ideal, square-cornered step
inputs occur naturally and are created by rotating machinery described by a Heaviside unit step function, to the extent
and vibrations or oscillations within the equipment or physi- the input cannot be reasonably modeled by the ideal step,
cal system, which forms the environment around the system then the engineer modeling the system has two choices. The
of interest. Sinusoids are also the basis of the Fourier transfor- system boundary can be expanded to include the power sup-
mation, which allows any periodic input to be approximated ply as a dynamic element within the model, and the system
by a summation of sinusoids to the accuracy desired. A very model revised to include the previously unmodeled dynamic
important special case is the steady-state response of systems aspects of the power source. Alternatively, if there is not
to a sinusoidal input, which is applicable to both vibration enough information to create a model of the power source
analysis and electrical systems powered by alternating cur- or to revise the system model, then the model of the input
rent. Mechanical vibrations are introduced in Chap. 4. We must be changed. If the time history of the input variable has
shall investigate the Fourier transform and the steady-state been measured, then a “step response function” can be used
response of systems to sinusoidal inputs in Chap. 10. In this to model the input of the system. We will defer use of step
chapter, we will consider the transient period up to steady- response functions as inputs until the introduction of “trans-
state. fer functions” in Sect. 2.10.
In order to “turn on” a sinusoid at time, t = 0, it is multi-
plied by the Heaviside unit step function per Eq. 2.7:
2.3 Linearity

f (t ) = sin (w t )us (t ) (2.7) Many engineering calculations rely on superposition. The
premise of superposition is that the response of a system to
where ω is the angular frequency in radians per second. each piece of a combined input can be calculated separately,
Multiplication by the Heaviside step function is not nec- and, then, the responses summed to calculate the response to
essary, if the problem is a textbook problem. It is necessary, the entire input. In other words, superposition is the decom-
even for a textbook problem, when the sinusoid is applied at position of an input into component parts, calculation of the
a time other than t = 0, Fig.  2.7. response of the system to each component of the input, and
then summing the responses to determine the responses of
the system to the input as a whole. The easy part is to create
2.2.6  Step Responses as Input Functions an input by summing different input functions. The hard part
is ensuring that the sum of the responses calculated for each
Modeling a power source as an ideal source described by a component of the input is reasonably close to the response of
Heaviside unit step function is a mathematical convenience the system to the entire input. Superposition (or summation)
which is always in error to some degree. All power sources of the resulting step responses only has validity, if we have
are physical devices with multiple energetic attributes. How restricted our model to linear elements.
50 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Fig. 2.8  a Elastic–perfectly


plastic material model. b Forces, a b
F1 and F2, superposed on a ma- σy
chine element loaded in tension ∆L
L
E F=F1 +F2 F=F1 +F2
Area A
ε E, σ y

30
iR 1 R 2 iR v12
VDC 20

Fig. 2.9   The schematic symbol of electrical resistance showing nodes


10 100 x R
of distinct values of voltage at either end of the element
-0.4 -0.3 -0.2 -0.1 0.1 0.2 0.3 0.4
-10 i, amps

Then we can use “superposition” to calculate the response -20

of the system to an input comprised of step functions. -30


Although superposition is a fundamental tool in engineer-
ing analysis, it is often misused. Superposition works only if Fig. 2.10   Linear electrical resistance. The slope of the line is the
phenomenon is reasonably approximated by a linear model resistance, R, when the units are volts and amperes
over the range of the calculation. Consider the elastic–
perfectly plastic material model, Fig. 2.8a, which we use to
describe the mechanical response of many metals and calcu- single straight line which passes through the origin of the
lation of the stress in a rod due to two axial forces, F1, applied plot of the power variables, v12 and iR. Conversely, the elas-
at time, t = ta , and F2, applied at time, t = tb, Fig. 2.8b. We tic–perfectly plastic model is comprised of two straight lines.
can use superposition to calculate the axial stress in the rod Consequently, it cannot be a linear model. It is a non-linear
by calculating the stress due to each force separately and model. Linear models must have a single straight line which
summing the stresses, thusly, passes through the origin, Fig. 2.10.
In addition to restricting our models of energetic systems
F1 F2 to linear models, we must also use linear mathematical oper-
σ1 = and σ2 = → σ 1 + σ 2 = σ axial ators, in order to use superposition. We can establish whether
A A
or not a mathematical operation is linear with two simple
if and only if σ axial ≤ σ y tests:
1. Can the operation by distributed onto the individual terms
since the maximum stress possible in the material model is in a sum?
σy. In other words, we must stay within the linear region of 2. Does doubling the input to the operation double the re-
the elastic–perfectly plastic material model, if we wish to use sult?
superposition. The result σ axial > σ y is nonsense when the Since we intend to derive a differential system equation, we
elastic-perfectly plastic material model is used, although it should establish the conditions needed for differentiation
is commonly observed in practice since many real materials with respect to time as a linear operation. Can the operation,
strain harden. d , be distributed onto the terms of a sum, such as x + x ?
dt a b
We will use the terms “element model” or “constitutive
model” for the equations which describe the relationship be-  d ( xa + xb ) dxa dxb
= + (2.8)
tween two variables in an element of an energetic system. dt dt dt
For example, the familiar equation of electrical resistance,
v12 = RiR , is a linear element equation. Voltage v12 is the dif- Yes. Does doubling the input to the differential operator
ference or drop in the voltage across the resistance in the double the result?
direction of the current, iR, flowing through the resistance.
The resistance, R, is a property of the element, Fig. 2.9.  d (2 x) dx
=2 (2.9)
The linear element model for an electrical resistor is a dt dt
significantly simpler model than the elastic–perfectly plastic
material model. The linear electrical resistance model is a Yes. Now try to think of circumstances under which these
tests fail. Specifically, what restrictions, if any, must be
2.5  Method of Undetermined Coefficients 51

placed on x? The input x cannot be raised to a power. For ex- 5τ


ample, define x ≡ xa2. Differentiating 2 x ≡ 2 xa2 with respect Transient Steady-State

to time yields the following:

=
( )
2
d (2 x) d 2 xa dx 2 dx dx
=2 a =4 a ≠2 a x(t)
dt dt dt dt dt

We will ensure that our dynamic models of energetic sys-


tems are linear, by using only linear equations to describe
the individual elements within the system, so that we can use
time, seconds
superposition. We must always remember to interpret our
results to evaluate the magnitude of the inaccuracy we in- Fig. 2.11  An oscillatory step response showing the transient period
troduce by the linear approximation of a non-linear physical and the beginning of steady-state at time equal to five time constants
relationship. We will develop models of mechanical system
elements in Chap. 4. For the time being, we will work with
models as presented. which acts on the system. Eventually, the transient response
decays to zero, and only the forced response, described by
the particular solution, remains. To reiterate, the general so-
2.4 Superposition lution of the system equation represents the actual, observed
response of a dynamic system to the inputs acting on it.
The general solution of linear ordinary differential equations The response of a system is the superposition of the transient
with constant coefficients is the superposition or sum of the response, created by disturbing the energetic equilibrium
solution to the homogeneous equation, created by setting the between the system and the power supply, and steady-state
output variable side of the differential equation to zero, and of response to each input.
the particular solution(s) for each forcing function. The forc-
ing functions are the inputs to the system. When an energetic
system has reached steady-state, there is equilibrium between 2.5  Method of Undetermined Coefficients
the energy storages in the system, Fig. 2.11. The solution of
the homogeneous equation represents the natural response of The Method of Undetermined Coefficients allows us to solve
a system to a disturbance. If the system is then disturbed by linear ordinary differential equations with constant coeffi-
an input of energy, a new equilibrium must be reached. cients, by remembering a generic solution for the unforced
The homogenous response of a dynamic system describes natural response, as well as generic solutions for a small num-
how the system reaches its new energy equilibrium. An ide- ber of physically common, forcing functions. The generic
ally linear system’s homogenous response decays exponen- solutions have adjustable constants (the undetermined coef-
tially and never equals zero. In reality, the homogenous re- ficients) to scale the amplitude, duration, frequency, etc., to
sponse of systems of practical interest eventually decays to fit the specific system equation, input, and initial conditions.
zero. We refer to the homogenous response as “transient” The general solution is constructed by the superposition, or
because of its finite duration. Since an ideally linear homog- summation, of the natural and forced responses.
enous response never equals zero, we must define the end Our purpose in reviewing the method of undetermined
of the transient period. Conventionally, we set time equal to coefficients is to provide a rigorous justification for shortcuts
either four or five times the system’s “time constant,” τ, as that we will develop. Our preferred solution method will be
the duration of the transient period. The time constant scales to extract the information contained in the differential equa-
the rate exponential decay of the transient response’s fac- tion, in order to scale the known forms of the solutions to
tors, and thus controls the rate at which a system’s response match a specific system equation, and set of inputs. In short,
reaches steady-state. we will avoid as much of the mathematical formality typi-
The transient response of a system is the sum of the ho- cally found in a course in differential equations, as possible.
mogenous response and the particular solution for each input In that vein, consider the nth differential equation below:


dnx d n −1 x d n−2 x dx
f1 (t ) + f 2 (t ) +  + f n (t ) = a 0 n + a1 n −1 + a 2 n − 2 +  + a n −1 + a n x (2.10)
dt dt dt dt
52 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Table 2.1   Common and Common form Standard form


standard forms of ordinary
differential equations with First-order
constant coefficients dv 1 dv a1
F (t ) = a 0 + a1v F (t ) = + v
dt a0 dt a 0

1 a 0 dv
F (t ) = +v
a1 a dt
1
τ
Second-order
d 2v dv 1 d 2 v a1 dv a 2
F (t ) = a 0 + a1 + a 2 v F (t ) = 2 + + v
dt 2 dt a0 dt a 0 dt a 0
nth-order
d nv d n−1v 1 d n v a1 d n−1v an
F (t ) = a 0 n
+ a1 n−1 +  + a n v F (t ) = n + ++ v
dt dt a0 dt a 0 dt n−1 a0

Although this nth order differential equation is in a general Step 1.  Separate the Input Functions and the Output Vari-
form, it is also more difficult to read than a differential equa- ables and Express the Differential Equation in Standard
tion of a specific order. As we will see later, a fourth-order Form.
differential system equation is a rarity of questionable accu- The input functions, also known as forcing, exciting, or driv-
racy. There is never a practical need for a higher-order equa- ing functions, describe what we plan to do, or did, to the sys-
tion. Consequently, we will illustrate the method of undeter- tem. Time is the independent variable in a dynamic system.
mined coefficients using a fourth-order differential equation, We must be able to describe our actions on the system in the
so as to avoid the “n” notation. language of mathematics as functions of time. We will iden-

d4x d3x d2x dx (2.11)
f1 (t ) + f 2 (t ) +  + f n (t ) = a 0 4
+ a1 3 + a 2 2 + a 2 + a4 x
dt dt dt dt

Unfortunately, there are two different conventions for num- tify the input functions as f (t ). The output variable, x, is also
bering the coefficients, each the reverse of the other. One a function of time. It should be written as x(t ), rather than as
method assigns the subscript, zero, to the coefficient of the simply x, but we normally don’t bother.


d4x d3x d2x dx
f1 (t ) + f 2 (t ) +  + f n (t ) = a 0 4 + a1 3 + a 2 2 + a 3 + a4 x (2.12)
   dt dt 
dt dt
 
Input Terms
Output Terms

highest-order term. The other method assigns the subscript, On the left of Eq. 2.12 is a summation of input functions.
zero, to the zero-order term. These conflicting conventions This summation is the “superposition” of the input func-
are a problem whenever there is a formula expressed in tions. The input summations occur for both physical and
terms of the subscripts of the coefficients. Never use a for- mathematical reasons. Physically, it is generally difficult to
mula which involves coefficient subscripts, unless you are isolate a system from the surrounding environment. If the
certain which numbering convention it is based on. action of the environment on the system is not negligible,
We will not limit the number of terms on the input side then there will be at least two physical inputs: what we do
of the differential equation. It is common and useful to to the system, and what the environment does to it. Multiple
sum (or superpose) any number of inputs. Therefore, we input terms are created mathematically by different orders
will retain the indefinite n for the number of input func- of differentiation of the input function. Also, a single physi-
tions or terms. cal input may require the sum of two or more functions to
The Method of Undetermined Coefficients is as fol- describe it. For example, a sinusoidal input with a non-zero
lows: average value is the sum of a sinusoid and a Heaviside step
2.5  Method of Undetermined Coefficients 53

function. A second example is the “construction” of finite c. Is the equation linear? This takes more thought. The ques-
duration pulses of various shapes, by summing scaled and tion of linearity refers to the dependent, or output, vari-
time shifted Heaviside step functions. If the differential able, x. The input or forcing functions can be non-linear.
equation is linear and has constant coefficients, as described In fact, an important forcing function is sin(w t ) . The
below, then we can solve for the response of each input func- input side can have non-linear terms. It is the output side,
tion separately, and sum (or superimpose) each of the results the side that describes the system, which must be linear.
to determine the overall response of the system. The differential operator is a linear operator, regardless of
We will use two different standard forms for differential the order of the differential because it distributes onto a sum,
system equations, one form for first-order system equations, Eq. 2.8., and doubling its input doubles its output, Eq. 2.9.
and the other for second-order and higher-system equations. For the output side to be linear, it cannot contain trig func-
The standard form for second-order and higher-system equa- tions, powers of the dependent (output) variable, products of
tions is as follows. After separating the input and output terms, the dependent variable and its derivatives, or products of the
clear the highest-order output term of its coefficient, Eq. 2.13. derivative of the dependent variable and another derivative.
This is the same standard form used with polynomials. Eq. 2.14 is non-linear because of the square on the output

1 1 1 d 4 x a1 d 3 x a 2 d 2 x a 3 dx a 4 (2.13)
f1 (t ) + f 2 (t ) +  + f n (t ) = 4 + + + + x
a0 a0 a0 dt a 0 dt 3 a 0 dt 2 a 0 dt a 0

The form we will use for first-order differential equations is variable, v.


called time constant form. As we will develop in depth, the
time constant of a system, τ (Greek for t), is the time scale  1 dF d 2 v b dv 2 K
= 2 + + v (2.14)
of its natural or unforced response. The time constant can be M dt dt M dt M
identified by units analysis, after the output variable term Non-
linear
of the first-order differential equation is cleared of its coef-
ficient. The time constant is also known as the “relaxation Most non-linear differential equations cannot be solved ana-
time,” a term with origins in elasticity, but now extended into lytically, because you cannot use superposition to break the
other applications. problem down into bite-sized pieces. The solution of non-
linear differential equations is an all or nothing operation. Do
Step 2.  Check the Type of the Differential Equation. not attempt to use linear methods on non-linear equations.
The system equation produced by reducing a lumped param- Go back to the beginning and linearize the model, by using
eter model of an energetic system will be a linear ordinary linear element or constitutive equations.
differential equation, if the models of the individual ener-
getic elements are all themselves linear models. In practice, Step 3.  Check the Units of Each Term in the Equation.
this means that equations which describe energy dissipation This is an important check. You are checking to see if you
in mechanical and fluid systems (i.e., friction), and some of are adding apples and apples. All terms in a summation must
the equations which describe elastic strain energy storage have the same units, or they cannot be summed. If the units
must be “linearized,” i.e., approximated by linear equations. don’t check, then you are working with nonsense. Check
We will examine the process more closely in Chap. 3, but the now, before you invest any more time in the solution. Con-
need to eliminate non-linear element equations is absolute. sider a second-order equation:
We cannot solve non-linear differential system equations
analytically. We must resort to computational (numerical) d 2v dv
F = a0 + a1 + a 2 v
methods, which we will develop in Chap. 8. Check the sys- dt 2
dt
tem equation before proceeding.
a. Is it an ordinary differential equation? If there are partial where F is the input force in Newtons, and v is the output
derivatives, then it is a partial differential equation, not an velocity in m/sec.
ordinary differential equation. What are the units of each term in this equation? New-
b. Are the coefficients a 0 , a1 , a 2 , … a n of the output side tons, since the right side of the equation is force in newtons.
constants, or are they functions of time? If they are not If this were a system equation, the coefficients, a0, a1, and a2
constant, stop now! You can’t solve a differential equa- would consist of the parameters of the elemental (constitu-
tion with non-constant coefficients, using the Method of tive) equations. We will use SI units in all of our calculations.
Undetermined Coefficients. Hence, the coefficients of mechanical system equations will
54 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

be terms with units of mass in kilograms, force in newtons, A fact of engineering mathematics you must know is the
displacement in meters, and time in seconds. solution of an ordinary differential equation with constant
One should check units after expressing the equation in coefficients homogeneous equation, Eq. 2.16. This is the one
standard form, since a unit check will reveal an error in that and only solution of the homogeneous equation:
process. Divide both sides by the coefficient of the highest-
order derivative of the output variable.  xH (t ) = A1e s1 t + A2 e s2 t +  + An e sn t (2.16)

d 2v dv 1 d 2 v a1 dv a 2 where n is the order of the differential equation. The coeffi-


F = a0 + a 1 + a 2 v → F = + + v
dt 2 dt a0 dt 2 a 0 dt a 0 cients An are the undetermined coefficients of the method of
undetermined coefficients. They are determined in the final
What are the units of the derivative terms? Recall the defini- step of the method. It will be easy to remember Ae st because
tion of a derivative from Calculus I. we will use this exponential function frequently.
The complex variable, s = σ + jw , of the exponen-
dv ∆v
= lim tial Ae st is a “characteristic” value of the system. In other
dt ∆t → 0 ∆t words, it is a solution of the system’s characteristic equation.
The characteristic equation is created by evaluating the ho-
v
Therefore, dv has the units,   , where the square brack- mogenous equation with the trial solution, x = Ae st . Recall
dt t  how to differentiate an exponential.
ets are read as “units of.” The derivative operator, d, means
“infinitesimal change of.” The derivative operator has no
 deu du
units, and does not affect the units of the term it operates = eu (2.17)
2 dx dx
on. Consider the second derivative, d 2v , and carefully note
dt
where the superscripts are. The superscript in the numerator A source of confusion is how to differentiate the Laplace
squares the derivative operator. The superscript in the de- variable, s, with respect to time. The Laplace variable s is
2
not a function of time. Consequently, it is treated as if it were
nominator squares the variable time, t. Consequently, d 2v
v dt a constant, and its time derivative is zero.
has the units,  2 .
 t 
dAe st de st dst  dt ds 
=A = Ae st = Ae st  s +t
Step 4.  Form and Solve the Homogeneous Equation. dt dt dt  dt 1 dt 0 
Set the input side to zero (i.e., no forcing functions) to create
the homogeneous equation. dAe st
= Ae st s
dt
 d 4 x a1 d 3 x a 2 d 2 x a 3 dx a 4
0= + + + + x (2.15)
dt 4 a 0 dt 3 a 0 dt 2 a 0 dt a 0 Although the mnemonic from Calculus places du/dx after the
exponential, we will place the derivative in front, to make it
This is called the homogenous equation, because “homoge- more prominent and, therefore, less likely to be lost during
neous” means “self-­generating,” and the trivial solution of transcription:
the homogenous equation, x = 0, equals the input side of
zero.  deu du u dAe st
= e → = sAe st (2.18)
The homogeneous equation represents the natural (un- dx dx dt
forced) response of the system, because the forcing function
is zero. Physically, the homogeneous solution represents the Example One  Formulate and solve the homoge-
response of the system, when it is given some amount of en- neous equation of the first-order differential equation,
ergy, then “released.” The energy flows through the system 1 dv a1 . Form the homogeneous equation by
a 0 F (t ) = dt + a 0 v
and into the environment, in a way that depends on the na-
setting the input side of the equation equal to zero.
ture of the system, or, in other words, is characteristic of the
system. This happens every time the system is “disturbed,”
which is whenever we apply an input or change an input to 1 dv a1 dv a1
F (t ) = + v → 0= + v
the system. Consequently, the homogeneous response of the a0 dt a 0 dt a 0
system is always present in the initial period of a system’s
response. The duration of the homogenous response defines The assumed form of homogeneous solution is always the
the “transient” period. A system has reached “steady-state,” exponential, Ae st . Substitute the homogenous solution,
when the homogeneous response has decayed to zero. vH (t ) = Ae st , into the homogeneous equation:
2.5  Method of Undetermined Coefficients 55

dvH a1 dAe st a1  d n Ae st
0= + vH → 0 = + Ae st = s n Ae st (2.19)
dt a0 dt a0 dt n

Evaluate the derivative. as repeated first-order differentiation. Each order of differ-


entiation yields the Laplace variable, s, raised to the power
a1
0 = sAe st + Ae st equal to the order of differentiation.
a0
Example Two Formulate and solve the homogeneous
The exponential, e st , can never equal zero. Factor out the equation of the second-order differential equation,
exponential, Ae st. 1 d 2 v a1 dv a 2 .
F (t ) = 2 + + v
a0 dt a 0 dt a 0
 a1  Form the homogeneous equation by setting the forcing
0 =  s +  Ae st
 a0  (input) function to zero.

The parenthetical term is the “characteristic function.” For 1 d 2 v a dv a2 d 2 vH a1 dvH a2


F (t ) = 2 + 1 + v → 0= + + vH
this equation to be true, the characteristic function must a0 dt a0 dt a0 dt 2 a0 dt a0
equal zero, since the exponential cannot. Setting the charac-
teristic function equal to zero yields the characteristic equa- Substitute the homogenous solution, vH (t ) = Ae st, into the
tion, which is easily solved for s. homogeneous equation.

a1 a1 d 2 Ae st a1 dAe st a 2
s+ =0 → s=− 0= + + Ae st
a0 a0 dt 2 a 0 dt a0

The result is the homogeneous solution with the unknown Evaluate these derivatives.
factor, A, the “undetermined coefficient” of the method of
undetermined coefficients. a1 a2
0 = s 2 Ae st + sAe st + Ae st
a0 a0
a1
− t
vH (t ) = Ae vH (t ) = Ae
st a0
→ Factor out the exponential Ae st .

Physically, the natural response of all first-order, stable en-  a1 a2 


0 =  s 2 + s +  Ae st
ergetic systems is an exponential decay. We will see that ho-  a0 a0 
mogenous, natural, or “unforced,” solution is present, even
when we drive a system with an input or forcing function. Set the characteristic function equal to zero, to form the char-
The exponential decay results from the system reaching equi- acteristic equation.
librium to the changed conditions. If we have a system which
is initially de-energized, and apply a step input, then energy a1 a2
s2 + s+ =0
will flow into the system, eventually leading to an equilibrium a0 a0
condition, in which the amount of energy in the system stays
unchanged, until the input is again changed. Any subsequent This algebraic equation is a second-order polynomial, in
change to the input will initiate another natural response. which the Laplace variable, s, has two solutions, which are
The homogenous solution of higher-order differential found using the following quadratic equation:
equations follows the same steps as the previous first-order
system example. Higher-order derivatives of the solution −b ± b 2 − 4ac
s1 , s2 =
Ae st are evaluated by repeated differentiation, 2a
2
d 2 Ae st d  dAe st  d st dAe st  a1 
a1 a2
= = ( s Ae ) = s = s 2 Ae st − ±   −4
dt 2 dt  dt  dt dt a1 a2 a0  a0  a0
s2 + s+ = 0 → s1 , s2 =
where the Laplace variable s can be factored out of the de- a0 a0 2
rivative, since it is not a function of time. The shortcut is to
recognize that evaluating higher-order differentiation of Ae st Because the characteristic equation of the system has two
yields the factor, s, raised to the order of the differentiation, solutions, s1 and s2 , the homogenous solution is the sum of
56 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

two exponential terms of the form, Ae st. Either term solves  d 4 x a 1 d 3 x a 2 d 2 x a 3 dx a 4 


the homogeneous equation, but only the sum of both terms L {0} = L  4 + + + + x
 dt a 0 dt 3 a 0 dt 2 a 0 dt a 0 
creates the homogenous response of the system.

vH (t ) = A1e s1 t + A2 e s2 t Although the Laplace transformation is only defined for


functions of time, zero is a special case, because of the linear-
The two unknown values, A1 and A2, lead to the requirement ity of the Laplace transformation. Constants can be factored
of two initial conditions to solve second-order differential out of linear operators. Therefore, we could, if we choose,
equations. multiply a function of time, f( t), by zero, and perform the
The significant difference between the homogeneous Laplace transformation on the product:
solution of a first-order system, and those of second- and
higher-order differential equations, is that some or all of the
solutions of characteristic equation of second- and higher-
L {0 ⋅ f (t )} = 0 ⋅ L { f (t )} = 0 ⋅ F ( s ) = 0
order equations may be complex numbers. Higher-order
differential system equations describe energetic systems Distribute the Laplace transformation operator onto each
with two or more “independent” energy storage modes or term of the summation.

d4x  a 1 d 3 x   a 2 d 2 x   a 3 dx   a 4 
L {0} = L  4  +L  3 
+L  2 
+L   +L  x
 dt   a 0 dt   a 0 dt   a 0 dt   a 0 

Factor out the coefficient terms.

 d 4 x  a1 d3x a2 d 2x a3  dx  a 4
L {0} = L  4  + L  3  + L  2  + L   + L { x}
 dt  a 0  dt  a 0  dt  a 0  dt  a 0

elements. An energy storage element is independent, if it The Laplace transformations of derivatives with respect to
is possible to store energy in it but no other energy storage time are Eqs. 2.20 and 2.21.
element in the system. The energy will not stay in that one
energy storage element. It will flow into the other elements 
 df (t ) 
of the system. The energy storage element is independent, L   = sF ( s ) − f ( 0) (2.20)
if it can be energized for an instant when all other energy  dt 
storage elements are de-energized. A familiar example of
independent energy storage elements is a spring-mass sys-   d 2 f (t )  df ( 0) (2.21)
L  = s F ( s ) − sf ( 0) −
2
tem. We can visualize the spring extended or compressed  2
 dt  dt
when the mass is stationary. A spring-mass system has two
independent energy storage modes or elements, strain or
The initial condition terms, f (0) and df (0) , are neglected
potential energy and kinetic energy. When there are two dt
or more independent energy storage elements in a system, and only product of the transformed variable and the Laplace
energy can be transferred internally between energy stor- variable s raised to the power of the order of the differentia-
ages within the system, as well as transferred into or out of tion is retained.
the system from the environment. Internal energy transfers Evaluate the Laplace transformation. An unknown time-
produce oscillations. Eventually, the oscillations die out, domain variable, x( t), transforms to an unknown Laplace-
as some of the energy during each internal transfer is dis- domain variable X( s).
sipated as heat, and the system reaches equilibrium with
the input. Systems that have complex characteristic values a1 a2 a3 a4
0 = s 4 X ( s) + s3 X (s) + s 2 X (s) + sX ( s ) + X (s)
will oscillate, when subjected to a step input. a0 a0 a0 a0
A more direct method of forming the characteristic equa-
tion is to perform the Laplace transformation on the homo- Factor out the Laplace-domain variable X( s).
geneous equation. The Laplace transform will be formally
introduced and its properties investigated in Sect. 2.10.  a1 a2 a3 a 
0 =  s 4 + s3 + s 2 + s + 4  X (s)
 a0 a0 a0 a0 
2.5  Method of Undetermined Coefficients 57

The summation is the “characteristic” function. For this third and fourth order are so time-consuming that, in the bad
equation to be true, either the characteristic function equals old days of slide rules, we used iteration (guessing and refin-
zero or X( s) equals zero. The latter is the trivial solution. ing guesses) to find the roots. In the twenty-first century, we
Setting the characteristic function equal to zero forms the use software, such as Mathcad, MATLAB, and Mathemat-
characteristic equation. ica, or high-end calculators, such as a TI-89 or TI-Nspire,
to solve polynomials. The software and calculators imple-
a1 a2 a3 a4
0 = s4 + s3 + s2 + s+ ment the Jenkins–Traub algorithm to solve polynomials. The
a0 a0 a0 a0 algorithm estimates a factor of the polynomial, divides the
polynomial by that factor, and repeats the process until the
Having recognized that the orders of s in the characteristic remainder is second order which is then solved by the qua-
equation equal the order of differentiation in the correspond- dratic formula.
ing differential equation, we can omit the mathematical for-
mality and write the characteristic equation directly. This Step 5.  Solve the Forced Equation for Each Forcing Func-
shortcut method can be dignified by calling it a “mapping,” tion.
which is a systematic substitution. There is a “particular” solution, xp, of the differential equa-
tion for each individual input or forcing function. The sum,
d d2 dn or superposition, of all of the particular solutions for the forc-
→s → s2 → sn
dt dt 2 dt n ing functions is the forced response. The forced response is
the steady-state response of the system. All physical systems
Applying the shortcut mapping method: try to follow their forcing functions. Linear systems actually
do follow theirs. The steady-state response of a linear system
d 4 x a1 d 3 x a 2 d 2 x a 3 dx a 4 will have the same functional form as the forcing function.
0= + + + + x Therefore, find the particular solution for each individual
dt 4 a 0 dt 3 a 0 dt 2 a 0 dt a 0
forcing function, by guessing a general solution that has the
same form as forcing function; substitute it into the differen-
 a1 a2 a3 a  tial equation; and evaluate the constants.
0 =  s 4 + s3 + s 2 + s + 4  X (s)
 a0 a0 a0 a0  The trial solution should contain the function, and all of
its derivatives that produce a term that differs from the initial
a1 a2 a3 a4 term by more than a constant. Each term is preceded by an
0 = s4 + s3 + s2 + s+ unknown constant, hence the name, “the method of unde-
a0 a0 a0 a0
termined coefficients.” Common forcing functions include
steps, ramps, polynomials, sinusoids, and exponentials. Re-
“Characteristic equation” is a good name, because the roots peat Step 5 for each type of forcing function.
of this equation (the s’s that make it equal to zero) give the The most common forcing function is a step function of
time constant for the exponential decay and the frequency some magnitude. For the time being, we will use step func-
of the oscillation, which indeed characterize the natural re- tions which “turn on” or transition at time, t = 0. A step
sponse of the system. A less meaningful but common name function is constant after it transitions. It is common to use
for this equation is the “eigen” equation. While Euler was an a constant instead of a step function, as a forcing function in
actual historical figure, there was no Herr Dr. Prof. Eigen. textbook problems. A ramp function is a step function multi-
“Eigen” is German for “innate,” or “one’s own.” The roots of plied by time, t. We will use, for the time being, ramp func-
the eigen (characteristic) equation are called the eigenvalues. tions which transition on at time, t = 0. A parabola function
The term “eigenvalue,” a mixture of German and English, is a ramp function integrated with respect to time, t.
is commonly used. An eigenvalue is a characteristic value. Summing (or superposing) a step and a ramp, creates a
The conventional notation for eigenvalues differs from that polynomial forcing function:
of characteristic values. Lowercase lambda λ is used for an
eigenvalue, and lowercase s is used for a characteristic value.  f1 (t ) = K1 + K 2 t for t > 0 (2.22)
Due to the likelihood of confusing the Laplace variable s
with the SI abbreviation s for seconds, we will depart from Function f1 (t ) is a first-order polynomial, since t is raised to
standard SI notation and abbreviate seconds as “sec.” the first power. Adding a parabola function makes the forc-
The hardest part of solving higher-order ordinary dif- ing function look more like a polynomial in t:
ferential equations with constant coefficients is solving the
characteristic equation. There are no general solutions for  f 2 (t ) = K1 + K 2 t + K 3t 2 for t > 0 (2.23)
polynomials above fifth order, and the general solutions for
58 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

A trial solution, xp, must include each forcing function term and
and its derivatives, if there are any. Considering the forc-
ing function, f 2 (t ), the first term, K1, is constant and has no  d (C2 cos(ω t ))
= −C2ω sin(ω t ) = C2 cos(ω t ) (2.27)
derivatives. dt

x p 1 = C1 since the coefficients C1 and C2 are undetermined, they can


absorb the angular frequency, ω, and the negative sign.
The second term, K 2 t, can be differentiated and expressed An important exception is the particular solution to be
as: used for exponential forcing functions. Exponential forc-
ing functions are a special case, because exponentials are
x p2 = C2 + C3t also the solution to the homogeneous equation. If the forc-
ing function is an exponential, f (t ) = Ae st , which, by rare
The third term, K 3t 2, can also be differentiated. It and its bad luck, happens to be exactly one of the exponential terms
derivatives are in the homogenous solution, then the trial solution is modi-
fied, by multiplying it by time, t, so that instead of using
x p3 = C4 + C5t + C6 t 2 x p = Ae st , one uses this form:

Superposing (or summing) the three trial particular solutions  x p = tAe st (2.28)

x p1 + x p2 + x p3 = C1 + (C2 + C3t ) + (C 4 + C5t + C6 t 2 ) Another special case occurs, if one root of the characteristic
equation is zero, and the system is subjected to a step input. If
and collecting coefficients of like powers of time t yields the s = 0, then Ae st = Ae s 0 = Ae0 = A, since any base raised to
following: the zero power equals one. Consequently, the constant term
in the homogenous solution has the same functional form of
x p1 + x p2 + x p3 = C1 + C2 + C4 + (C3 + C5 )t + C6 t 2 one of the input terms, the step input after it has transitioned
to a constant. The resolution to this problem is the same as in
Since the constants Cn are undetermined, terms of like pow- the case of an exponential forcing function, where you may
ers of t can be combined, and the sum of the undetermined multiply the trial particular solution term by t. In this case,
coefficients represented by a single unknown, before substi- the trial particular solution for a step input would be:
tuting the particular solution into the differential equation.
Summing and renaming the unknown constants  x p = Ct (2.29)

C1 + C2 + C4 ≡ C4 and C3 + C5 ≡ C5 The coefficients in the particular solutions are determined


by substituting the trial solution into the output side of the
Hence, the particular solution for the polynomial forcing equation, and collecting like powers of t or like trig functions.
function, f 2 (t ) , Neatness is important to keep from drowning in constants.
There will be a factor comprised of the constants, an, from the
 x p = C 4 + C5 t + C 6 t 2 (2.24) differential equation and the constants, Cn, from the particular
solution multiplying each term on the output variable side.
The similarity between the mathematical forms of the input, These factors must equal the coefficients of the corresponding
or forcing function, and the corresponding particular solu- terms on the input side. If there is no corresponding term on
tion is not a coincidence. Physical systems to respond to their the input side, then the factor equals zero.
inputs, and we will find that ideal linear systems replicate the Solve for each of the coefficients, Cn. There are two rea-
form of their forcing functions in their steady-state response. sons why one would never use initial conditions to find the
A very important family of forcing functions is sinusoids. constants in a particular solution. First, the initial conditions
Trial solutions for sinusoidal forcing functions require only apply to the entire solution, not to just a part of it. A particu-
two terms, i.e., lar solution is part of the general solution. Second, the initial
conditions are irrelevant to a particular solution. A particular
 x p = C1 sin(ω t ) + C2 cos(ω t ) (2.25) solution describes the “steady-state” response of an energetic
system to a specific input. A system has reached steady-state
because of the relationship between the derivatives when the homogeneous solution has decayed to insignifi-
cance. The effect of the initial conditions decays with the ho-
 d (C1 sin(ω t )) mogeneous solution. An energetic system in steady-state is in
= C1ω cos(ω t ) = C1 cos(ω t ) (2.26)
dt
2.6  Initial Conditions 59

Table 2.2   Initial conditions Differential equation Initial conditions needed


needed for first, second, and
nth-order differential equations First-order
1
F (t ) =
dv a1
+ v ( )
v 0+
a0 dt a 0

1 a 0 dv
F (t ) = +v
a1 a dt
1
τ
Second-order
1 d 2 v a1 dv a 2
F (t ) = 2 + + v ( )
v 0+ and
dv 0+( )
a0 dt a 0 dt a 0 dt

nth-order
1 d n v a1 d n−1v
F (t ) = n +
a ( ) , … d v (0 )
dv 0+ n −1 +

a0 dt a 0 dt n−1
+ + n v
a0 ( )
v 0+ ,
dt dt n −1

dynamic equilibrium with the power supplied by the forcing Step 8. Check Your Solution Against Your Expectation of
functions. The steady-state response is solely dependent upon the Response of the Physical System.
maintaining the new dynamic equilibrium. First, check the units in your solution. Are they consistent?
Then plot your solution. Does the solution agree with your
Step 6.  Assemble the General Solution. engineering judgment? In other words, does it seem to be
The general solution is the superposition (the sum) of the correct based on your understanding of the physical system?
homogeneous solution and the particular solution for each If your solution doesn’t make sense, then check it. Look for
forcing function, as seen here: the most common errors, such as sign errors and transcrip-
tion errors first. If needed, rederive the system equation.
 xh + x p = x (2.30) Never accept any computational result, until you believe it.
If you don’t believe it, first look for errors. If there are no
The general solution is the physically real solution. It is the mathematical errors, then the model is suspect. Reevaluate
response that we see on the oscilloscope. The initial condi- your model. You may need to include energetic aspects of the
tions apply to the general solution! Do not ­apply the initial physical system that you initially believed were negligible.
conditions to the homogenous solution. The homogenous so-
lution is not the complete solution.
2.6  Initial Conditions
Step 7.  Use the Initial Conditions to Find the Undetermined
Coefficients. Our purpose, when solving differential equations, will be
The only undetermined coefficients in the general solution to predict the dynamic response of an energetic system. It
are A1, A2, …, An from the homogeneous solution. You will is impossible to predict the response of a system, if we do
have as many unknown coefficients from the homogeneous not know its starting point. Hence, we need “initial condi-
equation, as you have orders of derivatives on the output tions.” We must answer the following three questions: (1)
side of the system equation. You must have as many initial “How many initial condition terms are needed?” (2) “What
conditions, as you have orders of derivatives; otherwise, you initial conditions are needed?” (3) “How are those values de-
cannot solve for the unknown coefficients. The initial condi- termined?”
tions you need are those of the output variable and its n −1 The number of initial conditions needed equals the order
derivatives. Determine the initial conditions using the energy of the differential equation. For example, a first-order dif-
storage variables at time, t = 0 − , and the energetic equations ferential equation needs one initial condition, a second-order
used to derive the system equation. Solve for the unknown differential equation needs two initial conditions, etc. We
coefficients by evaluating the general solution and its deriva- must always know the value of the output variable at time,
tives at time, t = 0+ . t = 0+, which is the instant immediately after the input is ap-
Remember the mnemonic, “Use the initial conditions plied. The additional initial condition(s), if needed, are the
last.” values of the derivative(s) of the output variable at time,
60 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

t = 0+ . Using “n” as the order of the differential system equa- imaginary number, j, as the unit vector which defines the
tion, we must also know the values of n −1 derivatives of the imaginary direction of a complex plane. However, we will
output variable at time, t = 0+. not use conventional notation for a unit vector, and express
We will address how values of the initial conditions are the imaginary number j as j for two reasons. First, as noted
determined in Chap. 3, after we have discussed the energy above, complex numbers behave as vectors for some—but
storage variables of an energetic system, and their role as not all—vector operations. Specifically, complex numbers
“state variables.” Briefly, if we wish to predict the response are added as vectors; consequently, they can be decomposed
of an energetic system to our input (a power source), we into real and imaginary components. However, we will see
must know how much energy is in each energy storage ele- that complex numbers do not multiply in the same manner
ment in the energetic system, the instant before we apply as vectors. Second, we would then also need to represent the
the input. We will see that knowing the input power vari- unit vector in the real direction. Unfortunately, the custom-
able (which we always know, since it is under our control) ary unit vector for the x-direction is i . If we were to use unit
and the values of the energy storage or state variables is vector notation, we would write the complex number as thus,
sufficient information to determine every other power vari- Z = xi + y j. This is unacceptable, since it is possible to con-
able in a system. fuse the unit vector in the positive real direction, i , with the
more common symbol for the imaginary number, i. Conse-
quently, to avoid confusion, we will omit the unit vector in
2.7  Complex Numbers and Variables the real direction. Also, to save time, we omit the vector no-
tation for the imaginary number j. We will write the complex
The coefficients of the method of undetermined coefficients number Z in Cartesian form as follows, Z = x + jy, with the
are functions of the characteristic values (eigenvalues) of the imaginary number j in front of the magnitude, y, to make it
system. The eigenvalues of second- and higher-order differ- more conspicuous.
ential equations may be purely real numbers, complex con- The terms “real,” “imaginary,” and “complex” are unfortu-
jugate pairs, or a combination of real numbers and complex nate, since the latter two can intimidate students. Historically,
conjugate pairs. Eigenvalues which are complex conjugate both real and complex numbers were discovered (or invented)
pairs yield a homogeneous solution which is a decaying si- thousands of years after positive integers. Real numbers are
nusoid of the form: created by the arithmetic operations of addition, subtraction,
multiplication, and division. Complex numbers are created by
 f (t ) = Ae at sin (w t + φ ) (2.31) exponentiation of a negative real number to a fractional power
or by arithmetic operations on complex numbers alone or with
Euler discovered the mathematics which relates complex real numbers.
numbers and trigonometric functions. The exponential func- 1
tion with a purely imaginary exponent, ( −9) 2 ≡ −9 = 3 −1 = 3 j

 f ( jx) = e j x (2.32) An important aspect of complex numbers is the complex


conjugate. Conjugate means paired. Complex conjugates are
has proved important in physics and engineering. It greatly complex numbers with the same real component and equal
simplifies the representation and manipulation of decaying but opposite imaginary components as shown below.
oscillations. We will illustrate the utility of the complex ex-
ponential, e st = e(σ + jw )t = eσ t e jw t , and the use of Euler’s sine  Z, Z* = x ± jy (2.33)
and cosine formula, in solving for the step response of an os-
cillatory second-order system, using the method of undeter- The superscript, *, which appears in Eq. 2.33, indicates the
mined coefficients. We will first review complex numbers, complex conjugate, as shown in Fig. 2.12. Complex conju-
and then develop the mathematics of complex exponentials. gates arise when solving polynomials. The roots of even-
powered polynomials with real coefficients can be complex
conjugate pairs. The common example is the quadratic for-
2.7.1  Cartesian Form  Z = x + j y mula. The roots are complex conjugates when the quantity
within the radial, the discriminant, is negative.
The familiar Cartesian form of a complex number is as fol-
lows, Z = x + yi . It is conventional in systems dynamics  2 − a1 ± a12 − 4a 2 a 0
to use j rather than i as the symbol for −1, because i ap- a 2 s + a1 s + a 0 = 0 → s1 , s2 = (2.34)
2a 2
pears in our equations as electric current. We will view the
2.7  Complex Numbers and Variables 61

Imaginary Imaginary
j3 Z 3 = Z1 + Z 2
jy Z = x + jy Z2 = 3+j
= (-2+j2)+(3+j)
j2 = 1+j3
φ
Z1 = -2+j2 j Z 2 = 3+j
x Real

-3 -2 -1 1 2 3 Real

-jy Z *= x - jy
Fig. 2.14   Complex numbers, Z1 and Z2, added using the parallelogram
rule vector construction
Fig. 2.12   The complex number, Z = x + jy , and its complex conju-
*
gate, Z = x − jy . Notice that the complex numbers of the conjugate
pair have equal but opposite angles
To use the parallelogram construction to perform subtrac-
tion, Z2 is multiplied by − 1, which rotates the vector 180°.
Imaginary The rotated vector, − Z2, is translated to the tip of Z1 to yield
Z 1 = -2+j2 the sum Z3, as shown in the vector construction, Fig. 2.15.
j2
There is a shortcut for constructing the difference be-
Z2 = 3+j tween two vectors, which is to draw the resultant Z3 from
j the tip of the vector being subtracted, Z2, to the tip of the
vector, from which it is being subtracted, Z1, Fig. 2.16.
This vector shortcut will prove useful in constructions we
-3 -2 -1 1 2 3 Real will make by using complex variables.

Fig. 2.13   The complex numbers, Z1 = −2 + j 2 and Z 2 = 3 + j 2.7.1.2 Magnitude (Modulus) of a Complex


Number
The mathematical term for the length or magnitude of a
2.7.1.1 Addition and Subtraction of Complex complex number is its “modulus.” This is unfortunate, since
Numbers many other uses of the term modulus leads to confusion.
Addition and subtraction of complex numbers are easiest to We will substitute the term “magnitude” instead. The Py-
perform when complex numbers are expressed in Cartesian thagorean Theorem yields the magnitude of the real vector
form, as shown here: v, v = x1 + y1:

  v = x12 + y12 (2.37)


Z = Z1 + Z 2 = ( x1 + jy1 ) + ( x2 + jy2 )
= ( x1 + x2 ) + ( jy1 + jy2 ) (2.35) The ancient Egyptians, Babylonians, and Greeks invented
= ( x1 + x2 ) + j ( y1 + y2 ) (or discovered) geometry, long before either imaginary
numbers or negative numbers were invented. Consequent-
Z = Z1 − Z 2 = ( x1 + jy1 ) − ( x2 + jy2 ) ly, Euclidian geometry and trigonometry use real numbers.
 Physically realizable geometry, that is, what we can build, is
= ( x1 − x2 ) + ( jy1 − jy2 ) (2.36)
restricted to lengths, which are positive real numbers. Some
= ( x1 − x2 ) + j ( y1 − y2 ) geometric relationships have been extended to include nega-
tive real numbers. Although the Pythagorean Theorem can
The graphical constructions, which are used to add or sub- be used with negative real numbers, it cannot be used with
tract position vectors or force vectors, are also used to add any imaginary numbers. The problem is the square of imagi-
or subtract complex numbers plotted as vectors in a com- nary number
plex plane. Complex numbers have magnitude, direction,
and obey vector rules for addition and subtraction, Fig. 2.13. 
x12 + ( jy1 ) =
2
x12 + j 2 y12 = x12 − y12 ≠ Z1 (2.38)
For example, the same graphical constructions based on the
“parallelogram rule,” used to add or subtract position vectors
or force vectors, are also used to add, Fig. 2.14, or subtract Notice that the Pythagorean Theorem, as applied to
complex numbers, as illustrated in the following examples. v = x1 + y1, yields the correct magnitude of Z1. Consequently,
62 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Fig. 2.15   The difference be- Imaginary


tween complex numbers, Z1 and
Z2, constructed using the paral- j2
-Z2 = -(3+j)
lelogram rule vector for vector
addition. Multiplying a vector or
Z1 = -2+j2
complex number by −1 reverses j Z 2 = 3+j
Z 3 = Z1 - Z2
its direction = (-2+j2)-(3+j)
= -5+j
-6 -5 -4 -3 -2 -1 1 2 3 Real

-Z2 = -(3+j) -j

Fig. 2.16   Vector shortcut to con- Imaginary


struct the difference, Z3 = Z1 − Z2.
The difference Z3 is drawn from Z 3= Z 1 - Z 2
j2
the tip of Z2 to the tip of Z1 -Z2 = -(3+j)

Z 1= -2+j2
Z 3= Z 1- Z 2 j
Z 2 = 3+j
= (-2+j2)-(3+j)
= -5+j
-6 -5 -4 -3 -2 -1 1 2 3 Real

-Z2 = -(3+j) -j

magnitude of a complex number is calculated using the Py- squared. We will consider the dot product of two complex
thagorean Theorem, after removing the imaginary number numbers to be undefined:

 Z1 ≡ x1 + jy1 = x12 + y12 (2.39)  Z1 · Z 2 = undefined (2.42)

Although the dot product of complex numbers can be de-


2.7.1.3 Multiplication and Division of Complex fined for some applications, the cross-product of two com-
Numbers plex numbers is not defined:
Although complex numbers add and subtract as vectors, they
are not, strictly speaking, vectors. Complex numbers both  Z1 × Z 2 = undefined (2.43)
lack operations of vectors and possess operations which vec-
tors do not. The dot product of a real vector against itself In one sense, we are splitting hairs, because there is no cross-
yields its magnitude squared. product defined for other quantities which we refer to as
vectors. For example, we may have a third-order dynamic
 v · v = ( x1 + y1 ) · ( x1 + y1 ) = x1 x1 + y1 y1 (2.40) system with the state vector x comprised of the state vari-
ables, pressure, translational velocity, and torque. The cross-
The dot product of a complex number against itself does not product of the state vector, x, and a position vector, r, is not
yield the square of its magnitude. defined either.
Complex numbers in Cartesian form are multiplied as an
 Z1 i Z1 = ( x1 + jy1 )i( x1 + jy1 ) = x1 x1 + jy1 jy1 algebraic expansion.

Z1 i Z1 = x1 x1 + j 2 y1 y1 = x1 x1 − y1 y1 (2.41) 
Z1Z 2 = ( x1 + jy1 )( x2 + jy2 ) = x1 ( x2 + jy2 ) + jy1 ( x2 + jy2 )
Consequently, the dot product of two complex numbers is ei- Z1Z 2 = x1 x2 + jx1 y2 + jy1 x2 + j jy1 y2
ther undefined, since it does not yield the magnitude squared,
or is defined as a special case, in order to yield the magnitude Z1Z 2 = x1 x2 − y1 y2 + j ( x1 y2 + y1 x2 ) (2.44)
2.7  Complex Numbers and Variables 63

Im Im
Z1 = -2+j2 Z1 = -2+j2
j2 j2

Z1 Z2 = 3+j Z1

|Z 1
|Z 1
j

|
|Z 2|
j

Z2 α
-2 -1 1 2 3 Re
-2 -1 1 2 Re
Fig. 2.17   Complex numbers, Z1 and Z2, represented in polar form
Fig. 2.18   Complex numbers should be sketched before computing
their angles. Complex numbers falling in the second and third quad-
A property which complex numbers share with scalars (but rant cannot be computed directly using the inverse tangent algorithm of
most calculators. A construction is needed
not with vectors) is division. Vector division is not defined.
The ratio of two complex numbers is evaluated in Cartesian
form, by multiplying the numerator and denominator by the inverse tangent of the magnitude of the imaginary compo-
complex conjugate of the denominator. nent over the real component.

 Z1 x + jy1 x + jy1 x2 − jy2  1


= 1 = 1 Z 2 = tan −1   = 0.32 rad
Z 2 x2 + jy2 x2 + jy2 x2 − jy2  3

x1 ( x2 − jy2 ) + jy1 ( x2 − jy2 )


= The magnitude (modulus) of Z2 is calculated with the Py-
x2 ( x2 − jy2 ) + jy2 ( x2 − jy2 ) thagorean Theorem, using the magnitudes of the real and
x1 x2 − jx1 y2 + jx2 y1 − j 2 y1 y2 imaginary components.
=
x2 x2 − jx2 y2 + jx2 y2 − j 2 y2 y2
Z 2 = 32 + 12 = 3.16
x1 x2 + y1 y2 + j ( − x1 y2 + x2 y2 )
= Z2 in polar form is
x22 + y22
(2.45)
Z 2 = ( Z 2 , Z 2 ) = ( 0.32 rad, 3.16)
2.7.2  Polar Form: Z = Z  Z
The inverse tangent algorithm, commonly used by handheld
There are two other forms for complex numbers. The first calculators, can introduce an error depending on the quad-
is the polar representation, familiar from cylindrical coordi- rant, in which the number complex falls.
nates, shown in Fig. 2.17. Recall that a polar representation The algorithm only returns angles between 90° and − 90°,
of a vector in a two-dimensional plane specifies the length, corresponding to complex numbers in the first and fourth
or magnitude, of the vector, and the angle that vector makes quadrants of a plane. To avoid error, always sketch a com-
with the positive x-axis. Likewise, the polar representation plex number as a vector, before calculating its angle. If the
of a complex number requires the length, or magnitude, of complex number falls in the second or third quadrants, then
the complex number, and the angle of the complex number. compute the angle using a geometric construction. For ex-
The magnitude (modulus) is a purely real number. The angle ample, the angle of Z1 is calculated by subtracting angle α
of a complex number is measured counter-clockwise from from 180° or π rad, as seen in Fig. 2.18 and Eq. 2.47.
the positive real axis. Unfortunately, the mathematical term
for the angle of a complex number is its “argument,” which  2
Z1 = 180o − α = 180o − tan −1   = 180o − 45o = 135o
is more commonly used to identify the operand or input to  2
a function. −1  2  π
 Z1 = π − α = π − tan  2  = π − 4 = 2.36 rad (2.47)
  opposite   y
Z = tan −1  = tan −1   (2.46)
 adjacent   x The magnitude of Z1 is

In this example, express the complex number, Z 2 = 3 + j, Z1 = 22 + 22 = 2.83


in polar form. Z2 falls in the first quadrant of the complex
plane. The angle (or argument) of Z2 is calculated as the Z1 in polar form is

Z1 = ( Z1 , Z1 ) = ( 2.36 rad, 2.83)


64 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

The angle of a complex number is not unique. It is impossible Im


to distinguish between an angle α and the angles, β = α ± n 2π
j sin(φ) e jφ
rad (or β = α ± n360o), since adding a complete revolution
about the origin of 2π rad (or 360°) to the angle yields the same

1
|=
orientation. The “principal angle” has two definitions. The prin-


cipal angle of a complex number is either the smallest positive

|e
angle, 0 ≤ φ < 2π rads ( 0 ≤ φ < 360o ), or the smallest angle, φ
−π ≤ φ ≤ π rads ( −180o ≤ φ ≤ 180o ), needed to describe the
orientation of the complex number. The latter definition allows cos(φ) Re
a negative, purely real number to be described by either a posi-
tive or negative angle, which is useful in evaluating functions Fig. 2.19   Euler’s Equation is ejφ = cos(φ) + j sin(φ). The projection of
ejφ onto the real axis is cos(φ). The projection onto the imaginary axis is
which reach limiting values approaching the negative real axis j sin(φ). The magnitude or length of ejφ is one
from either above or below.

where e jφ is a “complex exponential.” We will refer to it as


2.7.3  Euler’s Equations a “complex exponential unit vector,” since its magnitude is
unity. Notice that the angle φ is multiplied by the imaginary
Euler recognized that cos(φ ) + j sin(φ ) = e jφ . The proof of number j in the exponent. It is best to remember Euler’s
the expression uses power series for cos( φ), sin( φ), and ex: equation, Eq. 2.51, graphically, as shown in Fig. 2.19, rath-
 er than symbolically.
φ0 φ2 φ4 φ6
cos(φ ) = − + − + (2.48) The complex exponential unit vector is an important
0! 2! 4! 6! quantity, as we shall see. The product of the complex expo-
nential unit vector and a real number is a complex number
(2.49) φ φ φ φ 1 3 5 7
sin(φ ) = − + − + with the magnitude of the real number, and the direction or
1! 3! 5! 7!
angle of the complex exponential unit vector, Eq. 2.52. This
x 0 x1 x 2 x 3 is known as the complex exponential form of the complex
 ex = + + + + (2.50)
0 ! 1! 2 ! 3! number.

By definition, 0 ! = 1 . Recall φ 0 = e0 = x 0 = 1.  Z = Z e jφ where φ = Z (2.52)


Starting with the power series for ex, let x = jφ:
Again, it is important to note that the angle of a complex
e jφ = 1 + ( jφ ) exponential unit vector is every term in the exponent which

+
( jφ )2 + ( jφ )3 + ( jφ )4 + ( jφ )5 + ( jφ )6 + ( jφ )7  is multiplied by the imaginary number j. Further discussion
2! 3! 4! 5! 6! 7! follows in Sect. 2.7.4.
The trigonometric projection of a point on a unit circle
j 2 φ 2 j 3φ 3 j 4 φ 4 j 5 φ 5 j 6 φ 6 j 7 φ 7 centered on the origin onto the axes of a real plane is the
e jφ = 1 + jφ + + + + + +  same as the projection of the complex exponential, e jφ , onto
2! 3! 4! 5! 6! 7!
the axes of a complex plane, Fig. 2.20. The vertical axis of
Use j 2 = −1, the complex plane is the imaginary axis. Hence, the projec-
tion of e jφ onto the vertical axis is jsin( φ).
φ2 jφ 3 φ 4 jφ 5 φ 6 jφ 7
e jφ = 1 + jφ − − + + − − 
2! 3! 4! 5! 6! 7! 2.7.3.1  Euler’s Cosine Formula
Euler’s cosine formula, Eq. 2.53, is easiest to understand and
Collect terms with and without j. remember as a vector construction. The construction starts
with a complex exponential unit vector, e j φ . We will place it
 φ2 φ4 φ6   jφ 3 jφ 5 jφ 7  in the first quadrant. The cosine formula is derived by sum-
e jφ =  1 − + −  +  jφ − + −  ming e j φ with its complex conjugate, e − j φ , which is its mirror
 2! 4! 6!   3! 5! 7! 
image on the other side of the real axis with the opposite
angle. Summing complex conjugates yields a purely real re-
Use the series for cos( φ) and sin( φ) sult, since the imaginary components are equal but opposite.
The vector construction, Fig. 2.21, shows that the purely real
 e jφ = cos (φ ) + j sin (φ ) (2.51)
2.7  Complex Numbers and Variables 65

a y b Im Im
-jφ
sin(φ) jsin(φ) e
jφ 2jsin(φ) -e

1
|=
|

|e jφ
|1
φ φ
cos(φ) x cos(φ) Re -jφ
-e
jsin(φ) e jφ

1
|=
|e jφ

x 2 + y2 = 1 cos(φ) + jsin(φ) = e φ
Fig. 2.20  a The relationship between the trigonometry of a unit cos(φ) Re
circle in a real plane and b Euler’s formula in a complex plane,

e jφ = cos (φ ) + j sin (φ )

e-jφ
Im jφ
e e j φ − e− j φ
jsin(φ) Fig. 2.22   Euler’s sine formula sin (φ ) =
-φ 2j
1
|=
|ejφ

φ e -jφ ed as separate factors. However, the complex exponential


cos(φ) 2cos(φ) Re form is a mathematical expression, whereas the polar form
-φ is actually just the polar coordinates of the complex num-
ber. This distinction is illustrated below with the Cartesian
coordinates of a complex number versus the Cartesian form
-jφ
e of that number which is the sum of a real and an imaginary
number.
e j φ + e− j φ
Fig. 2.21   Euler’s cosine formula cos (φ ) =
2 Z = ( x, jy ) Z = x + jy

result is twice the length of the projection cos(  jφ). Conse- A similar distinction exists between the polar coordinates of
quently, the sum e j φ + e − j φ must be divided by two. a complex number and the complex number in exponential
form which is a product.
 e jφ + e− jφ
cos (φ ) = (2.53)
2 Z = ( Z, Z ) Z = Z e Z

2.7.3.2  Euler’s Sine Formula The magnitude |Z| is calculated using the Pythagorean The-
Euler’s sine formula, Eq. 2.54, is only slightly more compli- orem with the magnitudes of the real and imaginary com-
cated. Starting with e j φ , the complex exponential unit vector ponents, Eq. 2.39. Remember to drop j from the imaginary
is reflected relative to the real axis, by inverting the sign of component!
the exponent, i.e., the angle. The reflection is then rotated 180°
by multiplying it by negative one, creating the vector, which Z = x2 + y 2
when summed with the original complex exponential, yields a
resultant which is purely imaginary and twice the length of the The angle is calculated using the inverse tangent function as
projection, sin( φ), Fig. 2.22. Consequently, it must be divided shown below:
by 2j to yield
  opposite   y
e j φ + −e − j φ Z = tan −1   = tan −1  
sin(φ ) = (2.54)  adjacent   x
2j
The algorithm, used in most calculators for the inverse tan-
2.7.4  Complex Exponential Form Z e jφ gent, only reports angles in the first and fourth quadrants of
the complex plane. Sketch complex numbers, before calcu-
The exponential form of a complex number Z e jφ resembles lating the angle between them. This is demonstrated below
the polar form, in that the magnitude and angle are represent- using the complex numbers Z4 and Z5 shown in Fig. 2.23.
66 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Im Im
j0.32 Z 6= 2+j4 = 4.46 e j1.11
Z 4 = 1.41e j0.79 Z 5 = 3.16e j4
j

|Z 5|
Z4
4|
|Z

j3
Z5

|
46
1 2 3 Re

|4.
j2
Z 6 =1.11
Fig. 2.23   Complex numbers, Z 4 and Z5 , in complex exponential form

Z 4 = 1.41e j0.79 Z5 = 3.16e j0.32


j
 Z = 12 + 12 = 1.41
|Z 5|

4|
 4 Z4

|Z
Z4 = 1 + j  −1  1 0 π
Z 4 = tan   = 45 = = 0.79 rad Z5
 1 4
1 2 3 Re
j Z 4 j 0.79
Z4 = Z4 e → Z 4 = 1.41e
Fig. 2.24   The product of complex numbers, Z 4 and Z5 in complex ex-
ponential form
and

 Z = 12 + 32 = 3.16 The ratio of exponentials equals the base raised to the differ-
 5 ence of the exponents.
Z5 = 3 + j  −1  1  0
Z5 = tan   = 36.9 = 0.32 rad
 3 102 105
2−3 −1 3 5− 2
= 10 = 10 and 10 = 10 =
103 102
Z5 = Z5 e jZ5 → Z5 = 3.16e j 0.32
The fact that any base raised to the zero power equals one is
Addition and subtraction of complex numbers expressed as easy to prove
complex exponentials must be performed in vector form.
10
Consequently, addition and subtraction is easier in Cartesian = 10 1 ⋅10 −1 = 10 1−1 = 100 = 1
form than in complex exponential form, Fig. 2.14. However, 10
multiplication, division, and exponentiation are much easier
in complex exponential form. Recall these three properties 2.7.4.1 Multiplication in the Complex Exponential
of exponentials: Form
 The product of two complex numbers in complex exponen-
e a eb = e a + b (2.55) tial form is evaluated by first associating the real magnitudes
and complex exponential unit vector factors.
ea
 = ea −b (2.56)
eb
Z 4 Z5 = Z 4 e j Z4 Z5 e j Z5 → Z 4 Z5 = Z 4 Z5 e j Z4 e j Z5
 (e ) a n
= e n·a (2.57)
The magnitude of the product Z4 Z5 is the product of the
The properties of exponentials are independent of the con- magnitudes, |Z4| and |Z5|. The product of the complex expo-
stant we use as the base. We will illustrate the properties nential unit vectors is evaluated, using the property of ex-
using base 10, because its arithmetic is familiar, but the prop- ponentials expressed in Eq. 2.55, Fig. 2.24. The product of
erties apply to base e, as well. Further, the properties apply exponentials equals an exponential raised to the sum of the
whether the exponents are constants, variables, or functions. two exponents.
The product of exponentials equals the base raised to the
e j Z4 e j Z5 = e j Z4 + j Z5 = e ( 4 5 )
j Z + Z
sum of the exponents.

102 ·103 = 102 + 3 = 105 and 105 = 101+ 4 = 101 ·104 The imaginary number j has been factored and is multiplying
the sum of the two angles. Any quantity multiplied by j in the
2.7  Complex Numbers and Variables 67

Im “Transfer functions,” introduced in Sect. 2.10.4, are created


Z 1 = 2.83e j2.36 j2
by performing the Laplace transform on a system equation,
and then creating the ratio of the transformed output variable
Z1 over the transformed input variable. The transformation of
Z 2 = 3.16e j0.32
j the operation of differentiation, with respect to time, leads to
powers of the complex variable, s = σ + jw . A transfer func-
Z2 tion is thus a ratio of polynomials in the complex variable.
-3 -2 -1 1 2 3 Re
Z 7 = 4.24 rad Z 7 = -2.04 rad  N (s)
F (s) = (2.60)
D (s)
Z 7 = 1.12e -j2.04
In Eq. 2.60, N( s) and D( s) represent the numerator and de-
Fig. 2.25   The ratio of complex numbers Z2 over Z1, in complex ex- nominator polynomials. The general case is the following:
ponential form
 b0 s m + b1 s m −1 +  + bm
F (s) = where m ≤ n (2.61)
a 0 s n + a1 s n −1 +  + an
exponent of an exponential is the angle of the exponential.
The product Z4 Z5 is The denominator of a transfer function is the characteristic
function which, when set equal to zero, forms the charac-
 Z 6 = Z 4 Z5 = Z 4 Z5 e ( 4 5 ) teristic equation. Its roots are the characteristic values or ei-
j Z + Z
(2.58)
genvalues of the system. Systems with complex conjugate
Z 6 = (1.41)(3.16)e j (0.79 + 0.32) = 4.46e j1.11 eigenvalues have oscillatory homogeneous responses.
When the numerator and denominator polynomials of the
complex function Eq. 2.61 are evaluated for a specific com-
2.7.4.2  Division in the Complex Exponential Form plex value, such as s = σ + jw , the result is the ratio of two
The ratio of two complex numbers in complex exponential complex numbers.
form is evaluated in a similar fashion. First, associate the
real magnitudes and complex exponential unit vector factors.  Z N xN + jy N
F (σ + jw ) = = (2.62)
Z D xD + jyD
Z2 Z e  Z  e j Z2
j Z 2

= 2 j Z1 =  2  j Z1
Z1 Z1 e  Z1  e This ratio must be reduced to a single complex number. One
method is to perform the division per Eq. 2.45, i.e., to multi-
The magnitude of the ratio is the ratio of the magnitudes. The ply numerator and denominator by the complex conjugate of
ratio of the complex exponential unit vectors is evaluated the denominator, which will yield a purely real number as the
using the property of exponentials expressed in Eq. 2.56. denominator. Alternatively, the numerator and denominator
The angle of the resultant’s complex exponential unit vector can both be expressed as complex exponentials, that is, by
equals the angle of the numerator’s angle minus the denomi- using complex exponential unit vectors, and the ratio then
nator’s angle. simplified, as shown in Eq. 2.63.

e j Z2  ZN Z e jφ N Z
= e( 2 − 1 ) = e (Z2 − Z1 )
j Z j Z j F (σ + jw ) = Z = = N jφ D = N e j ( φ N − φ D ) (2.63)
e j Z1 ZD ZD e ZD

The ratio of the complex numbers Z1 over Z2 equals the ratio Unlike the roots of real numbers, the roots of complex ex-
of their magnitudes and the difference between the angles of ponentials are easy to calculate manually. We will illustrate
the numerator and denominator, Fig. 2.25. the properties of powers and roots of complex exponentials
using as an example the purely real number negative one ex-
 Z2 Z e j Z2 Z j − pressed as a complex exponential, Fig. 2.26. If the primary
= 2 j Z1 = 2 e (Z2 Z1 ) (2.59) angle is defined as the smallest angle which locates a com-
Z1 Z1 e Z1
plex number, then negative real numbers have two primary
angles, ±180o = ± π rad
Z 2 3.16e j 0.32 3.16 j (0.32 − 2.36)
Z7 = = = e = 1.12 e − j 2.04
Z1 2.83e j 2.36 2.83  −1 = e j π = e − j π (2.64)
68 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Im Im jφ1 _
π
φ=π e = ej 3
e jπ = -1 φ 3 = π = 3π
π
_
φ1 = 3

Re e = -1
e-jπ = -1 Re
φ = -π -1 -π
__
φ2 = 3
Fig. 2.26    Negative one expressed as a complex exponential,
−1 = e jπ = e − jπ jφ2 _
-j π
e =e 3

Fig. 2.28   The cube roots of negative one


Im π
_
j φ1 j2
j =e = e
we find two of the three roots. The third root is negative one
π itself. Negative one cubed equals negative one:
φ1 = _
2
e jπ = -1
( −1)3 = (e j π )
3
-jπ Re −1 = e j π →
e = -1 -π
φ2 = __
2 −1 = e j 3π → − 1 = e j (π + 2π )
π
_
j φ2 -j 2
-j = e =e The last expression can be written as the product of two com-
plex exponentials using Eq. 2.55,
Fig. 2.27   Negative one has two square roots. We choose to use the
positive square root as the imaginary number, j −1 = e j (π + 2π ) → − 1 = e j π e j 2π

from which the following holds:


When the operation of obtaining the square root of a quantity
is expressed as the quantity raised to the one-half power, i.e.,  e j 2π = e j (2π − 2π ) = e j 0 = e0 = 1 (2.66)
1
x = x , we see that the angle (argument) of the square root
2
Notice that negative one cubed is an “alias” of negative one.
of a complex number is one-half of the angle of the number,
Complex numbers expressed in polar or complex exponen-
Eq. 2.65 and Fig. 2.27.
tial form are “not unique,” meaning that there is not a single
 π π complex number described by an angle and magnitude. It is
j −j
−1 = e j π = e − j π → j = e 2
=e 2
(2.65) impossible to distinguish an angle θ from the angle θ ± 2nπ
since they overlie one another. An alias of a complex number
The imaginary axis is perpendicular to the real axis, thereby has the same magnitude (modulus) but with an angle 2nπ
confirming our calculation of the angle of j as one-half of the rads greater or less than its principal angle.
angle of minus one which is 90° or π/2 rad.
Notice that although we defined the imaginary number,
j, as the square root of negative one, there are actually two 2.7.5  Rotating Complex Exponential Unit Vector
square roots of negative one, j and − j. Our preference for
positive numbers leads us to use j. We developed Euler’s sine and cosine formula with a con-
Another interesting example is the derivation of the cube stant angle, φ. However, any quantity in the exponent of
roots of negative one. If we express negative one as a com- an exponential multiplied by the imaginary number j is an
plex exponential, and the cube root as a fractional exponent, angle. An angle which is the product of time, t, and a con-
stant frequency ω, θ (t ) = w t , where ω is the “angular” fre-
−1 = e j π = e − j π quency in radians per second, increases at a constant rate. If
π π
the angle has a non-zero value at time, t = 0, then the sum-
3
−1 = 3 e j π = 3 e − j π → 3
−1 = e
j
3
=e
−j
3 mation, θ (t ) = w t + φ is used, where the constant φ is the
angle at t = 0.
2.7  Complex Numbers and Variables 69

Im 1.0
e j(ωt+φ) eσt Exponential Decay Envelope

0.5
ωt eσt sin(ωt) = e-t sin(10t)
e jφ
f1(t) 0
φ

Re -0.5

Fig. 2.29 A rotating unit vector. The unit vector rotates in the positive, -eσt
counter-clockwise direction around the origin of the complex plane -1.0
0 1 2 3 4 5
t, sec
A rotating complex exponential unit vector, which ad- Fig. 2.30 Plot of Eq. 2.69 with M 1 = 1, σ = −1 , w = 10 , φ 1 = 0 , and
vances in the positive, counter-clockwise around the origin C1 = 0
of its complex plane as time increases, has a positive angular
frequency ω and a time-varying angle of either θ (t ) = w t or
θ (t ) = w t + φ , Fig. 2.29. 1.0
eσt Exponential Decay Envelope
(2.67)
e jwt = cos(w t ) + j sin(w t ) σ = -1
0.5
 e j (wt + φ ) = cos(w t + φ ) + j sin(w t + φ ) (2.68) eσt cos(ωt) = e-t cos(10t)

A great utility of complex exponential unit vectors is the ef- f 2(t) 0


ficiency with they which represent decaying oscillations. A
decaying oscillation is the product of an amplitude M, which
is a scaling factor, a real exponential decay eσt, and a sinusoid -0.5
with the form:
-eσt
 f1 (t ) = M 1eσ t sin (w t + φ1 ) + C1 (2.69) -1.0
0 1 2 3 4 5
t, sec
or
Fig. 2.31  Plot of Eq. 2.70 with M 2 = 1, σ = −1, w = 10, φ 2 = 0, and
 f 2 (t ) = M 2 e cos (w t + φ2 ) + C2
σt
(2.70) C2 = 0

where
150
f 2 (t ) = M σt
e cos (w t + φ2 ) + C2 σ = +1
Unstable Exponential
2     Growth Envelope eσt
Amplitude Decay Oscillation Steady 100
State
Value eσt cos(ωt) = et cos(10t)
50
and w is the angular frequency in radians per second, σ is
the decay rate of the real exponential, oscillation, and φ is f 3(t) 0
the “phase angle” or “phase shift”. The term “phase” means
position in a cycle, as in “phase of the moon”. -50
An oscillation decays, Figs. 2.30 and 2.31, if the real com-
-100
ponent of the eigenvalue, σ, is negative, since e − at = 1at . -eσt
e
Conversely, the oscillation grows, Fig. 2.32, if the real com- -150
5
0 1 2 3 4
ponent of the eigenvalue, σ, is positive. It is an unstable ex- t, sec
ponential growth and has no limit in theory. An unstable sys-
tem accumulates energy in the system at an increasing rate. Fig. 2.32 Plot of Eq. 2.70 with M 2 = 1, σ = +1, w = 10, φ 2 = 0 , and
If the homogeneous response of the system is oscillatory, the C2 = 0
70 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

accumulation of energy in the system increases the ampli- a b


tude of the oscillations. In a physical system, eventually, if f(t) = e σte jωt = e (σ+jω)t = e st Im[f(t)] = e σtsin(ωt)
the power supplied to the system is sufficiently large, the j1.0 1.0
amount of energy stored in the system exceeds the capacity
or strength of the system, thereby destroying it. If the power j0.5 0.5
supplied to the system is not large enough to destroy it, the
j0.0 0.0
systems will reach a “bounded” unstable state, in which its
oscillations remain at the limiting amplitude. A bounded un- -j0.5 -0.5
stable state is a failure mode for the response of a system,
since we want systems to follow the input we give them, not -j1.0 -1.0
-1.0 -0.5 0.0 0.5 1.0 0 1 2 3 4
to go into endless oscillations. time
The result of a dynamic calculation is presented in the
c Re[f(t)] = e σtcos(ωt)
form of Eqs. 2.73 or 2.74, as the product of a real exponen- -1.0 -0.5 0.0 0.5 1.0
0
tial and a sinusoid. In fact, the execution of the calculation
is greatly simplified when the oscillation is represented by 1
a complex exponential unit vector instead of a sinusoid. We

time
will develop the mathematical technique below, and illus- 2
trate its use in example 2.7.6.
3

(2.71)
f (t ) = Meσ t e j w t = Meσ t + j w t = Me(σ + j w ) t = Me st 4

This expression is a rotating exponential unit vector, with a Fig. 2.33   a Plot of a rotating complex exponential,
f (t ) = Me st = Me(σ + jw )t , where M = 1 and s = −1 + j 2. b The projection
time varying magnitude. If the real component, σ, is nega- of Me( ) onto the vertical axis is Eq. 2.68, Im  f (t ) = eσ t sin (w t ) ,
σ + jw t

tive, the vector’s magnitude decreases with time. The tip of c The projection of Me(σ + jw )t onto the horizontal axis is Eq. 2.69,
the vector traces a “logarithmic” spiral, Fig. 2.33a, a familiar Re  f (t ) = eσ t cos (w t )
natural shape. A cross-section of a snail’s shell is a logarith-
mic spiral, because a snail’s growth rate is proportional to its
size. The trace of a snail’s growth is a spiral from the origin Im
j j(ωt1+φ)
out. Conversely, the decay rate of a dynamic system’s oscil- e
lation is proportional to the amount of energy stored in the ωt 2
system. The trace of the response of a dynamic system is a j(ωt 2+φ)
e
spiral toward the origin. Software and calculators have the ωt1 jφ
e
functions Re [ ] and Im [ ], which return only the magnitude
of the real or imaginary components of a complex number,
φ
respectively. The imaginary component of the complex ex-
ponential is expressed, as Eq. 2.72, -1 1 Re

 f (t )] = Im[ Me st ] = Meσ t Im[e j wt ] = Meσ t sin(w t )
Im[ (2.72) -jφ
e
-ωt1
Note that Im[ f (t )] is a purely real number. It is the magni-
tude of the imaginary component. The magnitude of the real -j(ωt 2+φ) -ωt2
e
component is expressed as Eq. 2.73. -j(ωt 1+φ)
-j e
 f (t )] = Re[ Me st ] = Meσ t Re[e j wt ] = Meσ t cos(w t ) (2.73)
Re[
Fig. 2.34 Counter-rotating complex conjugate exponential unit vec-
When the characteristic equation of a second-order differen- tors, e j(wt +φ ) and e − j(wt +φ )
tial system equation yields complex conjugate eigenvalues
(characteristic values), then the homogeneous solution will Eqs. 2.49 and 2.50, reproduced below. The corresponding
be oscillatory. If input to the system is a step, then a solu- vector constructions are Figs. 2.22 and 2.23.
tion (the response of the system) will exceed or “overshoot”
the final, steady-state value of the output variable. The func- e j θ + e− j θ e j θ − e− j θ
tions, Re[ ] and Im[ ], are not as useful, when manipulating cos (θ ) = and sin (θ ) =
2 j2
complex conjugates, as Euler’s cosine and sine formula,
2.7  Complex Numbers and Variables 71

Euler’s cosine and sine formulas contain complex exponen- with constant coefficients which have complex conjugate ei-
tial unit vectors with equal but opposite angles. When the genvalues and, hence, an oscillatory, homogeneous solution,
angle θ is the time-varying angle θ = w t + φ then the com- with the following example:
plex exponential unit vectors e j (wt + φ ) and e − j (wt + φ ) counter-ro-
tate about the origin of a complex plane, Figs. 2.33 and 2.34. d 2 v dv
8= + + 6v
The resultant of e j (wt + φ ) + e − j (wt + φ ) lies on the real axis, dt 2 dt
dv(0+ )
e j (wt + φ ) + e − j (wt + φ ) = 2 cos (w t + φ )
(2.74) These are the initial conditions: v(0+ ) = 0 and = 10.
dt
or The units are not consistent in this textbook example. The
e j (wt + φ ) + e − j (wt + φ ) input is a constant, rather than a function of time, but we will
 = cos (w t + φ ) (2.75) ignore that physical nonsense, and proceed to the solution.
2
We will begin by solving the homogeneous equation. The
Likewise, the resultant, e j (wt + φ ) − e − j (wt + φ ) , drawn from the tip first step is to formulate and solve the characteristic equa-
of e − j (wt + φ ) to the tip of e j (wt + φ ), is parallel to the imaginary tion. Again, the solution of the homogeneous equation is al-
axis: ways the following:

 e j (wt + φ ) − e − j (wt + φ ) = j 2sin (w t + φ ) (2.76) vh (t ) = A1e s1t + A2 e s2t

or The number of terms equals the number of eigenvalues


which equals the order of the differential equation. Forming
 e j (wt + φ ) − e − j (wt + φ )
= sin (w t + φ ) (2.77) the characteristic function by means of the ­Laplace transfor-
j2 mation

The complex exponential unit vector may be intimidating at  d 2 v dv 


first, because it is new mathematics. However, it is much
L {0} = L  2 + + 6v 
 dt dt 
easier to deal with than the corresponding product of a real
exponential, and the sum of the product cosine and sine, to  d 2v   dv 
yield a sinusoid with phase shift, using one of the four fol-
L {0} = L  2  + L   + L {6v}
 dt   dt 
lowing trigonometric identities:

 sin (α + β ) = sin (α ) cos ( β ) + cos (α ) sin ( β ) (2.78) ( )


0 = s 2V ( s ) + sV ( s ) + 6V ( s ) → 0 = s 2 + s + 6 V ( s )

 sin (α − β ) = sin (α ) cos ( β ) − cos (α ) sin ( β ) (2.79) Assuming V( s) = 0 is the trivial solution. Set the characteris-
tic function equal to zero, to form the characteristic equation.
 cos (α + β ) = cos (α ) cos ( β ) − sin (α ) sin ( β ) (2.80)
s2 + s + 6 = 0
 cos (α − β ) = cos (α ) cos ( β ) + sin (α ) sin ( β ) (2.81)
The roots of the characteristic equation are the characteristic
The expansion of the expression from the complex expo- values or eigenvalues
nential form to the trigonometric form, with the phase shift
represented as the sum of products of sinusoids, is reason −1 ± 12 − 4 ⋅ 6
s1 , s2 = σ ± jw → s1 , s2 =
enough to work with complex exponentials: 2
eσ t (e j (w t + φ ) + e − j (w t + φ ) ) = 2eσ t cos(wt + φ ) s1 , s2 = −0.5 ± j 2.40
= 2eσ t (cos(wt ) cos(φ ) − sin(wt ) sin(φ ))
It is more efficient and less error-prone to perform as much
of the solution symbolically as possible. Substitute in the nu-
2.7.6 Example: Solution of a Homogeneous merical values of the eigenvalues only when necessary.
Equation using Complex Exponential We now solve the particular equation, using the form of
Unit Vectors the input as the form of the particular solution. The input
is a constant. Hence, the particular solution is the unknown
We will illustrate the use of complex exponential unit vec- constant C, as follows:
tors to solve a second-order ordinary differential equation,
vp = C
72 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Substitute and evaluate as follows: Now, evaluate dv(t ) for the known value at time, t = 0+ :
dt
d 2vp dv p d 2C dC
8=
dt 2
+
dt
+ 6v p → 8= 2
+
dt 0 dt 0
+ 6C ( ) = 10 = A s e
dv 0+ s1 0+
+ A 2 s2 e s2 0
+

1 1
dt
8
C= = 1.33 10 = A1 s1 e0 1 + A 2 s2 e0 1
6

Assemble the general solution: 10 = A1 s1 + A2 s2

v (t ) = vh (t ) + v p (t ) → v (t ) = A1e s1t + A2 e s2t + C We have two equations with two unknowns, A1 and A2.

v (t ) = A1e s1t + A2 e s2t + 1.33 −1.33 = A1 + A2

+
10 = A1 s1 + A2 s1
We now use the initial conditions, v(0 ) = 0 and
+
dv(0 ) / dt = 10, last with the general solution, to determine Resist the temptation to substitute the values of the eigenval-
the undetermined coefficients, A1 and A2: ues into the equations. Numbers are only needed for arithmetic.
Algebra is easier with symbols. It is easier to see the required
( )
v 0+ = 0 = A1e s1 0 + A 2 e s2 0 + 1.33 operations and requires less writing to execute them. Choose a
solution method from: (1) direct substitution and elimination,
0 = A1 e0 1 + A 2 e0 1 + 1.33 (2) Gaussian elimination, or (3) linear algebra. Direct substitu-
tion is easy for a system of equations with two unknowns.
0 = A1 + A2 + 1.33 → − 1.33 = A1 + A2
−1.33 = A1 + A2 → A1 = −1.33 − A2
+
In order to use the initial condition, dv(0 ) / dt = 10, dif-
ferentiate the general solution and then evaluate it for time,
t = 0+ . Reversing this order and evaluating the general solu-
(
10 = A1 s1 + A2 s2 → 10 = −1.33 − A2 s1 + A2 s2 )
tion, before differentiating it, will mistakenly lead to iden-
tifying quantities as constants, which are, in fact, variables 10 = −1.33s1 − A 2 s1 + A 2 s2
that happen to have a known value at the instant, t = 0+.
10 + 1.33s1
10 = −1.33s1 + A 2 ( s2 − s1 ) → = A2
dv (t ) dA1e s2 − s1
s1t s2 t
dA 2 e dC
= + +
dt dt dt dt 0
10 + 1.33s1
A1 = −1.33 − A2 → A1 = −1.33 −
dv (t )
s1t s2 t
dA1e dA1e s2 − s1
= + A2
dt dt dt
Now, substitute for s1 and s2 symbolically as s1 , s2 = σ ± jw .
Factor the undetermined constants, A1 and A2, out of the de-
rivatives 10 + 1.33s1 10 + 1.33 (σ + jw )
A2 = → A2 =
dv (t ) de s1t de s2t s2 − s1 σ − jw − (σ + jw )
= A1 + A2
dt dt dt 10 + 1.33σ + j1.33w
A2 =
de u
du − j 2w
Recall how to differentiate an exponential = eu .
dx dx
dv (t ) ds t ds t Express A2 as a single complex number rather than as a ratio.
= A1e s1t 1 + A2 e s2t 2 j 2w
dt dt dt Multiply by the unity ratio, .
j 2w
The eigenvalues, s1 and s2, are constants. Factor them out of
the derivatives: 10 + 1.33σ + j1.33w  j 2w 
A2 =
− j 2w  j 2w 
dv (t ) dt dt
= A1e s1t s1 + A 2 e s2t s2
dt dt dt j ( 20w + 2.66wσ ) + j 2 2.66w 2
A2 =
dv (t ) − j 2 4w 2
= A1 s1e s1t + A 2 s2 e s2t
dt
2.7  Complex Numbers and Variables 73

j ( 20w + 2.66wσ ) − 2.66w 2 Having determined the undetermined coefficients and ex-
A2 = pressed them as complex exponentials, we now substitute
4w 2

them and eigenvalues into the general solution symbolically.


2.66w 2  20w 2.66wσ 
A2 = − + j + 
4w 2  4w 2 4w 2  v (t ) = A1e s1t + A 2 e s2t + 1.33

 5 + 0.67σ  v (t ) = Me − jφ e(σ + jw )t + Me jφ e(σ − jw )t + 1.33


A 2 = −0.67 − j  
 w
Distribute time t in the exponents.
A1 = −1.33 − A 2
v (t ) = Me − jφ eσ t + jw t + Me jφ eσ t − jw t + 1.33
  5 + 0.67σ  
A1 = −1.33 −  −0.67 + j  
  w  Using Eq. 2.55, we express the complex exponentials as the
products of a real exponential and a complex exponential
 5 + 0.67σ 
A1 = −0.66 − j   unit vector.
 w
v (t ) = Me jφ eσ t e jw t + Me − jφ eσ t e − jw t + 1.33
Comparing A1 and A2, we see that they are complex conju-
gates. The discrepancy between real components is a round- Factor out the product Meσ t.
ing error, due to limited numerical precision. We now substi-
tute in the values, σ = −0.5 and w = 2.96 . (
v (t ) = Meσ t e − jφ e jw t + e jφ e − jw t + 1.33 )
 5 + 0.67σ 
A1 = −0.66 − j   Now use Eq. 2.55 to combine the products of two exponen-
 w tial.
 5 + 0.67 ( −0.5) 
A1 = −0.66 − j 
 2.40  (
v (t ) = Meσ t e − jφ + jw t + e jφ − jw t + 1.33 )
A1 = −0.66 − j1.94 Rearrange the order of the terms of the exponents and factor
out the imaginary number, j.
 5 − 0.67σ 
A 2 = −0.67 + j 
 w
 (
v (t ) = Meσ t e jw t − jφ + e − jw t + jφ + 1.33 )
A 2 = −0.67 + j1.94 (
v (t ) = Meσ t e j (w t −φ ) + e j (− w t + φ ) + 1.33)
Express A1 and A2 as complex exponentials. Make a simple The exponents are equal but of opposite sign, as needed to
sketch of the complex numbers to ensure that correct angles use Euler’s sine or cosine formula. It is easier to recognize
(arguments) are calculated, since A1 and A2 are in the third that the exponents are of opposite sign, which reduces the
and second quadrants of the complex plane, respectively. risk of a sign error with the phase shift, φ, if the negative sign
of the exponent of the second exponential is factored out.
 1.94 
A2 = −A1 = π − tan −1  = 1.90 rad
 0.67  (
v (t ) = Meσ t e j (w t −φ ) + e − j (w t −φ ) + 1.33)
The magnitude (modulus) of the complex A1 and A2 is the Note that the complex exponentials are counter-rotating
following: unit vectors with the phase angle, -φ. The sum of the coun-
ter-rotating complex exponential unit vectors is in the form
A1 = A2 = 0.67 2 + 1.942 = 2.05 of Eq. 2.74. After we multiply the first term by the unity
ratio of 2/2.
thus,
A1 = 2.05e − j1.90  and  A2 = 2.05e j1.90 v (t ) =
2
Me e (
2 σ t j (w t + φ )
+ e − j (w t + φ ) + 1.33 )
The angle of A1 is the phase shift of the response sinusoid.
e j (w t + φ ) + e − j (w t + φ )
We will express A1 and A2 as complex exponentials: v (t ) = 2 Meσ t + 1.33
2
A1 ≡ Me − jφ  and  A2 = Me jφ
74 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

6 Step 7. Find the coefficients by using the initial conditions.


-0.5t Step 8. Check your solution against your expectation of the
v(t) = 4.10 e cos(2.40t-1.90)+1.33
4 response of the physical system.

4.10 e-0.5t +1.33


2 2.8.1 Example Problem One: First-Order
v(t) Step Response
0
-4.10 e-0.5t +1.33 Solve the following first-order ordinary differential equation
-2
with constant coefficients:

dv
6 + 3v = 8
-4 dt
0 2 4 6 8 10
time
We need the initial conditions to calculate the response. In
2
d v dv physical systems, we must know, directly or indirectly, how
Fig. 2.35   Plot of the solution of 8 = + + 6v with the initial con-
dt 2 dt much energy is stored in the system’s energy storage ele-
( ) = 10
dv 0+ ments, immediately prior to the application of the input. It is
( )
ditions, v 0+ = 0 and
dt the flow of energy into and out of physical systems, which
changes the values of the variables. For this problem, we will
assume that the value of the output variable, v, is zero, at the
We use Euler’s cosine formula to obtain the form of Eq. 2.70: instant after we apply the following input:

e j (w t + φ ) + e − j (w t + φ ) ( )
v 0+ = 0
v (t ) = 2 Me σt
+ 1.33
2
where t = 0+ is the instant after the input is applied at t = 0.
v (t ) = 2 Me cos (w t + φ ) + 1.33
σt

Step 1. Separate the Forcing Functions and the Output


We now, finally, substitute for σ and ω from s = σ + jw Variable. Express the equation in standard form.
= −0.5 + j 2.40 and for M and φ from Me jφ = 2.05e − j1.90: The output variable is v. The independent variable is t. The
input (or forcing function), “8,” is nonsense. Again, a forc-
v (t ) = 2 ( 2.05) e −0.5t cos ( 2.40t + ( −1.90)) + 1.33 ing function (the input) must be a function of time; it can-
not be a constant. We will presume that the input is 8us (t ),
where the Heaviside unit step function us (t ) transitions from
v (t ) = 4.10e −0.5t cos ( 2.40t − 1.90) + 1.33
zero to one, when its argument becomes positive, at the time
t we define as t = 0+ . Express the differential equation in
The result is plotted in Fig. 2.35. standard form, by clearing the coefficient from the highest
derivative of the output variable.

2.8 Solved Problems Illustrating the Method dv dv 3 8


6 + 3v = 8us (t ) → + v = us (t )
of Undetermined Coefficients dt dt 6 6
dv
We shall illustrate the Method of Undetermined Coefficients + 0.50 v = 1.33us (t )
dt
by applying it in three examples.
Method of Undetermined Coefficients
Step 1. Separate the input functions and the output variable. Step 2.  Check Units of Each Term in the Equation.
Express the equation in standard form.
Step 2. Check the nature of the differential equation.  dv 
Step 3. Check the units of each term in the equation.  dt  + [ 0.50v ] = 1.33us (t )
Step 4. Solve the homogeneous equation.
 dv 
 dt  + 0.50 [ v ] = 1.33 us (t )
Step 5. Solve the forced equation for each input function to
yield the particular solutions.
Step 6. Assemble the general solution by summing (super-
v
posing) the homogeneous and particular solutions.  t  + [v] = [ ]
2.8  Solved Problems Illustrating the Method of Undetermined Coefficients 75

The units are inconsistent. This is not a valid equation. The To illustrate, we will first use substitution of the homoge-
units of each term in the summation and on both sides of the neous solution, vh (t ) = Ae st, followed by differentiation:
equation must be identical, or we are equating apples with
dv dvh (t )
oranges. We will ignore that, for this example. In a system + 0.50v = 0 → + 0.50vh (t ) = 0
equation, the coefficients carry units from the energy storage dt dt
and dissipation element equations, yielding a valid equation. dAe st
+ 0.50 Ae st = 0
dt
Step 3.  Check the Nature of the Differential Equation.
deu du
Recall how to differentiate an exponential, = eu .
dv dx dx
+ 0.50 v = 1.33us (t )
dt d ( st )
Ae s t + 0.50 Ae s t = 0
dt
This is an ordinary differential equation, since we are only
differentiating with respect to one variable, time = t. The co-  sdt tds 
Ae s t  + st
 + 0.50 Ae = 0
efficients of each term are constants. Hence, the differential  dt dt 
equation is an ordinary differential equation with constant
coefficients. We can use the Method of Undetermined Coef- The variable s is not a function of time t. Hence,
ficients to solve it.
td ( s )
= t ·0 = 0
Step 4.  Solve the Homogeneous Equation. dt
The homogeneous equation is formed by setting the input and
side of the equation to zero.
 sdt tds 
Ae st  + st st
 + 0.50 Ae = Ae ( s·1 + t ·0) + 0.50 Ae
st
dv  dt dt
+ 0.50v = 0
dt
Ae st = 0 → s + 0.50 = 0 → s = −0.50
( s + 0.50) 
We solve the homogeneous equation by knowing the form Never
Zero
of the solution. The homogeneous solution is always of the
form Ae st . The number of terms equals the order of the dif- Now, the alternative technique of mapping d → s
ferential equation. The homogeneous solution of this first- dt
order differential equation has one term. dv
+ 0.50v = 0 → sv + 0.50v = 0 → ( s + 0.50)v = 0
dt
vh (t ) = Ae st
The trivial solution, ( s + 0.50)v = 0, is v = 0. The trivial
The characteristic value or eigenvalue, s, is the solution of solution is what gives homogeneous solution its name, “ho-
the characteristic equation. There are two methods and one mogeneous” or “self-generating,” since 0 = 0. The charac-
shortcut for forming the characteristic equation from the ho- teristic equation is a first-order polynomial in the variable, s.
mogeneous equation: It has one root, s = −0.5, which is the characteristic value, or
1. Substituting the homogeneous solution, vh (t ) = Ae st , into eigenvalue, of the system.
the homogeneous equation yields the characteristic equa-
tion of the system. ( s + 0.50)v = 0 → s + 0.50 = 0 → s = −0.50
2. An alternative method to the characteristic equation is to
perform the Laplace transformation, neglecting the initial The homogeneous solution carries the unknown, yet-to-be-
condition terms. determined coefficient, A:
3. The shortcut is a “mapping” or substitution, based on the
Laplace transformation, in which each time derivative is vh (t ) = Ae st = Ae −0.5t
replaced with the corresponding power of s.

d d2 dn Step 5.  Solve the Forced Equation for Each Forcing Func-
→s → s2 → sn tion.
dt dt 2 dt n
The trial solution for each input is a summation. The first
term of the trial solution has the same functional form as the
76 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

input. The remaining terms, if present, are derivatives of the 4


first term. The input in this problem is a Heaviside step func-
tion, Eq. 2.1. A true Heaviside step change, in which a vari-
2
able has a value at time, t = 0 − , and a different value at time,
v(t)
t = 0+ , can never be obtained in an actual physical system, v(t) = 2.66-2.66 e-0.5t
above the atomic level of quantum mechanics. However, we vh (t) 0
closely approximate step changes in many circumstances.
Consequently, we will find the Heaviside step function ex- vp (t)
tremely useful mathematically. The derivative of a step, after -2 v h(t) = -2.66 e
-0.5t
the step has transitioned or “turned on” is zero, since the step
function remains constant. The derivative, at the instant the
step turns on, is either undefined, since the change of the -4
0 2 4 6 8 10
step takes place over zero time, or defined as the impulse time
singularity function, δ (t ), which also cannot exist physically Fig. 2.36  Homogeneous, particular, and general solution of Example
above the atomic level. All physical approximations of im- dv
One, 6 + 3v = 8
pulses are “pulses” of finite duration. However, we will also dt
find the impulse function useful mathematically.
We will omit the impulse from the forced (or particular) e( −0.5)( 0 ) = e0 = 1
solution, because our solution is only valid, after the input
has been applied to the system, i.e., for time, t > 0+ . The im- 0 = A + 2.66 → A = −2.66
pulse input serves to initialize the homogeneous response.
The homogeneous response is initiated by any change of the The general solution follows and is plotted in Fig. 2.36:
input to the system and, hence, is also present in the step
response. We would keep the impulse input it were the only v(t ) = vh (t ) + v p (t ) = −2.66e −0.5t + 2.66
input term, but otherwise it is not needed. Consequently, the
forced (or particular) solution is the same form as the step The expression is a stable exponential growth function. It is
input: stable because the eigenvalue, the multiplicative constant in
the exponent, is negative. The function approaches zero, as
v p (t ) = C us (t ) = C for t > 0+ time approaches infinity. If the exponent were positive, then
the response would be unstable, and the function would ap-
where C is a constant. proach infinity, as time approached infinity.
Substitute and evaluate derivatives. We will rewrite general solutions as,

dv p (t ) dC v(t ) = 2.66(1 − e −0.5t )


+ 0.50v p (t ) = 1.33 → + 0.50C = 1.33
dt dt 0
This form shows the amplitude (or magnitude) and the nature
0.50C = 1.33 → C = 2.66 of the response more clearly. The multiplicative constant is
the magnitude of the response. You will soon recognize (or
Step 6.  Assemble the General Solution. associate) the nature of the response by the function.
The general solution is the superposition (sum) of the fol-
lowing homogeneous and particular solutions
v (t ) = 2.66
 (1 − e )−0.5t

Magnitude   
v(t ) = vh (t ) + v p (t ) Stable Exponential
Growth Function

v(t ) = Ae st + C
Note that the constants are expressed as decimal fractions.
v(t ) = Ae −0.5t + 2.66 You convey the confidence you have in a result by the num-
ber of significant figures you report. Nothing in engineering
is exact. In fact, one objective of this text is to introduce
Step 7.  Use the Initial Conditions to Find the Undetermined some of the many approximations necessary in engineering
Coefficients. modeling. Carry as many significant figures in decimals as
The initial conditions allow us to evaluate the general solu- you wish as intermediate results during a calculation, but
tion for time t = 0+ . Any base raised to the power of zero never report results to more than three significant figures,
equals one, a 0 = b 0 = 1. Hence,
2.8  Solved Problems Illustrating the Method of Undetermined Coefficients 77

unless measurements of all of the parameters, the input vari-  d 2 v   dv 


able, and the initial value of the output variable were made to  2  + 3  + [1.5v ] = [ 4us (t ) ]
unusually high precision, and you have confidence that the  dt   dt 
model is accurate to that precision. v v
 t 2  + 3  t  + 1.5 [ v ] = 4us (t ) [ ]
2.8.2 Example Problem Two: Non-Oscillatory  v  v
Second-Order Step Response  t 2  +  t  + [v] = [ ]

Solve the following second-order ordinary differential equa- These units do not match, but we will ignore this, as it is a
tion with constant co­efficients: “textbook” problem.

d  dv  Step 3.  Check the Nature of the Differential Equation.


2  + 3v = 8us (t ) − 3v
dt  dt
d 2v dv
dv(0+ ) + 3 + 1.5v = 4us (t )
+
where the initial conditions are v(0 ) = 0, and = 0. dt 2 dt
dt
Step 1. Separate the Forcing Functions and the Output This is an ordinary differential equation with constant coef-
Variable. Express the differential equation in standard form. ficients, and we can use the Method of Undetermined Coef-
First, distribute the derivative operator, d : ficients to solve it.
dt

d  dv   d 2v dv  Step 4.  Solve the Homogeneous Equation.


2  + 3v = 2  2 + 3  Form the homogeneous equation by setting the input side of
dt dt  dt dt 
the equation to zero.
 d 2v dv 
2  2 + 3  = 8us (t ) − 3v d 2v dv
 dt dt  + 3 + 1.5v = 0
dt 2 dt

Now, separate the forcing function and the output variable: Since the differential equation is of second order, the homo-
geneous solution will have two terms.
 d 2v dv 
2  2 + 3  + 3v = 8us (t )
 dt dt  vh (t ) = A1e s1 t + A2 e s2 t

Finally, write the differential equation in standard form, Form the characteristic equation, using the method of your
where the highest-order derivative of the output variable has choice. Performing the Laplace transformation (neglecting
no coefficient: the initial condition terms) yields:

d 2v dv  d 2v   dv 
dt 2
+ 3 + 1.5v = 4us (t )
dt
L  2  + L 3  + L {1.5v} = L {0}
 dt   dt 

Step 2.  Check Units of Each Term in the Equation. s 2V ( s ) + 3sV ( s ) + 1.5V ( s ) = 0

 d 2 v   dv  ( s 2 + 3s + 1.5)V ( s ) = 0 → s 2 + 3s + 1.5 = 0
 2  + 3  + [1.5v] = [4us (t )]
 dt   dt 
v Solve the second-order polynomial in s. Second-order poly-
dv ∆v dv
Recall = lim . Hence dt has units  t  . The units nomials are solved using the quadratic equation:
dt t → 0 ∆t  
 d 2v 
 2  are less apparent. They are easier to identify in this −3 ± (3) 2 − (4)(1.5)
 dt  s1 , s2 =
 d 2 v   d  dv    v  2
form:  2  =     =  2  .
 dt   dt  dt    t  s1 = −0.63 and s2 = −2.37
78 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

yielding differentiation dv(t ) / dt is performed first, then the result is


evaluated using t = 0+ :
vh (t ) = A1e −0.63t + A2 e −2.37 t
dv (t ) d d d 2.67
We will use the initial conditions last to determine A1 and A2. dt
=
dt
(
A1e −0.63t +
dt
)
A 2 e −2.37 t + (
dt 0
)
dv (t ) d −0.63t d −2.37 t
Step 5.  Solve the Forced Equation for Each Forcing Func- dt
= A1
dt
e (
+ A2
dt
e ) ( )
tion.
Again, recall that exponentials are differentiated as
d 2v dv deu du
+ 3 + 1.5v = 4us (t ) = eu :
dt 2 dt dx dx

Cause and effect requires that the response of the system fol- dv (t ) d ( −0.63t ) d ( −2.37t )
lows the application of the input. Therefore, we need only = A1e −0.63t + A2 e −2.37 t
dt dt dt
evaluate the forcing function for the condition, t ≥ 0+ . The
Heaviside unit step function transitions from zero to one, dv (t ) d (t ) d (t )
when its argument becomes positive. = A1e −0.63t ( −0.63) + A2 e −2.37 t ( −2.37 )
dt dt dt

us (t ) = 1 for t>0 dv (t )
= −0.63 A1e −0.63t − 2.37 A2 e −2.37 t
dt
d 2 v p (t ) dv p (t )
2
+3 + 1.5v p (t ) = 4 Evaluate for t = 0+ :
dt dt

Assume the constant C as the particular solution, v p (t ) = C ( ) = −0.63 A


dv 0+
e
( −0.63)(0+ )
− 2.37 A 2 e
( −2.37)(0+ )
=0
1 1 1
2
dt
d C dC 4
+3 + 1.5 C = 4 → C= = 2.67 −0.63 A1 − 2.37 A 2 = 0
dt 2 0 dt 0 1.5

Step 6.  Assemble the General Solution. We now have two equations with two unknowns:
Superimpose (sum) the homogeneous and particular solu-
tions: A1 +A2 = −2.67
−0.63 A1 − 2.37 A2 = 0
v(t ) = vh (t ) + v p (t )
We can solve this system of equations with a variety of tech-
v (t ) = A1e −0.63t + A2 e −2.37 t + 2.67
niques. Direct elimination is straightforward. Eliminate A2
by rearranging the first equation, and substituting the result
Step 7.  Use the Initial Conditions to Find the Undetermined into the second:
Coefficients.
We must have as many initial conditions as the order of the A2 = −2.67 − A1
differential equation, because we will have that many unde-
termined coefficients. Each initial condition yields an equa- −0.63 A1 − 2.37 −2.67 − A1 = 0 ( )
tion, and we need as many equations as unknowns.
dv(0+ ) Solve for A1,
The initial conditions are v(0+ ) = 0 and = 0. Set
v(0+ ) = 0, dt
−(2.37)(2.67)
A1 = = −3.64
(2.37 − 0.63)
( −0.63)(0+ ) ( −2.37)(0+ )
( )
v 0+ = A1 e 1 + A2 e 1 + 2.67 = 0
Back substitute for A2
A1 + A 2 = −2.67
A2 = −2.67 − A1 = −2.67 − ( −3.64) = 0.97
We must differentiate the general solution to use the ini-
tial condition, dv(0+ ) / dt = 0. A common error is to reverse The general solution is
the order of the operations expressed in dv(0+ ) / dt = 0. The
v (t ) = vh 1 (t ) + vh 2 (t ) + v p (t )
2.8  Solved Problems Illustrating the Method of Undetermined Coefficients 79

4 Step 2.  Check Units of Each Term in the Equation.

 d 2v   dv 
v(t) 2  2  + 0.33  + [ 0.5v ] = [1.33 us (t ) ]
 dt   dt 
vh (t) v(t) = -3.64 e -0.63t + 0.97 e-2.37t+ 2.67  v  v
 t 2  +  t  + [ v ] = [ ]
1
0
vh (t)
2 -2.37t
vh (t) = -3.64 e The units do not check. If this were a system equation, we
vp (t) 2
would not proceed, until we eliminated the inconsistency in
-2
vh (t) = -3.64 e
-0.63t the units. You can multiply apples and oranges, but you can-
1
not add them.
-4
0 2 4 6 8 10 Step 3.  Check the Nature of the Differential Equation.
time It is an ordinary differential equation with constant coeffi-
cients. We can use the Method of Undetermined Coefficients.
Fig. 2.37   Homogeneous, particular, and general solutions of Example
d 2v dv
Two, 2 + 3 + 1.5v = 4 Step 4.  Solve the Homogeneous Equation.
dt dt
Form the homogeneous equation by setting the forcing
function(s) to zero:
v (t ) = −3.64e −0.63t + 0.97e −2.37 t + 2.67 d 2v dv
+ 0.33 + 0.5v = 0
dt 2 dt
and is plotted in Fig. 2.37.
d
Create the characteristic equation by mapping → s and
factoring out the output variable: dt
2.8.3 Example Problem Three: Oscillatory
Second-Order Step Response
s 2 v + 0.33sv + 0.5v = 0 → (s 2
)
+ 0.33s + 0.5 v = 0
Solve the following second-order ordinary differential equa-
tion with constant co­efficients: s 2 + 0.33s + 0.5 = 0

Solve the characteristic equation:


d 2v dv
6 2
+ 2 + 3v = 8us (t )
dt dt −0.33 ± (0.33) 2 − (4)(0.5) s1 = −0.16 + j 0.68
s1 , s2 = →
dv(0+ ) 2 s2 = −0.16 − j 0.68
+
where the initial conditions are v(0 ) = 0 and = 0.
dt
Note that the characteristic equation has complex conjugate
Step 1.  Separate the Inputs (Forcing Functions) and Output roots. Complex conjugate roots identify an “underdamped”
Variable. system which will oscillate, when disturbed by any input.
The forcing function and the output variable terms are al- The imaginary component of the root is the frequency of the
ready separated. We will express the differential equation in oscillation in radians per second. The real component of the
standard form by clearing the highest-order derivative of its root is the decay rate of the oscillation.
coefficient: There are two forms of the homogeneous solution. The
first is the product of a real exponential and the sum of a sine
d 2 v 2 dv 3 8 and a cosine term:
+ + v = us (t )
dt 2 6 dt 6 6
vh (t ) = e −0.16t ( A1 cos(0.68t ) + A2 sin(0.68t ))
d 2v dv
2
+ 0.33 + 0.5v = 1.33 us (t )
dt dt
80 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

The second form of the homogeneous solution uses the com-


dv(0+ )
plex exponential unit vector introduced in Sect. 2.7.4: Using =0
dt
st s t
vh (t ) = A1e 1 + A2 e 2 = A1e( −0.16 + j 0.68)t + A2 e( −0.16 − j 0.68)t dv (t ) d −0.16 t
dt
=
dt
e ( (
A1 cos ( 0.68 t ) + A 2 sin ( 0.68 t ) + 2.66 ) )
The complex exponential form is easier to work with, and
will be our preferred method. However, we will demonstrate The differential operator must be distributed on the right side.
both methods, starting with the trigonometric homogeneous Although not necessary, before we do so, we will distribute
solution. the exponential onto the cosine and sine terms to make the
subsequent differentiation slightly easier.
Step 5.  Solve the Forced Equation for Each Forcing Func-
tion. dv(t ) d
= ((e −0.16t A1 cos(0.68t ) + e −0.16t A2 sin(0.68t )) + 2.66))
dt dt
d 2v dv
+ 0.33 + 0.5v = 1.33us (t )
dt 2
dt dv (t ) d −0.16 t d −0.16 t
dt
=
dt
e (
A 1 cos ( 0.68 t ) +
dt
e ) (
A 2 sin ( 0.68 t ) )
The forcing function 1.33us (t ) = 1.33 for t > 0. Hence, the d
particular solution is a constant, C. + ( 2.66)
dt 0

d 2C dC 1.33 dv(t ) d d
+ 0.33 + 0.5C = 1.33 → C= = 2.66 = A1 (e −0.16t cos(0.68t )) + A2 (e −0.16t sin(0.68t ))
dt 2 0 dt 0 0.5 dt dt dt

Step 6.  Assemble the General Solution. dv (t )


dt
(
= A1 −0.16e −0.16 t cos ( 0.68 t ) − 0.68e −0.16 t sin ( 0.68 t ) )
v(t ) = vh (t ) + v p (t )
(
+ A 2 −0.16e −0.16 t sin ( 0.68 t ) + 0.68e −0.16 t cos ( 0.68 t ) )
v(t ) = e −0.16t ( A1 cos(0.68t ) + A2 sin(0.68t )) + 2.66
Now, evaluate for t = 0+ using the initial condition
Step 7.  Use the Initial Conditions to Find the Undetermined dv(0+ )
Coefficients. = 0:
dt

0 = A1 −0.16e( ( −0.16)(0+ )
( ( )) − 0.68e(
cos ( 0.68) 0+ (
( ) sin 0.68 0+
−0.16) 0+
( )( ) ))
(
+ A 2 −0.16e
( −0.16)(0+ )
( ( ))
sin 0.68 0+ + 0.68e
( −0.16)(0+ )
(
cos 0.68 0+ ( )))

(
0 = A1 −0.16 e0 1 cos ( 0) − 0.68 e0 1 sin ( 0)
1 0
) + A ( −0.16 e
2
0
1 sin ( 0) + 0.68 e0 1 cos ( 0)
0 1
)
0 = −0.16 A1 + 0.68 A2

We have a set of two equations with two unknowns. Solving


dv(0+ ) yields the coefficients A1 and A2:
The initial conditions are v(0 ) = 0 and = 0 . Using
+

v(0+ ) = 0 dt
A1 = −2.66
( −0.16 )( 0+ )
0 = A1 + 2.66 A1 = −2.66
0=e +
( A1 cos((0.68)(0 )) + A2 sin((0.68)(0 ))) + 2.66 +
→ 0.16 →
0 = −0.16 A1 + 0.68 A 2 A2 = A1 A 2 = −0.63
0.68
0 = e 0 ( A1 cos(0) + A2 sin(0)) + 2.66
Substituting the now determined coefficients, A1 and A2,
(
0 = e 0 1 A1 cos ( 0) + A 2 sin ( 0)
1 0
) + 2.66 into the general solution yields the result which is plotted in
Fig. 2.38:
0 = A1 + 2.66

v(t ) = e −0.16t ( −2.66 cos (0.68t ) − 0.63 sin (0.68t )) + 2.66


2.8  Solved Problems Illustrating the Method of Undetermined Coefficients 81

4 y
f3(t) = f1(t)+f2(t) f 2 (t) = -0.63 sin(0.68t) 1

0.5
M x = -0.63
2
f1(t) -2 -1.5 -1 -0.5 0.5 1 1.5 2 x
-0.5

f2(t) 0 -1
β=φ
-1.5
f3(t)
-2
-2 M
-2.5
M y = -2.66
-3
f1(t) = -2.66 cos(0.68t)
-4 Fig. 2.39  M is a vector sum, M = −0.63xˆ − 2.66yˆ
0 5 10 15 20
time

Fig. 2.38   Plots of f1 (t ) = −2.66cos ( 0.68t ), f 2 (t ) = −0.63sin ( 0.68t ) , The limits of sine and cosine are negative one and one. We
and the sum f3 (t ) = f1 (t ) + f 2 (t ). The sum of sinusoids is difficult to must insert a scaling factor M with an unknown magnitude
visualize before cos( β) and sin( β) in the identity, thus,

 M sin (w t + φ )
A drawback to the trigonometric form of the homogeneous = sin ( 0.68 t ) M cos ( β ) + cos ( 0.68 t ) M sin ( β ) (2.82)
solution is that the resulting waveform, the sum of a sine and
a cosine term, is difficult to visualize. We know the ampli- which yields:
tude of the individual sine and cosine terms, but what is the
amplitude of the sum? Further, what is the phase shift of the M cos(β ) = −0.63 and  M sin(β ) = −2.66
sum?
The sum of a cosine and a sine term with the same fre- We can interpret this pair of equations as the projection of
quency and, therefore, the same angle at any time t, can be a vector M onto the horizontal and vertical axes of a real
expressed as single sinusoid at that frequency, plus a phase plane:
shift φ. The calculation of the phase shift φ is based on one of
the following four trigonometric identities, Eqs. 2.78 to 2.81. M = M cos ( β ) xˆ + M sin ( β ) yˆ → M = − 0.63xˆ − 2.66yˆ
We will use the sine relationship, Eq. 2.78.
Note x̂ and ŷ denote unit vectors in the positive x and y direc-
sin (α + β ) = sin (α ) cos ( β ) + cos (α ) sin ( β ) tions, to avoid the potential confusion that the conventional
notation î and ĵ could cause. We will plot the vector, before
Equating the identity, Eq. 2.74, with the trigonometric factor we attempt to find its angle β. We find that the vector M lies
of our result: in the third quadrant, Fig. 2.39.
The angle β is calculated indirectly, as the difference be-
sin (α ) cos ( β ) + cos (α ) sin ( β ) tween − π and the angle between M and the negative x-axis:
= −2.66 cos ( 0.68 t ) − 0.63sin ( 0.68 t )  2.66 
β = −π + tan −1  → β = −π + 1.34
 0.63 
We see that to equate α with ωt and β with φ, we must rear-
range the terms: β ≡ φ = −1.80 rad

sin (w t + φ ) = sin ( 0.68 t )( −0.63) + cos ( 0.68 t )( −2.66) We calculate the magnitude of M via Pythagorean Theorem,
thus,
When we attempt to equate the factors terms, cos( β) and
sin( β), with the corresponding constants, we discover a M = M x2 + M y2 → M = ( −0.63)2 + ( −2.66)2
problem:
M = 2.73
cos (β ) = − 0.63  and  sin (β ) ≠ − 2.66
82 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

5.5 work with the variables, s1 and s2, as symbols rather than
-0.16t values whenever possible. It saves time and reduces the pos-
v(t) = 2.73e sin(0.68t-1.80)+2.66
sibility of a transcription or sign error.
4
2.73e-0.16t+2.66 Repeating Step 7.  Use the Initial Conditions to Find the
Undetermined Coefficients.
v(t) The initial conditions are v(0+ ) = 0 and dv(0+ ) / dt = 0 .
Using v(0+ ) = 0 :
2
( s1 )( 0+ ) ( s2 )( 0+ )
v(0+ ) = A1e + A2 e + 2.66 = 0 → A1 + A2 = −2.66
-2.73e-0.16t +2.66
Using dv(0+ ) / dt = 0:

0
dv (t )
0 10
time 20 30
dt
=
d
dt
( s t
A1e 1 +
d
dt
) d
(
A 2 e 2 + ( 2.66)
s t

dt 0
)
Fig. 2.40   General solution of Example Problem Three
d 2v dv dv(0+ ) ( s )(0+ ) ( s )(0+ )
6 2 + 2 + 3v = 8us (t ) with the initial conditions, v 0+ = 0 and ( ) = s1 A1e 1 + s2 A2 e 2 = 0 → s1 A1 + s2 A2 = 0
dt dt dt
dv 0+( )=0
We have a system of two equations with two unknowns:
dt
A1 + A 2 = −2.66
We check these results by evaluating M cos (β ) = −0.63 and s1 A1 + s2 A 2 = 0
M sin (β ) = −2.66 for M = 2.73 and β = −1.80 rad:
As illustrated in Sect. 2.7.5, we will not substitute in numeri-
2.73 cos( −1.80) = −0.62 and 2.73 sin( −1.80) = −2.66 cal values for s1 and s2 until necessary. Solve for A1 and A2.
Eliminate A2:
Substituting M and φ into Eq. 2.78:
A2 = − A1 − 2.66
2.73sin ( 0.68 t − 1.80)
= sin ( 0.68 t ) 2.73cos ( −1.80) + cos ( 0.68 t ) 2.73sin ( −1.80) ( )
s1 A1 + s2 − A1 − 2.66 = 0 → s1 A1 − s2 A1 = s2 2.66

2.66 s2
We can now express the general solution in terms of sine (s 1 )
− s2 A1 = s2 2.66 → A1 =
s1 − s2
with the phase angle φ, Fig. 2.40:

v(t ) = 2.73e −0.16t sin(0.68t − 1.80) + 2.66 Solve for A2:


2.66 s2
We now illustrate the alternative solution, in which the oscil- A2 = − A1 − 2.66 → A2 = − − 2.66
lation of the homogeneous solution is represented by com- s1 − s2
plex exponentials. We return to Step 6, and now use this ex-
pression, Substitute for s1 and s2 symbolically, where s1 = σ + jw and
s2 = σ − jw
st s t
vh (t ) = A1e 1 + A2 e 2 = A1e( −0.16 + j 0.68)t + A2 e( −0.16 − j 0.68)t

Repeating Step 6.  Assemble the General Solution. 2.66 s2 2.66 (σ − jw )


A1 = → A1 =
s1 − s2 σ + jw − (σ − jw )
v (t ) = vh (t ) + v p (t ) v (t ) = A1e 1 + A 2 e 2 + 2.66
s t s t
→ 2.66 (σ − jw )
A1 =
j 2w
v (t ) = A1e(−0.16 + j 0.68)t + A 2 e(−0.16 − j 0.68)t + 2.66

2.66 s2 2.66 (σ − jw )
Note how much more cumbersome is the general solution, A2 = − − 2.66 → A2 = − − 2.66
s1 − s2 σ + jw − (σ − jw )
when the values of s1 and s2 are substituted in. It is best to
2.8  Solved Problems Illustrating the Method of Undetermined Coefficients 83

2.66 (σ − jw ) Im
A2 = − − 2.6 A2 = -1.33 + j0.31
j 2w
j0.32
φ2
A1 and A2 contain ratios of a complex number over an imagi-
nary number. Both coefficients should be expressed as a α
single complex number. Multiply by the unity ratio j/j to -1.32 Re
eliminate the imaginary number in the denominator. Starting φ1
with A1: -j0.32
A 1 = -1.33 - j0.31
A1 =
2.66 (σ − jw )  j 
→ A1 =
(
2.66 jσ − j 2w )
j 2w  j  j 2 2w Fig. 2.41   Complex conjugate coefficients, A1 and A2.

2.66 ( jσ − ( −w )) 2.66 ( jσ + w )
A1 = → A1 = Calculate the magnitude (modulus) of the complex conju-
−2w −2w
gate coefficients, A1 and A2, by the Pythagorean Theorem.
Angles φ1 and φ2 should be calculated indirectly, to avoid the
j 2.66σ 2.66w 1.33σ
A1 = + → A1 = −1.33 − j error most calculators make in evaluating the inverse tangent
−2w −2w w of vectors in the second and third quadrants. Calculate angle
α using the inverse tangent and positive real values for the
Continuing with A2: lengths of the adjacent and opposite sides of the right tri-
angle, which includes α:
2.66 (σ − jw )  j 
A2 = −  j  − 2.6
j 2w  0.32 
φ2 = π −α → φ 2 = π − tan −1 
 1.33 
A2 = −
(
2.66 jσ − j 2w ) − 2.6
j 2w
2 φ 2 = π − 0.24 = 2.91 rad

 0.32 
2.66 ( jσ − ( −w )) φ 1 = −π + α → φ 1 = π + tan −1 
2.66 ( jσ + w )  1.33 
A2 = − − 2.6 → A2 = − 2.6
−2w 2w
φ 2 = −π + 0.24 = −2.91 rad
2.66σ 2.66 w 1.33σ
A2 = j + − 2.66 → A 2 = −1.33 + j
2w 2w w A1 = A2 = ( −1.33)2 + (0.31)2 = 1.37

Comparing A1 and A2, we see they are complex conjugates: Express A1 and A2 as complex exponentials:
j A1 j A2
1.33σ 1.33σ A1 = A1 e = 1.37e − j 2.91 and A2 = A2 e = 1.37e j 2.91
A1 = −1.33 − j A2 = −1.33 + j
w w
Substitute into the homogeneous solution the complex expo-
Complex conjugates have equal magnitudes (moduli) nential form of A1 and A2:
and equal but opposite angles (arguments), Fig. 2.41.
Substitute in numerical values for σ and ω where vh (t ) = A1e( −0.16 + j 0.68)t + A2 e( −0.16 − j 0.68)t
s1 , s2 = σ ± jw = −0.16 ± j 0.68,
vh (t ) = 1.37e − j 2.91e( −0.16 + j 0.68)t + 1.37e j 2.91e( −0.16 − j 0.68)t
1.33 ( −0.16)
A1 = −1.33 − j = −1.33 − j 0.31
0.68 Distribute the time variable, t, and then use the property of
exponentials, e a + b = e a eb , Eq. 2.51, to express the exponen-
1.33 ( −0.16) tials with exponents that are sums, as the product of expo-
A2 = −1.33 + j = −1.33 + j 0.31
0.68 nentials:

vh (t ) = 1.37e − j 2.91e −0.16t + j 0.68t + 1.37e j 2.91e −0.16t − j 0.68t


84 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

vh (t ) = 1.37e − j 2.91e −0.16t e j 0.68t + 1.37e j 2.91e −0.16t e − j 0.68t 2.9 Eigenvalues and Response
Characterization
Factor out the constant and the purely real exponential:
As we have seen, the homogeneous solution of an ordinary
(
vh (t ) = 1.37e −0.16t e − j 2.91e j 0.68t + e j 2.91e − j 0.68t ) differential equation with constant coefficients is the sum
of exponentials of the form Aest, where the variables, s, are
Use the property, e a + b = e a eb , to express the product of two the roots of a differential equation’s characteristic equation.
exponentials as a single exponential: These roots are the differential equation’s characteristic val-
ues or eigenvalues. The eigenvalues are the coefficients of
(
vh (t ) = 1.37e −0.16t e − j 2.91+ j 0.68t + e j 2.91− j 0.68t ) the variable time t in the exponentials of the homogeneous
solution. They determine the temporal or time scaling of the
Factor the imaginary number j out of the exponents: homogeneous solution. Eigenvalues are either purely real
numbers, e.g., s = σ , or complex conjugates, s1 , s2 = σ ± jw .
(
vh (t ) = 1.37e −0.16t e j (−2.91+ 0.68t ) + e j (2.91− 0.68t ) ) Complex conjugate eigenvalues indicate that the homoge-
neous response is underdamped and oscillatory. Oscillatory
We will rearrange frequency and phase angle terms in the ex- homogeneous responses have two time scales, the period of
ponents into the conventional order, w t + φ , where w = 0.68 the oscillation and the decay rate of the oscillation.
and φ = −2.91: The time scaling of step responses reveal the eigenval-
ues of the system. Consequently, the time scaling of a step
(
vh (t ) = 1.37e −0.16t e j (0.68t − 2.91) + e j (−0.68t + 2.91) ) response obtained from a test of an existing physical system
is used to develop a mathematical model of the system. This
Recall that any quantity multiplied by j in the exponent of an process is known as system “identification.” A second use
exponential is its angle. The angles are equal but of opposite of the eigenvalues time scaling is to establish the duration
sign. Factoring out the negative sign from the second expo- of the response. Further, when plotting a response function,
nential makes this easier to see: eigenvalues time scaling can be used to establish the “time
step,” which is the time interval between successive calcula-
(
vh (t ) = 1.37e −0.16t e j (0.68t − 2.91) + e − j (0.68t − 2.91) ) tions.

When the sum of the exponentials, e j ( 0.68t − 2.91) + e − j ( 0.68t − 2.91),


is compared with Euler’s cosine and sine formulas: 2.9.1  First-Order Step Responses

e jθ + e − jθ e jθ − e − jθ A purely real eigenvalue, or the real component of a complex


cos(θ ) = Eq. 2.53 sin(θ ) = Eq. 2.54
2 j2 eigenvalue in an exponent, yields a real exponential which,
as a factor, scales the amplitude of the output variable. We
we see that the cosine formula can be used, if the exponential have seen that a real eigenvalue or the real component of a
term is multiplied by the unity ratio 2/2: complex eigenvalue must be negative for the response func-
tion to decrease or decay with time and be “stable.” Con-
vh (t ) = 1.37e −0.16t
2
e (
2 j (0.68t − 2.91)
+ e − j (0.68t − 2.91) ) versely, if a real eigenvalue or the real component of a com-
plex conjugate eigenvalue is positive, the real exponential
e j (0.68t − 2.91) + e − j (0.68t − 2.91) formed by the eigenvalue grows with time, and the response
vh (t ) = 2.74e −0.16t is unstable.
2
The characteristic time of a real exponential is the inverse
vh (t ) = 2.74e −0.16t cos ( 0.68t − 2.91) of the absolute value of a purely real eigenvalue or the real
component of a complex eigenvalue, and is known as the
Assemble the general solution: “time constant.” The time constant’s symbol is tau, τ, Greek
for t, Eq. 2.83:
v (t ) = vh (t ) + v p (t ) 1
 τ≡ (2.83)
σ
v (t ) = 2.74e −0.16 t
cos ( 0.68t − 2.91) + 2.66
The magnitude or absolute value removes the negative sign
This result differs from our previous by a rounding error of of the eigenvalue or its real component, σ of stable systems.
0.01.
2.9  Eigenvalues and Response Characterization 85

Fig. 2.42   a Unit exponential


a b
decay with time scaled in 1
1
units of time constants τ.
b Unit stable exponential _t 63.2% _t
growth with time axis scaled e- τ 86.5%
95.0% 98.2% 99.3%
1- e- τ 95.0% 98.2% 99.3%
86.5%
in time constants τ. The 63.2%
responses decay 63.2 % the
remaining difference from 0 0
their current and steady- 0 τ 2τ 3τ 4τ 5τ 0 τ 2τ 3τ 4τ 5τ
state values during each time time
time interval, equal to one
time constant

A negative sign must be added to the exponent, when the   −t



exponential is written in terms of the time constant. ( )
f 2 (t ) = ∆ f 2 1 − eσ 2t + C2 → f 2 ( t ) = ∆ f 2  1 − e τ 2  + C2
 
−t
 eσ t = e τ for σ <0 (2.84) (2.87)

A property of exponents implied by Eq. 2.56 is that a nega- The dynamic range Δf is the absolute value of the difference
tive exponent expresses an inverse fractional relationship: between the initial and final values of the step response. We
will examine the step responses of first-order systems in de-
 1 tail in Chap. 3.
e− a = (2.85)
ea The real exponential of a first-order system’s step re-
t t sponse is characterized by its time constant τ. A simple
t −  1 − 
 
If a =   , then e  τ  =  t  . Evaluating e τ at the limits method to determine the time constant τ from data is to cal-
τ  
e τ  culate either 63.2 % of the change in the output variable from
of t = 0 and t = ∞ yields its initial to final value, or the corresponding 36.8 % of the
change remaining, and then to find the time corresponding
 0 to that value in the data, interpolating if necessary. Estimat-
− 
τ 1 1 1
e = = = = 1 for t = 0 ing a time constant from the duration of a first-order step
 0
  e0 1
e τ transient is usually inaccurate, because noise in the data and
the quantification imparted by digital instrumentation tend
and to obscure when the signal reaches steady-state defined as
either 98.2 %, 4τ, 99.3 %, or 5τ, of the step change.
 ∞
− 
τ 1 1 1
e = = = = 0 for t = ∞
 ∞
 
τ
e∞ ∞ Example One Determine the response function which
e describes the data plotted in Fig. 2.43.
t
− 
The response is a decay in the form of Eq. 2.86 with
 
Since e τ decreases from one to zero as time proceeds from C1 = 0, since the response decays to zero. The dynamic range
zero to infinity, it is a “decay” function. Although a decay is the initial value minus the steady-state value.
function requires infinite time to reach zero, it approaches
t
− 
∆ f1 = f1 ( 0+ ) − f1 ( ss ) → ∆ f1 = 4 − 0 = 4
τ
zero quickly. Evaluating e for time expressed in mul-
tiples of the time constant τ, Fig. 2.42 serves as illustration: The response decays 63.2 % of the dynamic range in the
A unit exponential decay, Fig. 2.42a, and a unit stable ex- first time constant, leaving 36.8 % of the dynamic range
ponential growth, Fig. 2.42b, range from zero to one. In a remaining.
step response, such as that shown in Fig. 2.36, these func-
tions are scaled in amplitude by a coefficient, the dynamic 0.632 ∆ f1 = ( 0.632)( 4) = 2.53
range, Δf, and can be offset from zero by the addition of a
constant, Eqs. 2.86 and 2.87. and
−t 0.368∆ f1 = ( 0.368)( 4) = 1.47
 f1 (t ) = ∆ f1eσ1t + C1 → f1 (t ) = ∆ f1e τ1 + C1 (2.86)
86 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Fig. 2.43   a Step response data


for an exponential decay. b The
a 4
b 4
time of the datum equal to 63.2 %
of the initial value and 36.8 % of 3
3 0.632 f1(0+)
the remaining change to the final
value equals the time constant τ,
τ1 = 0.6 sec f1(t) 2 f1(t) 2

1 1
0.368 f1(0+)

0 0
0 1 2 3 4 0 τ 1 2 3 4
t, sec t, sec

Fig. 2.44   a Step response data


for a stable exponential growth,
a 20
b
20
b The time of the datum equal to
63.2 % of the dynamic range plus 0.368 f2(ss)
the initial value, and the steady- 15 15
state value minus 36.8 % of the
dynamic range equals the time f2(t) 10 f2(t) 10
constant τ, τ2= 1.5 sec
0.632 f2(ss)
5 5

0 0
0 2 4 6 8 10 0 2 4 6 8 10
t, sec t, sec

The time at which the response equals 1.47 is the time con-  −t
  −t

f 2 ( t ) = ∆ f 2  1 − e  + C2
τ2
→ f 2 (t ) = 20 1 − e1.5 
stant, τ1 = 0.6 sec. The response function is    
−t −t
f1 (t ) = ∆ f1e + C1
τ1
→ f1 (t ) = 4 e 0.6 2.9.2 Non-Oscillatory Second-Order Step
Responses
Example Two Determine the response function which
describes the data plotted in Fig. 2.44. A non-oscillatory second-order step response, Eq. 2.88 and
The response is a stable exponential growth of the form Fig. 2.37, is also known as an “overdamped” second-order
of Eq. 2.87 with C2 = 0 since the response grows from zero. step response. An overdamped second-order step response
The dynamic range is the steady-state value minus the initial occurs when the two eigenvalues are both purely real, s1 = σ 1
value. and s2 = σ 2 . This response is difficult to characterize, be-
cause its features are less distinctive than an oscillatory step
∆ f 2 = f 2 ( ss ) − f 2 ( 0+ ) → ∆ f 2 = 20 − 0 = 20 response.

The response grows 63.2 % of the dynamic range in the 


M3  1 
first time constant, leaving 36.8 % of the dynamic range f 3 (t ) = 1 +
σ 1σ 2  σ 2 − σ 1
(
σ 1eσ 2t − σ 2 eσ1t  ) (2.88)
remaining. 

0.632 ∆ f 2 = ( 0.632)( 20) = 12.0 The eigenvalues or the corresponding time constants of the
system are established by curve fitting the step response
and
function f3( t), Eq. 2.88 to the data. If an iterative minimi-
0.368∆ f 2 = ( 0.368)( 20) = 8.0 zation routine is available, then any reasonable two initial
guesses for eigenvalues can be used. If the iteration must
The time at which the response equals 12.0 is the time con- be done manually, then it is worth the time to roughly esti-
stant, τ2 = 1.5 sec. The response function is mate initial values of the eigenvalues. The larger of the two
2.9  Eigenvalues and Response Characterization 87

Fig. 2.45  a Overdamped


second-order step response. a 30
b 30
b Quasi-linear portion of the f3(ss)=27
initial portion of response
extended to intersect the time 20 20
axis. The time interval between
that intersection and the time f3(t) f3(t)
corresponding to 63.2 % of
10 10
the steady-state value is the 0.632 f3(ss)
initial estimate of the larger time
constant
0 0
0 2.5 5 7.5 10 0 2.5 5 7.5 10
∆t=τ
t, sec t, sec

Fig. 2.46  a First iteration fitting


curve to overdamped second-or-
a b
30 30
der step response data. b Second
iteration reducing the value of
eigenvalue σ2 to improve the
fit in the initial portion of the f3(t) 20 f3(t) 20
curve. The values used to create
the “data” were σ 1 = −0.667, fest (t) fest (t)
10
σ1 = -0.549 10
σ1 = -0.549
σ 2 = −1.667 , and M 3 = 30
σ2 = -5.49 σ2 = -2.50
M 3 = 81.4 M 3 = 37.1
0 0
0 2.5 5 7.5 10 0 2.5 5 7.5 10
t, sec t, sec

time constants (the smaller eigenvalue) is estimated using The overdamped step response formula, Eq. 2.88, is written
the method described above, as if the response were a first- in terms of eigenvalues.
order step response. The quasi-linear portion of the response
curve is extended to intersect with the time axis, eliminating −1 −1 −1 −1
σ1 = = − 0.549 and σ 2 = = = −5.49
the initial curve due to the second exponential term. The time τ1 1.82 τ 2 0.182
constant of second real exponential will be smaller by some
unknown magnitude. However, the lower limit of its size is The steady-state value of the overdamped unit step func­
approximately one tenth of the larger time constant, if it con- tion is
tributes a recognizable reverse curve in the initial portion of
the response. The closer the eigenvalues or time constants
are in value, the greater the symmetry of the reverse curve of
f 3 ( ss ) =
M3 
1 +
σ 1σ 2 
1
σ 2 − σ1
(σ e 1
σ 2∞
0 − σ 2 eσ ∞
1
0 )
the overdamped step response.
M3
f 3 ( ss ) =
Example  Determine the response function which describes σ 1σ 2
the data plotted in Fig. 2.45.
The larger time constant is estimated by extending the The steady-state value of the step response data is 27. The
quasi-linear portion of the response to intersect the time constant M3 can be estimated, using the observed steady-
axis, Fig. 2.45b. The time at which the response has grown to state value and the estimates of the eigenvalues.
63.2 % of the steady-state value is then determined. The dif-
ference between these times is used as the initial estimate of
M 3 = σ 1σ 2 f 3 ( ss ) → M 3 = ( −0.549)( −5.49)( 27 ) = 81.4
the larger time constant. The smaller time constant is likely
no smaller than one-tenth of the larger time constant if the
reverse curve is visible. The result, Fig. 2.46a, has its greatest error early in the re-
sponse, where the larger magnitude eigenvalue σ2 has its ef-
1.82 sec fect. Reducing the magnitude to σ 2 = −2.50 improved the
τ1 ≈ 1.82 sec and τ 2 ≥ = 0.182 sec
10 curve fit, Fig. 2.46b. A few more iterations would be suf-
ficient for closely matching the data.
88 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

If the eigenvalues are equal, in other words, the character- 1.75


T
istic equation has repeated roots, then the system is known as Period
“critically damped.” If the amount of damping is decreased
or energy storage in the system is increased, then the step 1.00
response will be underdamped and overshoot its final value. fstep(t)
The term, “critical,” is a misnomer, since the situation is not
fimpulse(t)
critical, in the sense of producing an unstable response. It is,
in fact, difficult to discern from a step response, whether a 0.00
system slightly underdamped, critically damped, or slightly T
overdamped. Period
-0.75
0 T 2T 3T 4T 5T
time in periods T
2.9.3  Oscillatory Second-Order Step Responses
Fig. 2.47   Oscillatory second-order step and impulse responses. The
An oscillatory second-order step response, Figs. 2.31, 2.32, true period is measured from crossing to crossing of the steady-state
2.35, and 2.40, occurs, when the eigenvalues are a complex value, as shown on the impulse response trace. The usual method of
measuring peak to peak, shown on the step response trace, is less error
conjugate pair, s1 , s2 = σ ± jw , and reflects the physical situ- prone due to the ease in identifying peaks
ation of internal energy transfer between two independent
storage modes persisting long enough for the step response
to overshoot its final value at least once. An oscillatory peak of the response. The peak-to-peak time is not equal to
second-order step response may have a steady-state or final the period due to the decay of the oscillation, but the inher-
value which is either non-zero or zero. The latter is known as ent error is less than the typical error finding the true period
an impulse response and occurs when the only term input is using the steady-state crossing.
differentiated. The two general forms are: Instruments typically report frequency in cycles per sec-
ond or hertz, which is given the symbol f. A common and
 substantial error is to neglect to convert from a frequency
f step (t ) = Meσ t sin (w t + φ ) + C (2.89)
in hertz to the angular frequency in radians per second be-
f impulse (t ) = Meσ t sin (w t + φ ) (2.90) fore performing system dynamics calculations. All system
 dynamics calculations must be performed using angular fre-
quency in radians per second.
The real component of the eigenvalue σ is the coefficient The real component of the eigenvalue is determined from
in the exponent of the real exponential decay. The imagi- two values of the response. In principle, any two values can
nary component of the eigenvalue is the angular frequency be used, other than the crossing of the steady-state value,
of the oscillation in radians per second with the symbol, ω. where the sinusoid equals zero. In practice, values at two
An oscillatory homogeneous response has two characteristic peaks or two troughs are used to eliminate the sinusoid
times, the period of the oscillation T and the time constant of from the computation, since sin(w t + φ ) ≈ 1 at peaks, and
1 sin(w t + φ ) ≈ −1 at troughs. The sinusoids are only approxi-
the decay envelope, τ = .
σ mately equal to one and negative one due to the slight shift
The period T of an oscillation is the duration of a com- of the peaks and troughs created by the decay envelope but
plete cycle and is inversely proportional to the angular fre- that error is negligible.
quency,
 Eq. 2.88. The differences between peak values and steady-state
2π rad are identified as a and b, between trough values and steady-
T ≡ period = (2.91) state as c and d in Fig. 2.48. The time between the peaks or
rad
w troughs is approximately equal to the period T. In general,
sec
the coefficient Mis unknown until the real coefficient of the
Equation 2.91 is easy to remember if one equates a cycle eigenvalue is established. Using the value a at time t1 and b
with a circle. There are 2π rad in a circle and a cycle. at time t1 + T with sin(w t + φ ) ≈ 1
Measurement of the period T is illustrated in Fig. 2.47.
The period T should be measured as the time between suc- a = Meσ t1 and b = Meσ (t1 +T )
cessive crossings of the steady-state value, as shown for the
impulse response. Unfortunately, crossing of the steady-state Hence,
value can be difficult to determine on an oscilloscope. The
less error-prone method is to measure the time from peak to Meσ t1 − Meσ (t1 +T ) = a − b → M (eσ t1 − eσ (t1 +T ) ) = a − b
2.9  Eigenvalues and Response Characterization 89

1.75 should be calculated as the steady-state value minus the


trough value so that difference is positive.
a
Applying Eq. 2.92 to step response in Fig. 2.48 where
b
1.00 a = 0.62, b = 0.24, and T = 1.
t1
fstep(t)
 0.62 − 0.24 
ln 1 − 
fimpulse(t)  0.62 
t2 σ= = −0.95
0.00 1
d
c
2.9.4  Time Step for Response Calculations
-0.75
0 T 2T 3T 4T 5T
The real components of the eigenvalues of a system determine
time in periods T the rates of exponential decays in the homogeneous response.
Fig. 2.48   Oscillatory second-order unit step and impulse responses.
If a system has multiple eigenvalues, then the smallest mag-
The difference between peak and steady-state values, a and b, or the nitude real component σ min corresponds to the largest time
difference between steady-state and trough values, c and d, are used to constant τmax, due to their inverse relationship, and establishes
calculate σ, the real component of the complex conjugate eigenvalues the duration of the transient period of the response. The mag-
nitude or absolute value of the real component is used because
σ is negative. A response plot should extend into the beginning
a −b a −b of the steady-state response of the system. A first estimate for
eσ t1 − eσ (t1 +T ) = → eσ (t1 +T ) = eσ t1 −
M M the duration of a plot is tmax ≈ 7τ .
A “time step,” Δt or T, is the time between consecutive
Divide both sides by eσ t1 and use the property of exponen- calculations of a variable’s response, its time history. The
ea use of Δt for the difference between the times of consecu-
tials b = e a − b
e tive calculations is self-explanatory. The symbol T is also
used, both for the practical reason that MATLAB and other
e (1 )
σ t +T
eσ t1 a −b a −b programming languages are restricted to the Roman charac-
e ( 1 ) 1 = 1−
σ t + T −σ t
= − →
e σ t1
eσ t1
Me σ t1
Meσ t1 ters of a computer keyboard and because it is the symbol for
the period of a periodic function. The calculation is periodic
a −b since it is repeated at a constant time interval.
eσ T = 1 −
Meσ t1 When we repeatedly evaluate a response function for suc-
cessive values of time to create a plot, we want the time step
Substitute using a = Meσ t1 to be small enough to create a faithful representation of the
function. If the time steps are too large, then a smooth func-
a −b tion becomes broken into a piece-wise continuous function
eσ T = 1 −
a comprising line segments. If the time steps are too small,
then we create many more data than we need and take exces-
Take the natural logarithm of both sides. sive time doing so. Using 1/200 of the smallest characteristic
time of the system, either a time constant τ or a period T, will
yield a good plot.
 a − b  a − b
ln(eσ T ) = ln 1 −  → σ T = ln 1 −  We shall see in Chap. 8 that the magnitude of the time
 a  a 
step is critical when we use the Euler and Runge–Kutta al-
gorithms to solve a set of first-order differential equations.
This will yield These algorithms are “recursive,” meaning that the next
value is calculated using previous values. Specifically, the
  a − b Euler and Runge–Kutta algorithms estimate the first deriva-
ln 1 − 
 a  tive of output variable, multiply it by the time step to cal-
σ= (2.92)
T culate the change in the output variable, and then add that
change to the present value to estimate the next value of
If the peak or trough values are more than one cycle apart, the output variable. The estimate of the first derivative of
multiply the period T in the denominator of Eq. 2.92 by the the output variable improves as the time step is decreased.
number of cycles separating the values. Differences between If the time step between calculations is too large, then the
steady-state and trough values, such as c and d in Fig. 2.48,
90 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

estimate of the derivative can be far enough in error to create The Laplace transformation is defined as the integral:
a “numerical instability,” so called because numerical insta-

bilities yield the same shaped plots as physical instabilities. 
They blow up. When we use “numerical methods” to solve
L { f (t )} = F ( s ) = ∫ f (t ) e
−∞
− st
dt (2.93)
for the response of a dynamic system, the time step Δt will
be a fraction of the smallest characteristic time of the system. The function of time to be transformed, f (t ), is multiplied
by the complex exponential, e − st , and the interval is evalu-
ated from t = − ∞ to t = + ∞. All functions f (t ) of practical
2.10 Laplace Transformation and Transfer interest satisfy the necessary condition for transformability.
Functions We have omitted the dummy variable τ for time. The use of τ
as the time constant of first-order systems can lead to confu-
2.10.1  Laplace Transformation sion, when τ is also used as a dummy variable. It is conven-
tion to use lower case for time-domain variables and capital-
A transformation is an operation which changes the inde- ize Laplace-domain variables, although it is frequently not
pendent variable of a function. The Laplace transformation possible. A case in point is our use of capital F for force in
changes a function of time to a function of the Laplace vari- the time domain. The variable transformed to the Laplace-
able, s. The independent variable defines the domain of a domain is F( s), where the functional notation ( s), which is
function. Consequently, we speak of the Laplace transform often omitted in the time-domain, i.e., F( t), is essential to
as transforming a function from the time-domain to the indicate that the variable is in the Laplace-domain.
“Laplace-domain”, in other words, to the Laplace variable, We customarily choose to define time, t = 0, when we
s. The importance of the Laplace transformation is that time- apply an input to a system. The input function is zero for
domain differential equations are transformed to Laplace- time, t < 0. Consequently, there is no need to extend the lower
domain algebraic equations. limit of integration to t = − ∞. It is more appropriate to use a
The Laplace transformation was introduced in the context “one-sided” Laplace transformation where the lower limit of
of forming the characteristic function of a differential equa- integration is t = 0:
tion in Example Two of Sect. 2.5. We introduced the Laplace

transformation by stating without proof, that it is a linear op- 
erator and, as such, the Laplace transform can be distributed
L { f (t )} = F ( s ) = ∫ f (t ) e
0
− st
dt (2.94)
onto each term of a sum, and constants can be factored out of
the transform. We also stated that the Laplace transformation We will establish that the Laplace transformation is a linear
can be viewed as an operator which transforms the differ- operator by confirming it has the two essential properties of
entiation with respect to time operator, d/dt, to the Laplace linear operators. First, doubling the input f( t) to the Laplace
variable, s, when the initial condition terms, which apply to transformation doubles the output F( s):
the transformation of the derivatives with respect to time,
∞ ∞
Eqs. 2.20 and 2.21, are neglected. Further, we noted that an
unknown function of time, x( t), transforms into an unknown L {2 f (t )} = ∫ 2 f (t ) e− st dt = 2∫ f (t ) e− st dt = 2 F ( s )
0 0
function of the Laplace variable, s, which is X( s).
Our application of the Laplace transformation will be Second, the Laplace transformation can be distributed onto a
to transform an input function from the time-domain to the sum of transformable functions:
Laplace-domain, output functions from the Laplace-domain

to the time-domain, and differential equations both ways, as
needed. We will manipulate the algebraic equation of a trans- L { f (t ) + g (t )} = ∫ ( f (t ) + g (t )) e
0
− st
dt
formed differential system equation into a ratio known as a
∞ ∞
“transfer function,” which has the unique property of repre-
senting a system’s dynamic relationship in the form of a mul-
L { f (t ) + g (t )} = ∫ f (t ) e
0
− st
dt + ∫ g (t ) e − st dt
0
tiplicative operator. Multiplying a system’s transfer function
by the Laplace transform of an input function yields the out- L { f (t ) + g (t )} = F ( s ) + G ( s )
put variable’s response function in the Laplace-domain. Per-
forming the inverse Laplace transformation yields the output Therefore, the Laplace transformation is a linear operator.
variable’s time-domain response. We will discuss transfer We will evaluate three Laplace transform integrals to re-
functions in the next section, but we must first develop the veal a few of the properties of the Laplace transform, and
Laplace transformation of functions of time. to illustrate the construction of a Laplace transform table of
time-domain and Laplace-domain pairs. We will not evaluate
2.10  Laplace Transformation and Transfer Functions 91

the Laplace transform integral in practice. We will use a table We seek the integral, ∫e dt. Multiplying both sides of
− st

Laplace transform pairs instead, Table 2.3. our result by − 1s :

2.10.1.1 Example One: Laplace Transformation


1 1 1
of the Heaviside Unit Step Function − ( − s ) ∫ e − st dt = − e− st → ∫e
− st
dt = − e − st
f ( t ) = us ( t ) s s s


We have what we need.
L {u (t )} = U ( s ) = ∫ u (t ) e
s s
0
s
− st
dt
Returning to the Laplace transformation of the Heaviside
unit step function, we evaluate the integral:
Recall the definition of the Heaviside unit step function:

us (t − ta ) = 0 for t < ta ∞ ∞

L {u (t )} = ∫ u (t ) e
s
0
s
− st
dt = ∫ e − st dt
0
us (t − ta ) = 1 for t ≥ ta

e − st e − s⋅∞  e − s⋅0 
In the present case, ta = 0. Hence, L {us (t )} = − s
→ L {us (t )} = − s
−−
 s 
0

us (t ) = 1 for t ≥ 0
1
Using the properties of exponentials, e − a = and e0 = 1,
The unit step function is constant and equal to unity over the ea
yields the Laplace transformation of the Heaviside unit step
entire integral. Consequently, we can evaluate the unit step function.
function first and then integrate:
1  1
∞ ∞ ∞
L {u (t )} = − se( )( )
s −− 
 s
∫ u (t ) e
s ∞
s
− st
dt = ∫ 1e dt = ∫ e dt
− st − st
0
0 0 0
1
Integration and differentiation are inverse operations. If one us (t ) =
(2.95)
s
L{ }
remembers how to differentiate a function, the formula for
integration can be constructed by working backward. For ex-
ample, recall how to differentiate an exponential: An alternative presentation of this result is as a Laplace
transform pair:
du deu du deu
eu ∫e dx = ∫
u
= ∴ dx = eu  f (t ) = us (t )
dx dx dx dx 
 1
In our case, t = x and − st = u, where s is an unknown com-  F ( s ) = s
plex variable.

du d ( − st ) 2.10.1.2 Example Two: Laplace Transformation


∫e ∫e
u
dx = eu → − st
dt = e − st
dx dt of an Exponential Decay f ( t ) = e − st
The Laplace transformation integral of an exponential
 − sdt −tds 
∫e decay is:
− st − st
 +  dt = e
dt dt  ∞

L {e } = ∫ e
− at − at − st
e dt
Inspecting the left side, dt equals one. The complex vari- 0
dt
able, s, is not a function of time. Hence, ds equals zero. Begin the evaluation of the integral by combining the prod-
dt uct of exponentials using the property, e a eb = e a + b .
 dt ds 
∫e  − s dt − t dt  dt = ∫ − se dt = e
− st − st − st ∞ ∞

1 0 L{ } e − at = ∫ e − at e − st dt
0
→ L{ } e − at = ∫ e − (a + s )t dt
0

− s ∫ e dt = e
− st − st
92 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

We now make use of the integration formula we derived for The Laplace transformation of a function’s second deriva-
the integration of the Heaviside step function, tive with respect to time is here:

e − st 
∫e
− st
dt = −  d 2 f (t )  ( )
df 0+
s L  2  = s 2
F ( s ) − ( )
sf 0 +

dt
(2.98)
 dt 
replacing the variable, s, with the sum, a + s.
We will neglect the initial conditions terms, when performing
− ( a + s )t e − ( a + s )t the Laplace transformation on a differential system equation
∫e dt = −
a+s to form a transfer function, Sect. 2.10.4. In those cases, the La-
place transformation of a function differentiated with respect
We can now evaluate the integral: to time is
t =∞ d n f (t )
e − ( a + s )t

(2.99) = s n F (s)
L {e } = ∫ e
− at

0
− ( a + s )t
dt = −
a+s
dt n
t =0

− ( a + s )⋅∞
e  e − (a + s )⋅0 
L {e } = −  a + s
− at

a + s 
2.10.2  The Inverse Laplace Transformation

The inverse Laplace transformation is the process of deriving


 e − (a + s )⋅∞ e − (a + s )⋅0 1  the time-domain function, which corresponds to a Laplace-
L {e } − at
= −
 a + s
0

a + s 
 domain function. The analytical method involves integration
in the complex plane, which is graduate-level mathematics
1 for most mechanical engineers. Fortunately, there is no need
L {e } = a + s − at
to evaluate a complex integral, if the Laplace transform of
f (t ) has been evaluated yielding the Laplace transform pair,
Rewrite the denominator in standard polynomial notation which includes the relevant F ( s ). If we have established the
with the constant a last to yield the result in this form: transformation from the time-domain to the Laplace-domain,
we know the corresponding reverse or inverse transforma-
 1 tion.
L {e } = s + a − at
(2.96)

or, as a Laplace transform pair,


L { f (t )} = F ( s ) → L {L { f (t )}} = L {F ( s )}
−1 −1

 f (t ) = e − at

 f (t ) = L {F ( s )}
−1
(2.100)
 1
F (s) = We will refer to a time-domain function and its correspond-
 s+a
ing Laplace-domain function as a Laplace transform pair.
We will perform Laplace transformations and inverse La-
2.10.1.3 Laplace Transformation of Differentiation place transformations using Laplace transform pairs. Con-
with Respect to Time sequently, most of the effort involved in finding an inverse
The Laplace transformation of a function differentiated with Laplace transform of an output function will not be perform-
respect to time is, ing the inverse transformation. It will be the algebra needed
to create an exact match of the Laplace-domain function,
  df (t )  F ( s ), with an existing Laplace transform pair. A frequently
L  dt 
 = sF ( s ) − f 0
+
( ) (2.97)
needed operation is partial fraction expansion, introduced in
Sect. 2.10.5.
+
where f (0 ) is the time domain value of the function at time,
t = 0+ , the instant after a singularity input, such as a Heavi-
side unit step function, was applied to the system.
2.10  Laplace Transformation and Transfer Functions 93

2.10.3  Final Value and Initial Value Theorems Equation 2.103 is a transfer function. A transfer function
is a linear operator in the Laplace-domain. The numerator
Two useful Laplace transform theorems permit determina- polynomial contains the operations performed on the input,
tion of time-domain values without the need to perform the transformed into the Laplace-domain, where the Laplace
inverse Laplace transformation and evaluation of the time- variable, s, assumes the role of differentiation with respect
domain function. to time in the time domain. Correspondingly, the denomi-
The final value theorem is nator polynomial, a 0 s 2 + a1 s + a 2 , contains the operations
performed on the output variable in the differential equation.
 Notice the denominator of the transfer function is the char-
lim{ f (t )} = f ( ∞ ) = lim{ sF ( s )} (2.101)
t →∞ s→0 acteristic function of the differential equation. Setting the
characteristic function equal to zero forms the characteristic
The final value theorem must only be applied to functions equation of the differential equation:
which have a single final value. This excludes all periodic
functions and ramp functions. Beware that the final value a 0 s 2 + a1 s + a 2 = 0
theorem can be evaluated for functions which do not have a
final value. A transfer function is an unusual function, because it is used in
The initial value theorem is two very different ways. A transfer function can be evaluated
as a conventional function, by assigning a value to its argu-
lim+ { f (t )} = f 0+ = lim{ sF ( s )}
(2.102)
t →0
( ) s →∞
ment, the complex variable, s = σ + jw . Transfer functions
are complex functions, meaning, in the general case, the argu-
ment, s, and the result of evaluating the function are complex
2.10.4  Transfer Functions numbers. The second manner of using a transfer function is as
a multiplicative dynamic operator. Multiplying a transfer func-
A differential system equation is an input–output relation- tion by the Laplace transform of an input function yields the
ship. It is a dynamic operator which operates on the input Laplace transform of the output (response) function, Fig. 2.49.
to yield the output, where output is the response function of We are familiar with linear input–output relationships in
the dynamic system. We can create a second and very useful the time domain, such as the relationship between the dis-
form of the differential system equation by performing the placement and force of a spring, F (t ) = Kx(t ). If we rear-
Laplace transformation on it. The resulting algebraic equa- range this as a ratio of the output F( t) over the input x( t), we
tion is then manipulated into a “transfer function, as follows. have the form of a transfer function, where the left side is the
1. Use the Laplace transform of a differential system equa- ratio of output over input. What remains on the right side is
tion, neglecting the initial condition terms, to create an a multiplicative operator.
algebraic equation
F (t )
F (t ) = Kx (t ) → =K
x (t )
2 2
d u du d y dy
b0 + b1 + b 2 u = a 0 2 + a1 + a2 y
dt 2 dt dt dt
The spring equation, in the form of a transfer function, il-
 d 2u du   d2y dy  lustrates the use of a multiplicative operator, as follows. If
L b0 2 + b1
 dt dt
+ b 2u  =

L a 0 2 + a 1
 dt dt
+ a 2 y
 the input deformation, x( t), is known and the relationship be-
tween the spring deformation and spring force is expressed
b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s ) = a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s ) in the form of a transfer function:

x(t ) = 3 sin(2t ) F (t )
and =K
2. Factor out the transformed input and output variables x(t )
Then,
(b s
0
2
) ( )
+ b1 s + b2 U ( s ) = a 0 s 2 + a1 s + a 2 Y ( s )
F (t )
x (t ) = K 3sin ( 2t )
3. Create a ratio of the output variable over the input vari- x (t )
able
If we perform the Laplace transformation on the time-
Output Y ( s ) b2 s 2 + b1 s + b0
 = = ≡ G (s) (2.103) domain spring equation, where the unknown variables in
Input U ( s ) a2 s 2 + a1 s + a0 the time domain, x( t) and F( t), are transformed into the
94 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

U(s) Y(s) Next, distribute the variables.

Input Signal
G(s) Output Signal a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s ) = b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s )
Linear Operator
Now, apply the inverse Laplace transformation operator to
Fig. 2.49   A Laplace-domain block diagram, introduced in Chap. 9. both sides.
The lines represent Laplace-domain variables, (signals) U(s) and Y(s),
and the block contains the operator, the transfer function G(s)
L {a s Y ( s ) + a sY ( s ) + a Y ( s )}
−1
0
2
1 2

= L {b s U ( s ) + b sU ( s ) + b U ( s )}
−1 2
0 1 2
unknown variables, X( s) and F( s), in the Laplace-domain,
and then create the ratio of the output function over the input Distribute the operator, factor the coefficients out, and per-
function, we have a transfer function: form the inverse transformation.

L {F (t )} = L {Kx (t )} → L {F (t )} = KL { x (t )} a0 L {s Y ( s )} + a L {sY ( s )} + a L {Y ( s )}
−1 2
1
−1
2
−1

= b L {s U ( s )} + b L {sU ( s )} + b L {U ( s )}
−1 2 −1 −1

F (s) 0 1 2

F ( s ) = KX ( s ) → =K
X (s) The result is the time-domain differential equation.

2.10.4.1 Inverse Laplace Transformation d 2 y (t ) dy (t ) d 2 u (t ) du (t )


a0 + a1 + a 2 y (t ) = b0 + b1 + b 2 u (t )
of a Transfer Function dt 2 dt dt 2
dt
The inverse Laplace transformation of a transfer function
is simply the reverse of the process of creating a transfer 2.10.4.2 Use of Transfer Functions to Determine
function from a differential equation. The inverse Laplace Response Functions
transformation of an unknown Laplace-domain variable is We return to Example One of Sect. 2.8.1, to illustrate the
the corresponding unknown time-domain variable. use of transfer functions to determine the response of a sys-
tem to an input, i.e., to solve a differential equation. The first
L {Y ( s )} = y (t )
−1
example we solved using the method of undetermined coef-
ficients was:
The inverse transform of the product of the Laplace variable,
s, and a Laplace-domain variable is the time-domain variable dv
6 + 3v = F (t ) where F (t ) = 8us (t )
differentiated with respect to time. dt

d y (t ) We will again assume the initial condition, v(0+ ) = 0 .


L {sY ( s )} = dt
−1
We form the transfer function from the differential equa-
tion by performing the Laplace transformation, and then cre-
Likewise, the inverse Laplace transform of the product of ating the ratio to the output variable over the input variable.
the nth power of s and a Laplace-domain variable is the nth Recall that the initial conditions terms are neglected when
derivative. the Laplace transformation is applied to a differential system
equation to create a transfer function.
d n y (t )
L {s Y ( s )} =
−1 n

dt n
dv
L 6 dt + 3v = L {F (t )}
To determine the differential system equation which cor- dv
responds to a transfer function first, cross-multiply by the L 6 dt  + L {3v} = L {F (t )}
denominators to eliminate the ratios.
dv
Y ( s ) b 0 s + b1 s + b 2
2 6 L  dt  + 3L {v} = L {F (t )}
=
U ( s ) a 0 s + a1s + a 2
2

6sV ( s ) + 3V ( s ) = F ( s )
(a s
0
2
) (
+ a 1 s + a 2 Y ( s ) = b 0 s + b1 s + b 2 U ( s )
2
) V (s) 1
(6 s + 3)V ( s ) = F ( s ) → =
F ( s) 6s + 3
2.10  Laplace Transformation and Transfer Functions 95

Transfer functions are multiplicative operators. Multiplying a An exact match is needed to use a transform pair. Note there
transfer function by the Laplace transformation of the input is no coefficient multiplying s in the first-order factor in the
function yields the output function in the Laplace-domain. We denominator of the transform pair. The coefficient must be
will use the table of Laplace transform pairs, Table 2.3, to find removed from the response function.
the Laplace transform of the input function, F (t ) = 8us (t ) .
The relevant transform pair is  1
8   8
V (s) = → V (s) =  6 
f( t) F( s) s ( 6 s + 3) 1 s ( 6 s + 3)
 
us (t ) = 0 for t<0  6
1
us (t ) = 1 for t>0 s 8
6 1.33
V (s) = → V (s) =
The Laplace transformation of the input function is 6 3 s ( s + 0.5)
s s + 
6 6
8
L {F (t )} = L {8u (t )} s → F (s) = 8 L {u (t )} = s
s The coefficient 1.33 in the numerator of the response func-
tion can be factored out of the inverse Laplace transforma-
We now formulate the Laplace-domain response function, tion, since it is a linear operator.
V( s), as the product of the Laplace transformation of the
input function, F( s), and the transfer function, V( s)/F( s).  1.33 
L {V ( s )} = L
−1 −1
 
 s ( s + 0.5) 
Output ( s )
Input ( s ) = Output ( s )  1 
Input ( s ) v (t ) = 1.33L −1
 
 s ( s + 0.5) 
V (s)  8  1 
F (s) = V (s) →    = V (s)
F (s)  s   6 s + 3  1.33
v (t ) =
0.5
(
1 − e −0.5t ) (
→ v (t ) = 2.66 1 − e −0.5t )
To perform the inverse Laplace transformation, we peruse Note the above yields the same result we obtained from the
Table 2.3 for a Laplace-domain function of a similar form. method of undetermined coefficients in Sect. 2.81.
“Similar form” means similar products of polynomials in the
Laplace variable, s, in the numerator and denominator of the
output or response function V( s) and the Laplace-transform 2.10.5  Partial Fraction Expansion
table entry F( s).
The previous example used a first-order transfer function and
8 a step input. The simplicity of the transfer function and the
V (s) =
s ( 6 s + 3) input required only simple algebra to create the exact match,
with a Laplace transform pair needed for the inverse Laplace
The numerator of V( s) is a constant. (Constants are also re- transformation. Simple algebra may not be sufficient to yield
ferred to as zero-order polynomials, since any base, includ- a transform pair match, when the transfer function and/or the
ing a complex base, raised to the zero power equals one, input are of higher order. Although Laplace transform tables
8 = 8s 0 .) The denominator of V( s) is the product of the La- with hundreds of transform pairs exist, Table 2.3 is delib-
place variable, s, and a first-order polynomial in s. The rel- erately limited for didactic purpose, to the step response of
evant transform pair is second-order systems. Any response function with denomi-
nator more complicated than s (a 0 s 2 + a1 s + a 2 ) will not be
f ( t) F( s) found in Table 2.3.
1 1 Partial fraction expansion is used to split a Laplace-domain
(1 − e − at ) function (or signal) with a high-order denominator polynomial
a s( s + a)
into a sum of terms with lower-order denominators in order to
match entries in a Laplace transform table, and, thereby, de-
termine the inverse Laplace transform. The denominators of
96 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

the expansion terms are chosen to suit the engineer. The cor- Expand each product on the right side.
responding numerators are unknown, other than being polyno-
mials of s one order lower than the denominator, and must be (
xs 2 + ys + z = As 3 + A2ζw n s 2 + Aw n2 s )
found. For example,
(
+ Aas + Aa 2ζw n s + Aaw n2
2
)
N1 ( s ) + ( Bs 3
+ B 2ζw n s 2 + Bw n2 s )
F1 ( s ) =
(
s ( s + a ) s + 2ζw n s + w
2 2
n ) + (Cs 3 2
+ Cas + Ds + Das 2
)
A B Cs + D
= + + 2
s s + a s + 2ζw n s + w n2 Collect like powers of s.

where, N( s), a polynomial in s, is the numerator of the origi- xs 2 + ys + z = ( A + B + C ) s 3


nal transfer function. + ( A2ζw n + Aa + B 2ζw n + Ca + D ) s 2
Note, that the second-order denominator factor is written
in terms of a damping ratio ζ, and an ideal, undamped, natural (
+ Aw n2 + Aa 2ζw n + Bw n2 + Da s + Aaw n2 )
frequency ωn. This factor corresponds to a complex conjugate
pair of eigenvalues. Its inverse Laplace transform is a damped Equate the coefficients of like powers of s to create a set
oscillation. of simultaneous algebraic equations. Note that there is no
There is only one important special case, when the La- cubic term on the left side, which is equivalent to a coef-
place variable, s, occurs in the denominator of the original ficient equal to zero.
transfer function, as a factor raised to a power. In that case,
that denominator and all lower powers are included as terms s3 0 = A+ B +C
in the partial fraction expansion. For example, s2 is a factor
s2 x = A2ζw n + Aa + B 2ζw n + Ca + D
in the denominator of F2( s):
s1 y = Aw n2 + Aa 2ζw n + Bw n2 + Da
N2 (s) A B C D s0 z = Aaw n2
F2 ( s ) = + + +
s ( s + a )( s + b )
2
s 2
s s+a s+b
The unknowns are A, B, C, and D. Collect like unknowns and
move them to the right side of the coefficient terms.
The constants of the numerator terms are determined by
multiplying both sides by the denominator of the original s3 0 = A+ B +C
transfer function, canceling common factors, expanding into s2 x = ( 2ζw n + a ) A + 2ζw n B + aC + D
a polynomial in s, and equating coefficients of like terms to
form a set of algebraic equations.
s1 (
y = w n2 + a 2ζw n A + w n2 B + aD )
s 0
z = aw A 2
n
2.10.5.1  Example One: Partial Fraction Expansion
We will illustrate the technique using F1( s) above with The set of simultaneous equations can be solved by direct
N1 ( s ) = xs 2 + ys + z . First, multiply both sides by the de- substitution and elimination, Gaussian elimination, or using
nominator of the original signal. The denominators of the linear algebra. We will illustrate the use of linear algebra,
expansion terms are each a factor of the denominator of the the fundamentals of which are reviewed in Chap. 7. The left
original signal. Consequently, the multiplication will elimi- of the set of equations is expressed as a column vector. The
nate the denominators on both sides of the equation when the right side is the product of a four by four matrix and the col-
common factors are canceled. umn vector of the unknowns.

( xs 2
) (
+ ys + z s ( s + a ) s 2 + 2ζw n s + w n2 )
(
s ( s + a ) s 2 + 2ζw n s + w n2 )
=
(
A s ( s + a ) s 2 + 2ζw n s + w n2 ) + Bs ( s + a ) ( s 2
+ 2ζw n s + w n2 ) + (Cs + D ) s ( s + a ) ( s 2
+ 2ζw n s + w n2 )
s s+a s + 2ζw n s + w
2 2
n

( ) ( )
xs 2 + ys + z = A ( s + a ) s 2 + 2ζw n s + w n2 + Bs s 2 + 2ζw n s + w n2 + (Cs + D ) s ( s + a )
2.10  Laplace Transformation and Transfer Functions 97

s3 0  1 1 1 0  A  2.10.5.2 Example Two: Derivation


 x   2ζw + a of a Laplace Transform Pair
s2
2ζw n a 1   B 
 = n   Our second example partial fraction expansion will also il-
s1  y  w n2 + a 2ζw n w n2 0 a C  lustrate how a Laplace transform pair can be derived from
    
s0 z  aw n2 0 0 0  D an existing transform pair. We will derive the transform pair
for the signal:
Division of a matrix by a matrix is not defined in linear alge- 1
bra. The equivalent operation is to multiply a matrix and its F (s) =
inverse to yield the “identity matrix,” I. The identity matrix,
( s + a ) ( s + b)
I, consists of ones on the “main” diagonal, from top left to using the transform pair of an exponential decay, Eq. 2.95.
lower right and zeros elsewhere. The identity matrix, I, fills The first step is to express the function as the sum of two
the role that the number one fills in scalar multiplication. terms:
It can be inserted or removed as a factor without changing
 1 A B
a product. Matrix multiplication has restrictions. Except for F (s) = = + (2.104)
the identity matrix, I, and a matrix and its inverse, matrix ( s + a )( s + b) s+a s+b
multiplication is not commutative, meaning that the position
of factors in a product cannot be changed without affecting where A and B are constants to be determined in terms of the
the result. Further, excepting only the identity matrix, I, a constants a and b. Follow the technique of multiplying both
factor cannot be inserted into an existing product. We must sides by the denominator of the original signal; cancel terms;
multiply by appending a factor to the left or right end of a and then expand and collect the coefficients of like powers
product. The former is “pre-multiplication” and the latter is of s to create a set of simultaneous equations.
“post-multiplication.”
We solve for the vector of unknowns by pre-multiplying ( s + a ) ( s + b) = A ( s + a ) ( s + b) + B ( s + a ) ( s + b)
both sides by the inverse of the matrix on the right side. ( s + a ) ( s + b) s+a s+b

−1 −1
 1 1 1 0 0  1 1 1 0  1 1 1 0  A 
 2ζw + a 2ζw n a 1   x   2ζw + a 2ζw n a 1   2ζw + a 2ζw n a 1   B 
 n  = n  n  
w n2 + a 2ζw n w n2 0 a  y  w n2 + a 2ζw n w n2 0 a w n2 + a 2ζw n w n2 0 a C 
        
 aw n2 0 0 0 z  aw n2 0 0 0  aw n2 0 0 0  D

The product of a matrix and its inverse is I.

( s + a )( s + b) A ( s + a ) ( s + b) B ( s + a ) ( s + b)
 1 1 1 0
−1
 0  1 0 1 0  A  = +
 2ζw + a ( s + a )( s + b) s+a s+b
 2ζw n a 1   x  0 1 0 0  B 
n  =
w n2 + a 2ζw n w n2 0 a  y  0 0 1 0  C  1 = A ( s + b) + B ( s + a )
      
 aw n2 0 0 0  z  0 0 0 1  D 

Expand and collect the coefficients of like powers of s.


which can be removed, yielding the vector of unknowns.
1 = As + Ab + Bs + Ba → 1 = ( A + B ) s + ( Ab + Ba )
−1
 1 1 1 0  0  A
 2ζw + a 2ζw n a 1   x  B  Equate coefficients of like powers of s on both sides of the
 n  =  equation to yield a set of two simultaneous equations with
w n2 + a 2ζw n w n2 0 a  y C 
      two unknowns, A and B.
 aw n2 0 0 0  z   D
s1 0 = A + B
s 0 1 = Ab + Ba
98 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Solve for A and B using direct elimination, Gaussian elimi- −1 − at 1 − bt


f (t ) = e + e
nation, or linear algebra. Using linear algebra, we find: a −b a −b

s1 0 = A + B  0   1 1   A Combine terms and order the exponentials alphabetically by


→  =  
s 0 1 = bA + aB  1  b a   B  their exponents:

Premultiply both sides by the inverse of the matrix. 1  −1 1


f (t ) = (e − bt − e − at ) → f (t ) =   (e − bt − e − at )
a −b  
−1 a − b
−1 −1
1 1   0 1 1   1 1   A
b a   1  = b a  b a   B  to yield
        

1 1 
−1
 0  A  1  − at
f (t ) =  (e − e − bt )
b a  1  = I  B   b − a 
     
 −1  Hence,
 A  a − b 
−1
 1 1   0   A
b a   1  =  B  →  B  =  1   1 
          L {F ( s )} = L
−1 −1
 
 a − b   ( s + a )( ) 
s + b

1
The partial fraction expansion of F( s) is L {F ( s )} =  b − a  (e
−1 − at
)
− e − bt = f (t )
1 A B
F (s) = = + or
( s + a )( s + b) s+a s+b

 −1   1  
 1  − at − bt  1
   
a − b  a − b L 
 b a
(
 e − e  =
( s a
)
)( s + b)
(2.105)
F (s) = +  −  +
s+a s+b

We now use the Laplace transform pair for an exponential We have derived the Laplace transform pair:
decay:
  1  − at
1 (
 f (t ) =  b − a  e − e
− bt
)
f (t ) = e − at
↔ F (s) = 
s+a 
F (s) = 1
  −1   1  

 ( s + a )( s + b)

L {F ( s )} = L
−1 −1  
 a − b  a − b 
 s + a + s + b  Again, the constants, a and b, can be purely real or com-
plex. If they are complex conjugates, then the time-domain
The inverse Laplace transformation is a linear operator. It expression is a decaying sinusoid. The purely real, time-do-
can be distributed onto a sum. main expression is derived using the logic we used to solve
oscillatory responses with the method of undetermined coef-
  −1    1   ficients. Beginning by substituting a complex conjugate pair
   
L −1
{ F ( s )} = L −1
 a − b  + L −1
 a − b  for a and b, a = σ + jw and b = σ − jw
 s + a   s + b 
 1  − at
The numerator terms are constants and can be factored out of
f (t ) = 
 b − a 
(
e − e − bt )
the inverse Laplace transformation.
  − (σ + jw )t
−1  1  1  1 
f (t ) = 
 (σ − jw )
1
− (σ + jw )  e

(
− e − (σ − jw )t )
f (t ) = L L
−1 −1
 +  
a−b s + a a − b s + b
We next use the property of exponentials, e a + b = e a eb .
We have an exact match between the argument of the inverse
Laplace transform, and the transform pair for a decaying
exponential. Therefore, we can perform the inverse Laplace
f (t ) =
j 2w
e(
−1 − (σ + jw )t
− e − (σ − jw )t )
transformation using that transform pair.
−1 −σ t − jwt
f (t ) =
j 2w
(
e e − e − σ t e jw t )
2.10  Laplace Transformation and Transfer Functions 99

Fig. 2.50  a Spring-mass-damper


system driven by an input force, a x,v b 25
F( t). b Plot of the input force
b
( )
F (t ) = 10us (t ) + 10 1 − e −0.25t 20

F(t) F(t) 15
M N 10
5
K
0
0 5 10 15 20
t, sec

We now factor out the purely real exponential, and group the Begin by creating the transfer function of the system
purely imaginary exponentials, the complex exponential unit equation. Perform the Laplace transformation on both sides
vectors, to use Euler’s sine formula. of the system equation, neglecting the initial condition terms.

−1  e − jw t − e jw t  1 − σ t  e jw t − e − jw t   d 2v 
f (t ) = e −σ t  → f (t ) = e   dF  dv
w  j2  w  j2  L  =
 dt 
L  M 2 + b + Kv 
 dt dt 
1
f (t ) = e −σ t sin(w t ) Distribute the Laplace transformation operator onto each
w
term of the sum.
Hence,
dF  d 2v  dv

 1 −σ t  1 L  dt  = L M dt 2 
+ L b dt  + L {Kv}
L  e sin (w t ) = (2.106) 
w  ( s + σ + jw )( s + σ − jw )
Factor out the parameter terms, which are constants.
or
dF  d 2v  dv

 1 −σ t  1 L  dt  = M L  dt 2 
+b L  dt  + KL {v}
L  e sin (w t ) = 2 (2.107) 
 w  s + 2σ s + σ 2 + w2
The Laplace transformation of an unknown time-domain
2.10.5.3 Example Three: Response variable is an unknown Laplace-domain variable. Differen-
of a Spring-Mass-Damper System tiation with respect to time transforms to the Laplace vari-
This example illustrates the use of transfer functions and par- able, s.
tial fraction expansion, to find the response of an oscillatory
second-order system to an input which is superposition of a sF ( s ) = Ms 2V ( s ) + bsV ( s ) + kV ( s )
step and a stable first-order exponential growth. Determine
the velocity of the mass of the spring-mass-damper system Collect the transformed output variable V( s). A transfer
shown in Fig. 2.50a. The system equation relating the input function is the ratio of the output over the input:
force to the velocity of the mass is
(
sF ( s ) = Ms 2 + bs + k V ( s ))
dF d 2v dv
= M 2 + b + Kv Output V ( s ) s
dt dt dt = =
Input F ( s ) Ms 2 + bs + k

with b = 1, M = 2, and K  = 6. The force input is Note that the operators from the output variable side of the
F (t ) = 10us (t ) + 10(1 − e −0.25t ) , Fig. 2.50b. system equation are in the dominator of the transfer function.
100 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

The standard form for a transfer function requires the coef-


1
ficient of the highest-order term of the denominator to be Factor into the second transform:
0.25
cleared. This form corresponds to the standard form of a dif-
ferential equation. 1
1 L {F (t )} = 10L {u (t )} + 2.5L  0.25 (1 − e
s
−0.25t
)
V (s) s
= M
F (s) b k We now have the exact matches needed to perform the trans-
s2 + s+ formations using the Laplace transform pairs.
M M

1 10 2.5
L { f (t )} = 10L {u (t )} + 2.5L  0.25 (1 − e
s
−0.25t
) → F (s) = +
s s ( s + 0.25)

Substitute in the parameter values. A transfer function is a multiplicative operator. Multiply the
transfer function by the Laplace transform of the input to
1 yield the Laplace transform of the output.
V (s) s V (s)
2 0.5s
= → = 2
F (s) 1 6 F ()
s s + 0.5s + 3
s2 + s + V ( s )  10 2.5  0.5s 
2 2 F (s) = +   = V (s)
F ( s )  s s ( s + 0.25)   s 2 + 0.5s + 3 
Next, determine the Laplace transformation of the input:
Cancel the Laplace variable, s, where it occurs in both the
(
F (t ) = 10us (t ) + 10 1 − e −0.25t
) numerator and denominator.

 10   0.5 s   2.5  0.5 s 


The relevant Laplace transform pairs from Table 2.3 are:  s   s 2 + 0.5s + 3  +  s s + 0.25   s 2 + 0.5s + 3  = V ( s )
 ( )
f ( t) F( s)
us (t ) = 0 for t<0
The Laplace transform of the output variable is
1
us (t ) = 1 for t>0 s
5 1.25
+ = V (s)
1
a
(1 − e − at )
1
s(s + a)
(
s + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3
2
)
We now wish to perform the inverse Laplace transforma-
Apply the Laplace transformation operator to both sides, and tion to return to the time-domain. Note the denominators
distribute to the terms of the right side. of the two terms contain the common factor, second-order
factor, s 2 + 0.5s + 3 . This is the characteristic function of
L {F (t )} = L {10u (t ) + 10 (1 − e
s
−0.25t
)} the spring-mass-damper system. In order to use the Laplace
transform pairs, we must identify whether the characteristic
function represents an oscillatory or non-oscillatory homo-
L {F (t )} = L {10u (t )} + L {10 (1 − e
s
−0.25t
)} geneous or natural response. If the eigenvalues of the system
are real, then the homogeneous response is not oscillatory.
We must create exact matches with the time-domain func- If the eigenvalues are a complex conjugate pair, then the re-
tions of the Laplace transform pairs. Factor the coefficient 10 sponse is oscillatory.
out of both the transformations. Multiply the second trans- Set the characteristic function equal to zero, to form the
form by the unity ratio, 0.25/0.25. characteristic equation and solve it using the quadratic equa-
tion.
L {F (t )} = 10L {u (t )} + 10 0.25 L {(1 − e )}
0.25 −0.25t
s
−0.5 ± (0.5)2 − ( 4)(3)
s 2 + 0.5s + 3 = 0 → s1 , s2 =
2
s1 , s2 = −0.25 ± j1.71
2.10  Laplace Transformation and Transfer Functions 101

The eigenvalues s1 , s2 = −0.25 ± j1.71 are a complex conju- nator of the second term into a summation with denomina-
gate pair. Therefore, the homogeneous or natural response of tors that match Laplace transform pairs. The denominators
the system is oscillatory or underdamped. Accordingly, we of the partial fraction expansion terms are factors of the left
must use Laplace transform pairs which correspond to oscil- side. The numerators of the expansion terms are polynomials
latory time-domain responses. These are simple to identify in s, one order lower than their denominators. The constants,
in Table 2.3; the time-domain response contains a sinusoid. A, B, and C must be determined.
The denominator of the corresponding Laplace-domain sig-
nal is written in terms of the ideal, undamped natural fre- 1.25 A Bs + C
= + 2
quency, ωn, and the damping ratio, ζ. Specifically, the two ( s + 0.25) ( s 2
+ 0.5s + 3) s + 0.25 s + 0.5s + 3
terms of the output
Multiply both sides by the denominator of the left side,
5 1.25
+ = V (s) and cancel factors which appear in the numerator and
(
s 2 + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3 ) denominator.

(
1.25 ( s + 0.25) s 2 + 0.5s + 3 ) (
A ( s + 0.25) s 2 + 0.5s + 3 ) + ( Bs + C )( s + 0.25) ( s 2
+ 0.5s + 3 )
=
( s + 0.25) ( s 2 + 0.5s + 3) s + 0.25 2
s + 0.5s + 3

( )
1.25 = A s 2 + 0.5s + 3 + ( Bs + C )( s + 0.25)

correspond to Distribute the unknown constants.

5
= 2
5 ( ) (
1.25 = As 2 + A0.5s + A3 + Bs 2 + Bs 0.25 + (Cs + C 0.25))
s + 0.5s + 3 s + 2ζw n s + w n2
2

Collect like powers of s.


and
1.25 = ( A + B ) s 2 + ( A0.5 + B0.25 + C ) s + A3 + C 0.25
1.25 1.25
= Create a set of three equations, by equating the coefficients
( s + 0.25) ( s 2
+ 0.5s + 3 ) (s + a)(s 2
+ 2ζw n s + w n2 ) of s on the left with that on the right side. If a power of s is
missing from the left side, then its coefficient equals zero.
Checking the forms of the denominators of the terms against
the forms of the denominators of the transform table pairs, s2 0 = A+ B
1
we find a match for the first term. The relevant transform s 0 = 0.5 A + 0.25 B + C
pair is 0
s 1.25 = 3 A + 0.25C
f (t ) F (s) We will solve for the constants with linear algebra. First, ex-
wn −ζw nt w n2 press the set of equations as a single vector-matrix equation.
e sin(w n 1 − ζ t )
2

1−ζ 2 s + 2ζw n s + w n2
2

 0  1 1 0   A
 0  = 0.5 0.25 1   B
    
There is not a transform pair table match for the second term. 1.25  3 0 0.25 C 
We must use partial fraction expansion to break the denomi-
Premultiply both sides by the inverse of the matrix.

−1 −1
1 1 0   0  1 1 0  1 1 0   A
0.5 0.25 1   0  = 0.5 0.25 1  0.5 0.25 1   B 
    
 3 0 0.25 1.25  3 0 0.25  3 0 0.25 C 
102 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

5 0.426 0.426 s 0.106


The product of a matrix and its inverse equals the identity 2
+ − 2 − 2 = V (s)
matrix, I, which can be removed. s + 0.5s + 3 s + 0.25 s + 0.5s + 3 s + 0.5s + 3
Inspecting the output signal, we notice that the first and last
−1
1 1 0   0  1 0 0  A terms have a common denominator and can be summed,
0.5 0.25 1       eliminating a term.
  0  =  0 1 0  B 
 3 0 0.25 1.25 0 0 1  C 
4.894 0.426 0.426 s
2
+ − 2 = V (s)
−1 s + 0.5s + 3 s + 0.25 s + 0.5s + 3
1 1 0   0   A
0.5 0.25 1   0  = I  B The output variable, V( s), is now a sum containing three differ-
    
ent transform pairs which are in the Laplace transform table,
 3 0 0.25 1.25 C 
Table 2.3. The relevant Laplace transform pairs are:
−1
1 1 0   0   A  0.426   A f (t ) F (s)
0.5 0.25 1         
  0  =  B →  −0.426 =  B  1
e − at
 3 0 0.25 1.25 C   −0.106 C  s+a

Substitute the values of A, B, and C into the partial fraction


ωn
1−ζ 2 (
e −ζω nt sin ω n 1 − ζ 2 t ) ω n2
s + 2ζω n s + ω n2
2

expansion:

1
1−ζ 2
(
e −ζω nt sin ω n 1 − ζ 2 t − φ ) s
s 2 + 2ζω n s + ω n2
1.25 A Bs + C
= +
( s + 0.25) ( s 2 + 0.5s + 3) s + 0.25 s 2 + 0.5s + 3  ωd   1−ζ 2 
φ = tan −1   = tan −1  
 σ   ζ 
1.25 0.426 −0.426 s − 0.106
= +
( s + 0.25) ( s 2 + 0.5s + 3) s + 0.25 s 2 + 0.5s + 3 Again, we must make exact matches against the table trans-
form pairs to perform the inverse Laplace transformation
using the transform pairs.

 4.894   0.426   0.426 s 


L {V ( s )} = L
−1 −1
 2 +
 s + 0.5s + 3 
L −1
 −
 s + 0.25 
L −1
 2 
 s + 0.5s + 3 

Factor out the numerator coefficients.

 1   1   s 
v (t ) = 4.894 L L L
−1 −1 −1
 2  + 0.426   − 0.426  2 
 s + 0.5s + 3   s + 0.25   s + 0.5s + 3 

The second term has a first-order numerator. Write the sec- The second and third terms match their Laplace transform
ond term as two terms. pair. Multiply the first term by the unity ratio, 3/3.

1.25
( s + 0.25) ( s 2 + 0.5s + 3)  3
v (t ) = 4.894  

L
 2
−1 1 
 + 0.426 L −1  1 
 
 3  s + 0.5s + 3   s + 0.25 
0.426 0.426 s 0.106
= − 2 − 2 −1  s 
s + 0.25 s + 0.5s + 3 s + 0.5s + 3 − 0.426 L 2
 s + 0.5 s + 3


Now, assemble the output signal, V( s).
Factor three into the numerator of the first term and evaluate
5 1.25 the remaining ratio.
+ = V (s)
(
s 2 + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3 )
2.10  Laplace Transformation and Transfer Functions 103

 3   1   s 
v (t ) = 1.63 L L L
−1 −1 −1
 2  + 0.426   − 0.426  2 
 s + 0.5s + 3   s + 0.25   s + 0.5s + 3 

To evaluate the oscillatory time-domain terms, we must cal-


culate the system’s ideal, undamped, natural frequency ωn
v1 (t ) = 1.63
1.73
1 − ( 0.144)
2 (
e − (0.144)(1.73)t sin 1.73 1 − ( 0.144) t
2
)
and the damping ratio ζ. The simplest method is to equate
coefficients of like powers of s of the denominators of an v1 (t ) = 2.85e −0.249t sin (1.71t )
oscillatory term, and the corresponding Laplace-transform
table expression. This term happens to be the Laplace transform pair we de-
rived as Example Two. The exponent is the real component
s 2 + 0.5s + 3 = s 2 + 2ζw n s + w n2 of the eigenvalues, s1 , s2 = −0.25 ± j1.71 . The discrepancy
is due to numerical precision. The frequency 1.71 is the mag-
First, calculate the ideal, undamped, natural frequency ωn, nitude of the imaginary component of the eigenvalues. The
which is the square root of the constant term. units are radians per second. Again, eigenvalues are the char-
acteristic values of a system. They do indeed characterize the
w n = 3 = 1.73 system’s natural or homogeneous response, which is excited
by any change in the input to the system.
Then, using the value of the natural frequency ωn, calculate The inverse Laplace transformation of the second term is
the damping ratio ζ.
 1 
v2 (t ) = 0.426 L v2 (t ) = 0.426 e −0.25t
−1
0.5   →
0.5 = 2ζw n → ζ= → ζ = 0.144  s + 0.25 
2w n
The Laplace variable, s, in the numerator of the third terms
The damping ratio of an oscillatory factor is in the range of creates a constant φ, known as the phase shift, which is sub-
zero to one, 0 < ζ < 1. If the calculated damping ratio is equal tracted from the time-varying angle of the sine, w n 1 − ζ 2 t .

 
v3 (t ) = −0.426L
−1 
 2
s 
 = −0.426  −
 s + 0.5s + 3  
1
1−ζ 2
(
e −ζw t sin w n 1 − ζ 2 t − φ 
n


)
 
v3 (t ) = −0.426  −

1
1 − ( 0.144)
2 (
e − (0.144)(1.73)t sin 1.73 1 − ( 0.144) t − φ
2
)
where the phase shift φ is

w   1−ζ 2   1 − ( 0.144)2 
φ = tan  d  = tan −1 
−1
 → φ = tan 
−1
 = 1.43 rad
 σ   ζ   0.144 

to or greater than one, then the factor is not oscillatory, and a Yielding
different Laplace transform pair must be used.
We now use the Laplace transform pair table to perform v3 (t ) = 0.430 e −0.249t sin (1.71t − 1.43)
the inverse Laplace transformation of each term. The first
term yields: The time-domain response function plotted in Fig. 2.51, is
 3  v (t ) = v1 (t ) + v2 (t ) + v3 (t )
v1 (t ) = 1.63L −1
 2 
 s + 0.5s + 3 
v (t ) = 2.85e −0.249t sin (1.71t ) + 0.426 e −0.25t
v1 (t ) = 1.63
wn
1−ζ 2
e −ζw n t
(
sin w n 1 − ζ t 2
) + 0.430 e −0.249t sin (1.71t − 1.43)
104 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

3 dv (t )
2.3.a 3F ( t ) = 9
dt
+ 6v ( t ) F (t ) = 14 us (t ) ( )
v 0+ = 0
2 dF (t ) dv (t )
2.3.b 3
dt
=9
dt
+ 6v ( t ) F (t ) = 14 us (t ) ( )
v 0+ = 34
1
m
v(t), ___ dv (t )
sec 2.3.c 3F ( t ) = 9
dt
+ 6v ( t ) F (t ) = 14 us (t ) ( )
v 0+ = 34
0
dv (t )
2.3.d 3F ( t ) = 9
dt
+ 6v ( t ) F (t ) = 14 us (t ) ( )
v 0+ = 22
-1

-2 Problem 2.4 Use the method of undetermined coefficients to


0 5 10 15 20
t, sec solve the ordinary differential equations with constant coef-
ficients below for the input F (t ) = 10 us (t ). Plot the homo-
Fig. 2.51   Response v(t) of the spring-mass-damper system to the input geneous, particular, general solution on the same plot using
force Mathcad or MATLAB.

2.4.a dF (t ) d 2v (t )
= +3
dv (t )
+ v (t ) ( )
v 0+ = 7
( ) = 14
d v 0+
Problems dt dt 2 dt dt

2.4.b F (t ) =
d 2 v (t )
+3
dv (t )
+ v (t ) ( )
v 0+ = 0
( )=7
d v 0+
Problem 2.1 Use the method of undetermined coefficients dt 2 dt dt
to solve the ordinary differential equations with constant co-
F (t ) =
d 2 v (t ) dv (t )
+ v (t ) ( )
v 0+ = 6
( )=0
d v 0+
efficients below. Plot the homogeneous, particular, general 2.4.c +3
dt 2 dt dt
solution on the same plot using Mathcad or MATLAB.
2.4.d F (t ) =
d 2 v (t )
+3
dv (t )
+ v (t ) ( )
v 0+ = 10
( )=3
d v 0+
2
dF (t ) dv (t ) dt dt dt
2.1.a
dt
=3
dt
+ v (t ) F (t ) = 10 us (t ) ( )
v 0+ = 7

dv (t ) Problem 2.5 Use the method of undetermined coefficients to


2.1.b F (t ) = 3 + v (t ) F (t ) = 10 us (t ) ( )
v 0+ = 0
dt solve the ordinary differential equations with constant coef-
dv (t ) ficients below for the input F (t ) = 12 us (t ). Plot the homo-
2.1.c F (t ) = 3
dt
+ v (t ) F (t ) = 10 us (t ) ( )
v 0+ = 6
geneous, particular, general solution on the same plot using
dv (t ) Mathcad or MATLAB.
2.1.d F (t ) = 3
dt
+ v (t ) F (t ) = 6 u s (t ) ( )
v 0+ = 10

2.5.a 2 F (t ) =
d 2 v (t )
+8
dv (t )
+ 3v (t ) ( )
v 0+ = 5
( ) = 18
d v 0+
Problem 2.2 Use the method of undetermined coefficients dt 2 dt dt
to solve the ordinary differential equations with constant co-
2.5.b 2 F (t ) =
d 2 v (t )
+8
dv (t )
+ 3v (t ) ( )
v 0+ = 0
( ) = 12
d v 0+
efficients below. Plot the homogeneous, particular, general dt 2 dt dt
solution on the same plot using Mathcad or MATLAB.
2.5.c 2
dF (t ) d 2 v (t )
= +8
dv (t )
+ 3v (t ) v 0 = 9
+
( ) ( )=0
d v 0+
dv (t ) dt dt 2 dt dt
2.2.a 2 F (t ) = 8
dt
+ 3v (t ) F (t ) = 12 us (t ) ( )
v 0+ = 5
d 2 v (t ) dv (t ) ( )=9
d v 0+
2.5.d 2 F (t ) = +8 + 3v (t ) ( )
v 0+ = 18
dv (t ) dt 2 dt dt
2.2.b 2 F (t ) = 8
dt
+ 3v (t ) F (t ) = 12 us (t ) ( )
v 0+ = 0

dF (t ) dv (t ) Problem 2.6 Use the method of undetermined coefficients to


2.2.c 2
dt
=8
dt
+ 3v (t ) F (t ) = 12 us (t ) ( )
v 0+ = 9
solve the ordinary differential equations with constant coef-
dv (t ) ficients below for the input F (t ) = 14 us (t ). Plot the homo-
2.2.d 2 F (t ) = 8
dt
+ 3v (t ) F (t ) = 6 u s (t ) ( )
v 0+ = 18 geneous, particular, general solution on the same plot using
Mathcad or MATLAB.

Problem 2.3 Use the method of undetermined coefficients 2.6.a 3F (t ) =


d 2 v (t )
+9
dv (t )
+ 6v ( t ) ( )
v 0+ = 0 ( ) = 18
d v 0+
to solve the ordinary differential equations with constant co- dt 2 dt dt
efficients below. Plot the homogeneous, particular, general
2.6.b 3
dF (t ) d 2 v (t )
= +9
dv (t )
+ 6v (t ) v 0 = 34
+
( ) ( ) = 12
d v 0+
solution on the same plot using Mathcad or MATLAB. dt dt 2
dt dt

2.6.c 3F (t ) =
d 2 v (t )
+9
dv (t )
+ 6v ( t ) ( )
v 0+ = 34 ( )=0
d v 0+
2
dt dt dt
Chapter 2 Appendix 105

d 2 v (t ) dv (t ) ( )=9
d v 0+ stant coefficients below for the input F (t ) = 10 us (t ) . Plot
2.6.d 3F (t ) =
dt 2
+9
dt
+ 6v ( t ) ( )
v 0+ = 22
the solution using Mathcad or MATLAB.
dt

Problem 2.7 Use the method of undetermined coefficients to d 2 v (t ) dv (t )


2.10.a 2 F (t ) = +8 + 3v (t )
solve the ordinary differential equations with constant coef- dt 2 dt
ficients below for the input F (t ) = 10 us (t ). Plot the homo- d 2 v (t ) dv (t )
2.10.b F (t ) = +3 + 5v (t )
geneous, particular, general solution on the same plot using dt 2
dt
Mathcad or MATLAB. d 2 v (t ) dv (t )
2.10.c 2 F (t ) = +4 + 9v ( t )
dt 2 dt
2.7.a
dF (t ) d 2 v (t )
= +3
dv (t )
+ 5v (t ) v 0 = 7
+
( ) ( ) = 14
d v 0+
d 2 v (t ) dv (t )
dt dt 2 dt dt 2.10.d 3F ( t ) = +5 + 18v (t )
dt 2 dt
2.7.b F (t ) =
d 2 v (t )
+3
dv (t )
+ 5v (t ) ( )
v 0+ = 0 ( )=7
d v 0+
dt 2 dt dt

2.7.c F (t ) =
d 2 v (t )
+3
dv (t )
+ 5v (t ) ( )
v 0+ = 6 ( )=0
d v 0+ Problem 2.11 Use Laplace transforms and transfer functions
dt 2 dt dt to solve the ordinary differential equations with constant coef-
d v (t ) dv (t ) ( )=3   −t

d v 0+ ficients below for the input F (t ) = 10 us (t )  2 + 3 1 − e 2   .
2
2.7.d F (t ) =
dt 2
+3
dt
+ 5v (t ) ( )
v 0+ = 10
  
dt
Plot the solution using Mathcad or MATLAB.
Problem 2.8 Use the method of undetermined coefficients to
d 2 v (t ) dv (t )
solve the ordinary differential equations with constant coef- 2.11.a 2 F (t ) = +8 + 3v (t )
dt 2 dt
ficients below for the input F (t ) = 12 us (t ). Plot the homo-
geneous, particular, general solution on the same plot using d 2 v (t ) dv (t )
2.11.b F (t ) = 2
+3 + 5v (t )
Mathcad or MATLAB. dt dt
d 2 v (t ) dv (t )
2.11.c 2 F (t ) = +3 + 8v (t )
2.8.a 2 F (t ) =
d 2 v (t )
+3
dv (t )
+ 8v (t ) ( )
v 0+ = 5 ( ) = 18
d v 0+ dt 2 dt
dt 2 dt d 2 v (t ) dv (t )
dt 2.11.d 3F ( t ) = +5 + 9v ( t )
dt 2 dt
2.8.b 2 F (t ) =
d 2 v (t )
+3
dv (t )
+ 8v (t ) ( )
v 0+ = 0 ( ) = 12
d v 0+
dt 2 dt dt

2.8.c 2
dF (t ) d 2 v (t )
= +3
dv (t )
+ 8v (t ) ( )
v 0+ = 9 ( )=0
d v 0+
2
dt dt dt dt
Chapter 2 Appendix
2.8.d 2 F (t ) =
d 2 v (t )
+3
dv (t )
+ 8v (t ) ( )
v 0+ = 18 ( )=9
d v 0+
dt 2 dt dt
Table 2.3   Laplace Transform Pairs
f( t) F( s)
Problem 2.9 Use the method of undetermined coefficients to
solve the ordinary differential equations with constant coef- Unit impulse: δ (t ) = 0 for t ≠ 0
1
(Dirac delta function) δ (t ) = ∞ for t = 0
ficients below for the input F (t ) = 14 us (t ). Plot the homo-
geneous, particular, general solution on the same plot using us (t ) = 0 for t<0 1
Mathcad or MATLAB. Unit step:
s
us (t ) = 1 for t>0

d 2 v (t ) dv (t ) ( ) = 18
d v 0+ ur (t ) = 0 for t<0 1
2.9.a 3F ( t ) =
dt 2
+9
dt
+ 6v ( t ) ( )
v 0+ = 0
dt
Unit ramp:
s2
ur (t ) = t for t>0

2.9.b 3
dF (t ) d 2 v (t )
= +5
dv (t ) +
( )
+ 9v (t ) v 0 = 34
( ) = 12
d v 0+
1
dt dt 2 dt dt e − at
s+a
2.9.c 3F ( t ) =
d 2 v (t )
+5
dv (t )
+ 9v ( t ) ( )
v 0+ = 34
( )=0
d v 0+
1 1
dt 2 dt dt
a
(
1 − e − at ) s (s + a)
2.9.d 3F ( t ) =
d 2 v (t )
+5
dv (t )
+ 9v ( t ) ( )
v 0+ = 22
( )=9
d v 0+
1 1
dt 2 dt dt
b−a
(
e − at − e − bt ) ( s + a )( s + b)
Problem 2.10 Use Laplace transforms and transfer func- 1 s
b−a
(
be − bt − ae − at ) ( s + a )( s + b)
tions to solve the ordinary differential equations with con-
106 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Table 2.3  (continued) purpose computer programming languages in use during that


f ( t) F( s) early period were written to implement mathematical algo-
1  1  1 rithms for engineers and scientists. MathWorks foresaw the
1+
ab  a − b
(
be − at − ae − bt 

) s ( s + a )( s + b ) time when general purpose programming languages would
evolve into “object-oriented” languages to manage the ex-
ωn
1−ζ 2
(
e −ζω nt sin ω n 1 − ζ 2 t ) ω n2
s + 2ζω n s + ω n2
2
ecution of enormous programs, and no longer permitted
simple programming of mathematical algorithms. MATLAB
is a “procedural” language intended to encode algorithms.

1
1−ζ 2
(
e −ζω nt sin ω n 1 − ζ 2 t − φ ) The syntax and execution of MATLAB resemble that of the
s original BASIC computer language from the 1960s, which
 ωd   1−ζ  2 s 2 + 2ζω n s + ω n2 was developed as an introductory programming language.
φ = tan −1  −1
 = tan  ζ 
 σ   

Plotting in Mathcad
1−
1
1−ζ 2 (
e −ζω nt sin ω n 1 − ζ 2 t + φ )
ω n2
Mathcad Prime is engineering computational software de-
ω   1−ζ 2  s ( s 2 + 2ζω n s + ω n2 )
φ = tan  d  = tan −1 
−1 signed to resemble the layout and appearance of manual

 σ   ζ  computations. Equations, graphics, and text can be placed at
will in the worksheet, with the only restriction that constants
ω and functions must be defined above where they are used,
sin (ω t )
s2 + ω 2 because the worksheet is computed from top to bottom, left
s to right. Mathcad strives to use conventional mathematical
cos (ω t ) notation, making the worksheet significantly easier to read
s + ω2
2

than computer code. Plotting a function in Mathcad requires


little more than defining the function to be plotted using an
assignment statement.
Mathcad and MATLAB
Mathcad Assignment Statements
In the early 1980s, shortly after personal computers were in- Assignment statements are instructions to a program to as-
troduced, there was a flurry of activity to develop software sign a value or expression to a variable. Assignment state-
for the emerging market. Two Massachusetts companies tar- ments are read from right to left. In other words, the quantity
geted engineering and scientific computation with radically on the right side of an assignment statement is assigned to
different approaches. Mathsoft, the originator of Mathcad, the variable or function on the right. Although most com-
chose to develop software which emulated the mixture of puter languages, including MATLAB, use the convention,
mathematics, graphics, and text, which typifies manual engi- equal sign i.e., =, as the “assignment operator,” Mathcad
neering calculations. A Mathcad “worksheet” presents equa- does not. The Mathcad assignment operator is :=, which is
tions in as close to standard mathematical notation as pos- created by typing a colon. It is supposed to connote an arrow
sible. What the user sees is actually a graphical interface de- pointing from right to left. The reason Mathcad uses := rath-
fined by user created “regions” or objects that support math- er than = as the assignment operator is because assignment
ematics, graphics, or text. The regions within the worksheet statements are not equalities. Assignment statements are
can be dragged around, resized, copied, and edited. Mathcad similar to a “store” command on a calculator. An assignment
worksheets are evaluated automatically from top to bottom, statement is an instruction to store the value or expression on
left to right when any change is made. the right within the variable or function name on the left. For
MathWorks, the developer of MATLAB took a differ- example, the following are three valid Mathcad assignment
ent approach from Mathsoft. Rather than creating an au- statements.
tomated engineering computation worksheet, MathWorks
chose to create a “high-level” programming language with a: =1
the functions and functionality needed for engineering and a : = a +1
scientific computing. A high-level programming language
F(t) : = 10 ⋅ sin ( 2 ⋅ t )
performs many of the tasks needed to translate an algorithm
expressed as human-readable computer code into the in-
structions required by the computer’s processor to execute The first and third assignment statements also make sense
the program. Computer programs from 1950’s through the as equalities. However, the second statement written as an
early 1980’s were tiny, by today’s standards. Most general equality, a = a + 1, is nonsense.
Mathcad and MATLAB 107

In Mathcad, a conventional equal sign is a command to Upper


evaluate a variable or an expression, and display the result. Limit
The use of an equal sign makes sense in this context, because
the quantities on either side of the equal sign are equal.
Dependent
Plotting in Mathcad Variable
Let us say you wish to plot the result of example Sect. 2.8.1,
v (t ) = 2.66 (1 − e −0.5t ) . You must first define v( t) in the
Mathcad worksheet above, or to the left of where you plan to Lower
place the plot. When you click on an empty area of a Mathcad Limit
worksheet and begin to type, the program presumes that Lower Independent Upper
you are entering an equation. If, in fact, you are typing text, Limit Variable Limit
Mathcad will recognize you are entering text by the space
entered between letters. There are no spaces in equations. Fig. A2.1   A Mathcad x–y plot object showing the black rectangular
When entering an equation, the space bar is used in lieu of place holders. The variables can be scalars, vectors, expressions, or
functions. The limits can be numerical values or expressions
the mouse to move the shaped cursor out of exponents and
denominators. Alternatively, if you wish to enter text, type “
as the first character to create a “text region.” 3
Click in an empty area of a Mathcad worksheet and type 2.658
2.5

v(t) : 2.66*(1 − e ^ −0.5* t ) 2


Space Bar Space Bar v( t)
1.5
Then either type tab, or click outside of the equation object.
1
You will see the following: 0.689
0.5
(
v(t) : = 2.66 ⋅ 1 − e −0.5⋅t ) 0
0
5
t
10
7⋅ τ

Mathcad is case and font sensitive. T, t, t, τ, and t are all


different variables. Greek characters can be entered using (
Fig. A2.2   A Mathcad plot of the function,v (t ) = 2.66 1 − e −0.5t )
the Greek alphabet “pallet” (produced by clicking on the αβ
button), or by typing the combination of the Ctrl key and g
simultaneously, when the cursor is immediately to the right button bar inserts an empty x–y plot into the worksheet. To
of the Roman character one wishes to change to Greek. use the menu, click on the Insert menu, and follow the drop
The duration of a response plot should be six or seven down menu to the item X–Y Plot.
times the largest time constant in the system, Sect. 2.9.4. We
are plotting a first-order response, where the time constant is Insert
the inverse of the magnitude of coefficient in the exponent, Graph
σ = −0.5. Type X–Y Plot.

t Ctrl+g : | 1/ −0.5 An empty x–y plot will display with solid black rectangles

Simultaneously termed “place holders” for the independent and dependent
variables and the axes limits, Fig. A2.1.
Press the Ctrl and g keys simultaneously. Type tab or click To plot v(t) vs. t, enter t in the independent variable
outside the equation object and you will see place holder on the x-axis, 0 in the lower limit x-axis place
holder, 7*t Ctrl + g (to see 7 ⋅ τ) in the upper limit x-axis
1 place holder, and either the function, v(t), or the expression,
τ:=
−0.5 2.66 ⋅ (1 − e −0.5⋅t ), in the dependent place holder on the y-axis.
Clicking outside the region causes Mathcad to display the
To insert an x–y plot into a Mathcad worksheet, first place plot, Fig. A2.2.
the cursor outside an existing Mathcad text or math region. Mathcad autoscales axes, when the limit place holders
One can then either use icons or the menus. The first time are blank. Often, as in this example, autoscaling leads to the
you click on a plot icon, Mathcad displays the plotting pal- display of only a portion of the response. The limits can be
let or button bar. Clicking on the icon for an x–y plot in the edited, after the plot is produced to change either the vertical
108 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Response Function v(t)


2.658 4
2.66
2

velocity, m/sec
v( t) 2
v( t)
− 0.5 t
− 2.66
⋅ ⋅e
1 0
2.66
0
0 −2
0 5 10 −3

0 t 7 ⋅τ 0 5 10
0 t 6 ⋅τ
time, sec
Fig. A2.3   The Mathcad plot of the function, v (t ) = 2.66 1 − e −0.5t , ( )
with the lower limit of the vertical axis edited to read zero −0.5t
Fig. A2.5   The Mathcad plot of the function, v(t ) = 2.66(1 − e ), the
expression, −2.66e , and the constant, 2.66
−0.5t

Response Function v(t)


100
2.658 2.642
0
velocity, m/sec

2 − 100
v( t)
v( t)
− 200
1
− 300
− 392.119
0 − 400
0 − 10 −5 0 5 10
0 5 10
0 t 7 ⋅τ
− 10 t 10

time, sec Fig. A2.6   The Mathcad plot of the function,v(t ) = 2.66(1 − e −0.5t ), auto-
scaled from t = − 10 to t = + 10
Fig. A2.4   The Mathcad plot of the function, v (t ) = 2.66 1 − e ( −0.5t
) , for-
matted adding gridlines, axes labels, and a plot title

in a place holder produces another place holder. For exam-


or horizontal range shown. Clicking on the lower limit of ple, we plot v(t ), −2.66e −0.5t , and 2.66 on the same axis, by
the vertical axis 0.689 introduces the shaped cursor into typing a comma after entering v(t ), and again after entering
that region. Editing 0.689 to read 0, and then either typing −2.66e −0.5t , Fig. A2.5.
tab, or clicking outside the plot region causes Mathcad to We chose to set the limits of the time axis at zero and six
re-evaluate and display the plot, Fig. A2.3. or seven time constants. If we create a plot by editing the
Plot regions can be dragged around the worksheet, and place holders for the independent and dependent variables,
the frame of a plot region can be dragged to resize the plot. but leave the axes limits blank, Mathcad will autoscale the
Plots can also be “formatted” to add grid lines, change the horizontal axis from −10 to +10, and evaluate the dependent
grid spacing, and to change the width, color, and type of line, variable within those limits. Since the input was applied at
Fig. A2.4. Right clicking within a plot region brings up a time, t  = 0, the plot will show a response before the input
context-sensitive menu which includes the three items, For- acted on the system, Fig. A2.6. Even though we did not in-
mat…, Trace…, and Zoom… The format dialog box has five tend for the function to be evaluated for negative time, it
tabs which are self-explanatory, with the exception of the can be. The polite term for the resulting plot is “non-causal”
“secondary Y-axis.” Clicking the check box “Enable second since it violates cause and effect. The more common terms
Y-axis” produces a set of place holders on the right side of include nonsense and garbage.
the plot, the middle place holder is for a second dependent Do not plot negative time, unless (1) the input is applied
variable, and the other two are the limits of the secondary Y- at the negative time of the lower limit, or (2) the response
axis. We will find a secondary Y-axis useful, since the power function is multiplied by the Heaviside unit step function to
variables of energetic systems have different units and, im- zero out the response function, until the time the input is ap-
portantly, different magnitudes. Plotting the responses of plied. Mathcad’s notation for the Heaviside unit step func-
two different power variables on one axis can lead to one tion is capital phi, Φ( ). The f(x) button brings up a dialog
response appearing flat, since its vertical range is misscaled. with all of Mathcad’s built-in functions. The Heaviside unit
When we wish to plot two or more “traces” on the same step is in the “Piecewise Continuous” submenu, or can be
axis, typing a comma after entering a variable or expression found in the alphabetical list. Multiplying the response func-
Mathcad and MATLAB 109

3 Response Function v2(t)


3
4.5
2 4

Φ( t) ⋅ v( t) 1 3

v, m/sec
v.2( t)
0 2

−1
−1 1
− 10 −5 0 5 10
0.013
− 10 t 10 0
0 10 20 30
0 t 6⋅ τ.2
Fig. A2.7   The Mathcad plot of the product of the response function, t, sec
v( t), and the Heaviside step function, us( t). Mathcad’s notation for the
Heaviside step function is φ (t) Fig. A2.8   The Mathcad plot of the response function, v2( t), formatted
with gridlines, axes labels, and a plot title

tion by the Heaviside unit step function zeros out the value of
the response function, until the moment when the Heaviside Mathcad permits assignment statements to be evaluated.
step function transitions from zero to one, Fig. A2.7. Type
As an example of a second-order oscillatory step re-
t Ctrl+g .2 : | 1/ −0.16 =
sponse, we will plot the result of Sect. 2.7.3 
Simultaneously

to see
v (t ) = 2.74e −0.16t cos ( 0.68t − 2.91) + 2.66
1
τ2 : = = 6.25
Recall the exponent of the real exponential is σ, the real com- −0.16
ponent of the eigenvalues of the system, and the frequency
ω is the magnitude of the imaginary component of the ei- Create a plot by clicking the X–Y Plot button, which should
genvalues. be visible in both the button bar below the menus and in the
Graph pallet. Get into the habit of entering the independent
s1 , s2 = σ ± jw → s1 , s2 = −0.16 ± j 0.68 variable, t, and its limits, before entering the function name
or expression as the dependent variable. Reversing the order
Create an assignment statement defining the response vari- leads to Mathcad trying to be helpful and autoscaling using
able, v2( t). Note the subscript 2, which is part of the func- its standard range of − 10 to 10, which is rarely the range we
tion’s name. Mathcad refers to a subscript which is part of will want. Format the plot, adding gridlines, axes labels, and
a variable of function name as a “literal” subscript, to dis- a plot title, Fig. A2.8.
tinguish it from a subscript which represents the index of a We can reuse function and variable names. The assign-
vector. A literal subscript is created by typing a period im- ment operator := is a “local” assignment, meaning that it can
mediately before the literal subscript. A “vector” subscript or be overwritten by a new assignment operator which appears
index is created by typing a left square bracket [ immediately to its left or below it in the worksheet. If we reuse a vari-
before the subscript. Type

v.2(t) : 2.74*e ^ −0.5* t Space Bar Space Bar *cos(0.68* t − 2.91) + 2.66

to see
able name, Mathcad underlines it with a green squiggle, to
v 2 ( t ) : = 2.74 ⋅ e −0.16⋅t ⋅ cos ( 0.68 ⋅ t − 2.91) + 2.66 alert us, in case we thought the variable name was unique.
Mathcad handles units as if they were variables, and has vir-
The time constant, which scales the duration of the plot, is tually every engineering unit predefined. Click on the mea-
the time constant of the decay envelope. The upper limit of suring cup symbol to bring up the unit dialog box. Many
the time axis should be six or seven τ, where common choices for variable names are predefined units.
Consequently, variables may be underlined with a green
1 1 squiggle, even though they are unique, because they are also
τ= → τ= → τ = 43.75 sec ≈ 44 sec
σ −0.16 the abbreviation of a unit.
110 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Mathcad permits mixed units in calculations. Mathcad the MATLAB’s default variable type is an “array.” An array
converts all units to SI prior to computation, and then pres- is an ordered set of data where an individual element is iden-
ents the results in SI (or sometimes metric) but adds a blank tified by an index. The indices locate elements in an array
placeholder. If the user enters a unit into the placeholder, as if they were Cartesian coordinates. A one-dimensional
Mathcad recomputes and expresses the result those units. array, also called a vector, because it resembles mathemati-
For example, type in the following volume computation, cal vectors, is a sequence where an element is identified by
where the three lengths are expressed in inches, feet, and a single index. Two-dimensional arrays, which resemble ma-
centimeters. trices, require two indices to identify an element. In standard
mathematical notation, the index of an element in a vector or
5*in *0.6*ft *14*cm = matrix would be subscripted. There are no subscripts, super-
You will see, scripts, or Greek characters in computer code, only the key-
board characters. Array indices are contained in parentheses.
5 ⋅ in ⋅ 0.6 ⋅ ft ⋅14 ⋅ cm = 3.252 L � For example, element 14 of the vector y is written as y(14).
A vector index may be an expression, such as y(n + 1). The
Edit the place holder. Type in^3. You will see values of vector indices must be positive integers. A restric-
tion of MATLAB is that the smallest vector index is one, not
5 ⋅ in ⋅ 0.6 ⋅ ft ⋅14 ⋅ cm = 198.425 in 3 zero.

Assignment Statement
Plotting in MATLAB The assignment operator in MATLAB is a conventional
equal sign, = . Assignment statements are read from right to
MATLAB’s Environment left. The right side can be a numerical value, a previously
MATLAB is a programming “environment.” The default defined variable, or an expression. The left side must be a
configuration opens with a tabbed tool bar across the top of variable.
the screen and five windows below it, with the Command
Window in the center. MATLAB is an “interpreter” which For-End Loop
means ­MATLAB translates and executes code line by line. MATLAB provides a number of “flow-control” instructions,
The practical effect is that the “command line,” identified or commands, which allow the execution sequence of repeat
by “prompt” > > in the Command Window, can be used as a or skip blocks of code. We will use a logic structure called
calculator. For example, typing 2 + 4 Enter at the command a for-end loop which, as one might guess, begins with the
prompt yields word for on the first line of the code block and ends with the
word end on the last. A variable which serves as a counter
 > > 2 + 4 is defined in the for line and given an integer range. The
ans = counter variable consecutively assumes values of a defined
6 range. The lines of code between the for and end lines are
repeatedly executed, until each counter variable defined in
MATLAB keeps a record of the commands and the variables the for line and incremented by one passes through, and the
used in calculations and scripts. The Command History win- loop reaches its upper limit.
dow shows a history of commands entered which extends Example. We will wish to repeat the calculation of a re-
to prior uses of MATLAB. It is convenient for an individual sponse function, perhaps one thousand times. We choose a
using MATLAB on a personal machine, but it also means name, say n, for the counter variable. We can use the counter
that users of a public machine have access to the prior user’s variable in the code between the for and end statements, but
command history. The command history can be cleared by we must not change its value. MATLAB will increment the
right clicking on the title bar at the top of the Command His- value of the counter variable, and check it against the upper
tory window to bring up a context menu and selecting the limit of the range. The syntax of a for-end loop that executes
item “Clear Command History.” 10 times is

Array Variables for n=1:10


The record of the variables used is shown in the window instruction;
titled, Workspace, with three columns, Name, Value, and instruction;
Min. Note that MATLAB created the variable named ans and
displays its Values and Min as 6. The icon at the left of row instruction;
is a square subdivided into four squares. This is a clue that end
Mathcad and MATLAB 111

We can create a vector variable and assign it values using a sonal storage device or location. Click on the green triangu-
for-end loop. Say we need a time vector t from t = 0 to t = 1.0 lar Run icon in the tool bar at the top. A dialog will appear
with the time step or increment ∆t = 0.1. How many ele- stating that the script is not in the current folder or on the
ments are in the vector t? There are 11. The common error is MATLAB path. Click on the button, Add to Path, which adds
to divide the interval by the increment, your personal storage location to those locations that MAT-
LAB checks for scripts. It will also run the script. Nothing
tend − tstart 1.0 − 0 appears to happen, because you are still in the Editor. Navi-
= = 10 Intervals
∆t 0.1 gate to the MATLAB environment. You will see the result in
the Command Window. The elements of vector t are written
This is the correct number of intervals of one-tenth between to the Command Window for each iteration through the loop.
zero and one, but we are short of one value needed to create Notice that t is now listed in the Workspace window, and
the end point for the last interval. The calculation of interval has the Value, < 1 × 11 double >, describing it as a one-dimen-
end points is sional array with 11 elements of a data-type double. Data
type refers to how data are stored. Double refers to “double
tend − tstart 1.0 − 0 precision floating point” number. Double precision uses 64
+1 = + 1 = 11 Interval End Points
∆t 0.1 bits of computer memory. Floating point refers to scientific
notation.

MATLAB’s Editor Plotting


We will use MATLAB’s editor to write a “script” or program We will now edit First.m to create a second vector y and
to create the vector t., stored with the extension.m, and then plot it. Although MATLAB has pi as a constant, it does not
run at the command line by typing the file’s name. Search as have Euler’s number e. Exponentiation to the base e is per-
you may, you will not find a tab or window for MATLAB’s formed with the function exp(). MATLAB’s function, plot(),
editor from the default environment. The editor is opened, produces an interactive plot which can be resized, formatted,
when you choose the icon, “New Script,” in the Home tool and “interrogated” using the “data cursor.” The arguments of
strip tab. The Editor opens as a second program. The only plot() are pairs of vectors which contain the x and y-axes co-
way to navigate from the Editor back to the MATLAB en- ordinates of the data plotted, e.g., plot(x, y). In our example,
vironment is by clicking on the MATLAB icons in the Win- the independent variable is the vector t. Add the plot com-
dows task bar, and selecting the M ­ ATLAB environment. mand after the end statement.
Comments are labels, explanations, and notes added to Our output is now graphical in the form of a plot. We do
computer code for the programmer’s and future user’s ben- not need to see the values of t and y written to the screen. In
efit. The longer and more complicated the code, the more fact, writing each iteration of the loop to the screen slows
important are comments for structuring the program and execution of the script significantly. A semicolon at the end
making it understandable. Comments begin with a percent of line of code is an instruction not to write that line on the
sign % and are colored green in ­MATLAB’s editor. A com- screen.
ment may occupy the end of a line of code, or may be a line
by itself. Comments which identify variables and describe % First.m  06-02-14
the function of blocks of code are essential. The purpose of for n=1:11;
writing a script is to save time by automating tasks. Poorly t(n)=0.1*(n-1);
commented code is difficult to understand and use, and de- y(n)=exp(-t(n));
feats the purpose. end;
We will begin our script with a comment giving the name plot(t,y)
of the script, First.m, and the date. MATLAB names can-
not have spaces. Use an underscore instead. Our simple pro- Save and run this script. The MATLAB plot titled Figure 1
gram is will appear as a new document, Fig A2.9. If Figure 1 does
not exist, it will be displayed on top of the Editor. If Figure 1
% First.m  06-02-14 already exists, it will be overwritten with the current plot but
for n=1:11 will not be brought to the front. Navigate to it by clicking
t(n)=0.1*(n-1) on the MATLAB icons in the task bar at the bottom of the
end screen.
MATLAB figures can be formatted after they are created,
Write and save the script. Although you can accept MAT- by opening the “Tools” menu item in the menu bar of the
LAB’s default location, it is best to save your script to a per- figure, and choosing Edit Plot. It is also possible to format
112 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

1 1

0.9 0.9

0.8
0.8
0.7
0.7
0.6
0.6
0.5
0.5
0.4

0.4
0.3

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Fig. A2.9   Plot of the MATLAB script, First.m Fig. A2.10   Plot of the MATLAB script, Second. m, showing a kink
due to the final two values, which were retained in the vectors, t and y,
from the previous execution of a script, First.m
figures with commands in the script. Although more com-
plicated, it allows the formatting to be automated. The tech-
nique is addressed in the Programming in MATLAB appen- The plot, Fig. A2.10, is correct from t = 0 to t = 0.8, corre-
dix to Chap. 8. sponding to the portions of the vectors, t and y, which were
overwritten with new values. The final two elements of vec-
The Dark Side of MATLAB’s Workspace tors t and y were not overwritten, and still hold values cre-
There is a dark side to MATLAB’s workspace. Variables re- ated by the previous script First.m, leading to the kink in the
main in the workspace, until they are deleted, by selecting curve. Although this example may seem contrived, it is not.
the variable in the Workspace window, right-clicking, and Engineers tend to use and reuse the same meaningful vari-
choosing delete, or the entire workspace is cleared by select- able names. MATLAB does not create a new instance of an
ing the title bar of the Workspace window, right clicking, and existing variable. Consequently, when MATLAB executes
choosing Clear Workspace. The existence of variables in the an error, or the results of successful execution of a script do
workspace can mask errors and omissions in scripts. For ex- not make sense, clear the workspace and run the script again.
ample, say you write a script to calculate a response function.
You intend to define a variable named tau equal to a certain Plotting Flow Chart and Script
value, but you simply forget and omit that line of code. If you We now have the basics to write a MATLAB script to plot
run the script and tau was not used in any previous calcula- response functions. We must develop the logic to establish
tions and, consequently, is not in the workspace, MATLAB the duration of the plot and the time step of the calculation.
identifies the error. A ding sounds, and the Command win- The variable is tau_max, the largest time constant of the
dows displays an error message indicating that an undefined system. tchar is the smallest characteristic time of the sys-
variable was used in line x of the script. On the other hand, tem, which may be either a time constant, or, if the response
if you omit the definition of tau in the script, but tau exists in is oscillatory, a period of oscillation T, Sect. 2.9.4. The time
the workspace with a different value, MATLAB will use the step, dt, is 1/200 of tchar. The duration of the calculation,
workspace value. The script will execute but produce an er- tend, should be approximately six or seven tau_max. The
roneous result. number of calculations for the plot N is
Minor editing of the script, First.m, will provide a
graphical example of the effect of MATLAB’s workspace. tend
Second.m uses the same variables, named t and y, as First.m. N=
dt
The upper limit of the for-end loop is reduced to 9 from 11.
Finally, the argument of exp() is multiplied by two, increas-
ing the exponential decay rate. The for-end loop is similar, and the plot() command is identi-
cal to scripts, First.m and Second.m. Figs. A2.11 and A2.12
% Second.m  06-02-14 are flow charts of the script.
for n=1:9; We cannot name the script, plot.m, because plot () is a
t(n)=0.1*(n-1); MATLAB function, so we will name it Plotting.m. Because
y(n)=exp(-2*t(n)); we must edit the script each time we wish to plot a new re-
end; sponse function, it is helpful to uniquely identify each ver-
plot(t,y) sion of the script with a date code, i.e., plotting_060314.m.
The plot produced by the code, Fig. A2.13, is then formatted
and annotated using MATLAB's plot editing features.
Mathcad and MATLAB 113

Fig. A2.11   Upper half of the


flowchart of the MATLAB Start
script, Plotting.m

Calculate maximum time constant from the


minimum magnitude real eigenvalue component
sigma_min = mininimum |σ|
tau_max = 1/abs(sigma_min)

Calculate the plot duration as 6 or 7 tau_max


tend = 6*tau_max

No Oscillatory
System?

Yes

Calculate the smallest time constant from the Calculate the characteristic time tchar as the
maximum magnitude real eignevalue component period of the highest frequency oscillation
sigma_max = maximum |σ| omega = ω
tchar = 1/abs(sigma_max) tchar = (2*pi)/omega

Calculate the time step dt

dt = tchar/200

Continued in Fig. A2.12

% plotting_060214A.m % system. If the smallest characteristic time is a


% Response Function Plotting Script 6-2-14 % period, use the following block to calculate the
% % period from an angular frequency in radians/sec.
% tau_max is the largest time constant in the system.
% If the eigenvalues are known then use the smallest % omega = 0.68
% magnitude real component sigma. % tchar = 2*pi/omega

sigma= -0.5 % If the smallest characteristic time is a time


tau_max = abs(1/sigma) % constant then comment-out the two lines above and
% use the following line.
% tend is the plot duration. Should be 6 or 7 tau_max.
tchar = tau_min
tend = 6*tau_max
% dt is the time step between calculations. Use 1/200
% tchar is the smallest characteristic time of the % of the smallest characteristic time.
114 2  Differential Equations, Input Functions, Complex Exponentials, and Transfer Functions

Continued from Fig. A2.11 Response v(t)


3

v, m/sec
Calculate the number of interations of for-end loop
N = tend/dt 1

0
0 2 4 6 8 10 12
Initialize for-end loop counter variable
t, sec
n=1
Fig. A2.13  The MATLAB plot of the response function, v (t), from the
script, plotting_060214A.m, after formatting

plot(t,y)
Compute time vector element
t(n) = (n-1)*dt % plotting_060214B.m
% Response Function Plotting Script 6-2-14
%
% tau_max is the largest time constant in the system.
Compute output vector element
% If the eigenvalues are known then use the smallest
y(n) = response(t(n))
% magnitude real component sigma.

sigma = -0.16
tau_max = abs(1/sigma)
Increment counter by one No
n == N ?
n=n+1 % tend is the plot duration. Should be 6 or 7 tau_max.

tend = 6*tau_max
Yes

% tchar is the smallest characteristic time of the


Plot vector pairs
% system. If the smallest characteristic time is a
plot(t,y)
% period, use the following block to calculate the
% period from an angular frequency in radians/sec.

omega = 0.68
Finish tchar = 2*pi/omega

% If the smallest characteristic time is a time


Fig. A2.12   Lower half of the flowchart of the MATLAB script,
plotting.m % constant then comment-out the two lines above
and use
% the following line.
dt = tchar/200
%tchar = TimeConstant
% Calculate the upper limit of the for-end loop. This
% value is the number of calculations. % dt is the time step between calculations. Use 1/200
% of the smallest characteristic time.
N = tend/dt
dt = tchar/200

for n=1:N % Calculate the number of calculations N. This is the


t(n)=(n-1)*dt; % upper limit of the for-end loop.
y(n) = 2.66*(1-exp(-0.5*t(n)));
end;
References and Suggested Reading 115

Response v2(t) 3
Response Plot with Homogeneous and Particular Solutions
4.5

4
2

1
v2(t), m/sec

v, m/sec
0
2

−1
1
−2

0
0 10 20 30 40 −3
t, sec 0 2 4 6 8 10 12
t, sec

Fig. A2.14  The MATLAB plot of the response function, v2(t), from the
Fig. A2.15  The MATLAB plot of the response function, v1(t), −2.66−0.5t
script, plotting_060214B.m, after formatting
and 2.66, from the script, plotting_060214C.m, after formatting

N = tend/dt % If the smallest characteristic time is a time


% constant then comment-out the two lines above
for n=1:N % and use the following line.
t(n)=(n-1)*dt;
y(n)=2.74*exp(-0.16*t(n))*cos(0.68*t(n)-2.91)+2.66; 2 tchar = tau_min
end;
% dt is the time step between calculations. Use 1/200
plot(t,y) % of the smallest characteristic time.

Plotting additional traces on the y-axis requires the cre- dt = tchar/200


ation of additional output vectors, and then adding pairs of
independent, dependent variable vectors to the plot state- % Calculate the upper limit of the for-end loop. This
ment. We will rename the script, plotting_060314A.m, % value is the number of calculations.
as the following: plotting_060314C.m, and edit it to plot,
v (t ) = 2.66(1 − e −0.5t ), −2.66e −0.5t and 2.66 on the vertical N = tend/dt
axis. The revisions are in the for-end loop and in the com-
mand plot (). The resulting plot is Fig. A2.14. for n=1:N
t(n)=(n-1)*dt;
% plotting_060214C.m y1(n) = 2.66*(1-exp(-0.5*t(n)));
% Response Function Plotting Script 6-2-14 y2(n) = 2.66*(exp(-0.5*t(n)));
% y3(n) = 2.66;
% tau_max is the largest time constant in the system. end;
% If the eigenvalues are known then use the smallest
% magnitude real component sigma. plot(t,y1,t,y2,t,y3)

sigma= -0.5
tau_max = abs(1/sigma) References and Suggested Reading

% tend is the plot duration. Should be 6 or 7 tau_max. Hildebrand FB (1976) Advanced calculus for applications, 2nd edn.
Prentice-Hall, Englewood Cliffs
Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
tend = 6*tau_max Cliffs
Ogata K (2009) Modern control engineering, 5th edn. Prentice-Hall,
% tchar is the smallest characteristic time of the Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
% system. If the smallest characteristic time is a Prentice-Hall, Upper Saddle River
% period, use the following block to calculate the Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
% period from an angular frequency in radians/sec. dynamics. Addison-­Wesley, Reading

% omega = 0.68
% tchar = 2*pi/omega
Introduction to the Linear Graph Method,
Step Responses, and Superposition 3

Abstract
System dynamics predicts the responses of physical systems to inputs of energy. In this
chapter, we examine the response of first-order systems to step changes of the input power
variable. The response of a system to a step input, called the system’s “step response,” is
both common and important. It is common because many sources provide a reasonably
constant value of the input variable, if power draw is not too large. The importance of the
step response is twofold. First, the step response reveals the homogeneous response of the
system. We can experimentally determine the elemental parameters of the system’s step
response. Second, the superposition (or summing) of steps inputs of different amplitudes
and shifted in time allows us to approximate any arbitrary input. Superposition is then used
to calculate the response of a system to that input by summing the individual responses to
each step.

3.1 Introduction step, and sinusoidal inputs to be extended to arbitrary inputs


by using superposition. Lastly, and most importantly, it al-
We revisited differential equations in Chap. 2 and thoroughly lows much of the content of this course to be used in engi-
reviewed the method of undetermined coefficients. In this neering practice.
chapter, we will apply a decidedly engineering perspective
to the solution of first-order differential equations with step
inputs. The approach we develop will support the engineer- 3.2  Introduction to the Linear Graph Method
ing design process. Specifically, our approach will allow us
to make relatively quick decisions using simple models, be- Chapter 1 introduced the definitions of mechanical work and
fore we commit ourselves to the time and expense of a more power, the concepts of mechanical and energetic models,
accurate but involved mathematical model. We will focus on and the techniques and notation needed to describe power
the relationship between the input to the system and the re- flow through an energetic network. We will now develop
sponse of the system. From this perspective, the differential these basics into the linear graph method. A linear graph is
system equation is seen as an “operator,” where the system a circuit-like representation of an energetic system drawn
equation operates on the input variable to yield the output from a schematic representation of an energetic model. The
variable, Fig. 3.1. Although we use the term “input variable” technique of drawing a linear graph from the schematic of an
in order to predict the behavior of a physical system, we need energetic system is straight forward. We shall see in Chaps. 4
to know how the input variable changes with time. In other and 5 that the same methodology is applied for mechanical,
words, we need the input function, also known as the forcing fluid, thermal, and electrical systems. The greatest benefit
function, to calculate the response function. of a linear graph is the ease of joining dissimilar subsystems
This perspective is attractive for three reasons. First, it into a single energetic system which will allow us to model
simplifies the solution of the differential system equations in machines, as we shall see in Chap. 6.
many cases. Second, it allows the solution of impulse, pulse,

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_3, © Springer Science+Business Media New York 2014 117
118 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Input Variable Dynamic Output Variable


x,v
Forcing Function System Response Function
Surface 1
Fig. 3.1   A dynamic system “operates” on an input yielding a response. v1
The equation which describes the response is the response function
h Fluid Shear
v2
x,v Surface 2

F(t) Mass Fig. 3.3   Shear of a fluid between two planar surfaces separated by
Lubricating
M fluid film distance, h, moving with horizontal velocities, v1 and v2
Damping b
e­ nergetic attribute of the physical system. In physical reality,
an object with mass has additional properties or attributes,
Fig. 3.2   Schematic energetic model of a force source, F( t), sliding a such as volume, compliance, etc. All physically real objects
mass M on a lubricating film with damping b
have multiple energetic attributes. However, an element
in an energetic model has just one energetic attribute. The
We will introduce the linear graph method with the ex- consequences of a model element having a single energetic
ample of a force source sliding a mass on a lubricating fluid property are contrary to our everyday experience with real
film, Fig. 3.2. The lubricant film supports the mass to prevent physical objects. A mass element in an energetic model can
it from making solid to solid contact with the surface. In order only store kinetic energy. It cannot store strain energy nor
for the mass to slide, the fluid must shear, which dissipates can it dissipate energy as heat through inelastic deformation.
mechanical power, thus heating the fluid. This phenomenon of Consequently, a mass element is ideally rigid. It cannot de-
“viscous friction” is called damping, and is represented by the form. Likewise, a fluid which shears and dissipates energy as
parameter b. We will present the basics of the viscous shear of heat can only dissipate energy. It cannot store energy. Conse-
fluids, as we examine the energetic properties of the model. quently, it has no mass and is incompressible.
If a physical object has two significant energetic attri-
butes, then two energetic elements are needed to model the
3.2.1  Energetic Model one physical object. Judging the significance of various en-
ergetic properties of a single physical object or system must
First, it is important to note that an energetic model need not be deferred for now, since it requires interpretation of the re-
resemble the physical object in the slightest. The energetic sults of the modeling process. As a rule of thumb, for the first
model shown in Fig. 3.2 can represent any number of physi- and simplest model, an attribute is “significant” and should
cal systems of vastly different appearance, size, and func- be included in the model if it stores or dissipates at least 10 %
tion, such as a fuel injector for an automobile or a massive as much energy or power as any other similar element in
table of numerically controlled machine tools. Appearance the model. Any model should follow the “KISS” principle,
is not a physical attribute we wish to represent in our model, where KISS stands for Keep It Simple, Stupid. Modeling
except in those cases, when the geometry of a device affects should always progress from simple to more complicated.
its energetic behavior. The size of a system is conveyed by It is possible to add elements to a model, if the preliminary
the magnitudes of the system’s energetic parameters, M and results indicate it is necessary.
b, which need not be established, until the model is actually We shall see that energy dissipation in mechanical sys-
used to predict the response of a specific physical system. tems is difficult to describe mathematically. The mass of
An energetic model is an “abstract” model. Engineering a solid object can be represented within the accuracy with
students are intimidated by the term, abstract, since it is asso- which we can determine the object’s mass. Conversely, all
ciated with intellectually challenging aspects of mathematics descriptions of friction are at best approximations of the
and physics which would more correctly be titled, “arcane,” actual energy dissipation, and are invariably uncertain to a
rather than abstract. A more common usage of abstract de- greater degree than other energetic properties.
notes “greatly simplified.” The abstract of this chapter is a The fluid film provides lubrication, that is, supports the
summary of the main points. mass preventing contact between the two solid surfaces. For
A difficult aspect of modeling for most students is the the mass to slide horizontally, the fluid must shear, Fig. 3.3,
fact that an element in an energetic model represents a single and that shear dissipates energy in “viscous” friction.
3.2  Introduction to the Linear Graph Method 119

­Figure  3.3 shows the gap height to be h. In reality, the height a b ∆x


of the gap, h, between the mass and the supporting surface
τ yx
varies with the velocity of the mass, increasing with increas-
ing translational speed. Also, there is a minimum velocity, y y
below which the “asperities” (or peaks) on the solid surface
of the mass would contact the asperities on the supporting τxy x τxy y x
surface, creating Coulomb friction. We will model the sys-
tem as having a constant gap height, h. The value we would
use for h would be that which we believe is representative τyx
of our system’s “operating” condition. In other words, a gap
Fig. 3.4  a Shear stresses, τxy and τyx, on the surfaces of a cubic element
height should be typical of that which occurs, when the mass
of fluid. b Shear in the x–y plane of a fluid cube
is sliding in an expected range of speed.
The conventional fluid mechanics model for liquid flow
assumes that the liquid in direct contact with a solid surface much with) temperature, Eq. 3.4 could be rewritten with the
“sticks” to that surface. This is known as the “no-slip” as- shear rate raised to the power n:
sumption, and results in the fluid in contact with a given n
 dγ  (3.5)
surface having the velocity of that surface. Macroscopic τ = µ0 
  dt 
shear displacement deforms the fluid through shear strain,
γ , which is given as where µ0 and n are constants. This is a common form of a
non-linear relationship between shear rate and shear stress
∆x in fluids.
γ =
(3.1)
y When viscosity changes little over the anticipated rang-
es of shear rate and temperature which the fluid will expe-
where ∆ x is the difference in the horizontal displacement of rience, then viscosity can be modeled as an ideal, constant
parallel surfaces of a cube of fluid, and y is the length of the viscosity. The shear rate and shear stress are then linearly
side of the cube and, thus, the distance between the cube’s related, Eq. 3.6,
parallel faces, Fig. 3.4.
The shear strain rate is the time derivative of shear strain: dγ
τ=µ
(3.6)
dt
dγ 1  ∆ x  1 dx
= lim
(3.2) =
dt ∆t → 0 ∆t  y  y dt A fluid modeled as having a constant viscosity is known as
a Newtonian fluid, and yields the linear velocity distribution
If the shear strain is uniform across the height h of the gap, shown in Fig. 3.3. We will assume an ideal, constant viscos-
then dx / dt = v and the shear strain rate is ity, that is, Newtonian fluid, as the lubricant, which is the
simplest model of the fluid mechanical effects.
dγ v The free body diagram of the forces acting on the mass is
=
(3.3)
dt y shown in Fig. 3.5. The shear stress on the bottom surface of
the mass opposes the motion of the mass. If the input force,
The shear stresses within the fluid, Fig. 3.4, are related to the F( t), acts in the positive direction, then the viscous friction
strain rate of the fluid by the parameter viscosity given the shear force, Fshear, acts in the negative direction. The magni-
symbol, µ. The viscosity of most common liquids is a func- tude of the shear force, Fshear, is integral of the shear stress,
tion of strain rate and temperature. Therefore, in the general τyx, over the bottom area of the mass. Similarly, integration
case, the relationship between shear stress and strain rate of of the fluid pressure across the bottom of the mass yields
a liquid is a force equal to weight, W, of the mass acting to support
the mass, FR. Notice that F( t) and Fshear form a couple that
  dγ  dγ
τ = µ ,T  (3.4) rotates the mass about the z-axis. This rotation cannot hap-
 dt  dt pen without violating the geometry we have assumed for
our model. What we have neglected is the variation of gap
Equation 3.4 is non-linear, because the viscosity µ is a func- height h and the pressure acting on the mass with position.
tion of the shear rate, d γ / dt . It is also not “stationary,” mean- The variation in h and pressure would be the result of a more
ing that the equation does not remain the same. In this case, rigorous fluid mechanics model of the motion of the mass on
it changes with temperature. If fluid temperature was held a fluid film with constant viscosity. A fluid mechanics model
constant, or the viscosity was “insensitive” to (did not vary would reveal that, if the mass moved fast enough, the rota-
120 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

If the heat dissipated in the fluid is removed quickly, by


F(t) removing either the heated fluid, or just the heat, then the
M
y W temperature increase will be minimized. If heat accumulates
in the fluid, then the temperature increase will be maximized.
x
Hence, for a given fluid, the magnitude of the effect of the
FShear energy that is dissipated in the fluid on the fluid’s viscosity
FR
depends on these following factors: the thermal properties
of the fluid and the solid surfaces; whether it is an “open”
Fig. 3.5   Free body diagram of the forces acting on mass M
or “closed” fluid system (does the fluid recirculate or does
new fluid enter and heated fluid leave?); and the operating
x,v
conditions. Thermal systems and heat transfer are discussed
in Chap. 5.
F(t) Mass There are two ways to include a more accurate model
of viscosity in our energetic model. We can maintain a
M
linear model by approximating a non-linear function. The
method, known as “linearizing about an operating point,”
FShear is developed in Sect. 4.2.4.1, when we revisit damping in
Chap. 4. The second approach is to formulate a more ac-
Fig. 3.6   Free body diagram showing only the forces acting on mass M curate non-linear description of viscosity, such as Eq. 3.4
that do work
or 3.5. Non-linear models yield non-linear differential sys-
tem equations, which have a drawback, although they pres-
tion would be clockwise, opposite of the couple shown on ent more realistic descriptions of dynamic systems. Linear
the free body diagram, leading to a gap height that decreased system equations can be solved by a variety of classical
from the leading side to the trailing side of the mass. We mathematical techniques. Conversely, non-linear differen-
are constructing, however, an energetic model of the system. tial system equations must be solved by “numerical meth-
We need not include these details, unless they have a sig- ods,” which approximate differential equations as “finite
nificant effect on the energetic behavior of the system. We differences.” Numerical methods, discussed in Chap. 8, are
will neglect the rotation of the mass. We will examine how widely used in engineering, but each analysis is its own
to couple translational and rotational motions in mechanical special case. The first model of any system should be a
systems when we develop “transducers” in Chap. 6. simple linear model. If a non-linear model is found to be
We can further simplify the free body diagram by elimi- necessary, then the simple linear model provides the start-
nating forces or force components which do no work on the ing point for the more complicated non-linear model. Since
system. In this model, motion is restricted to the x-axis. The we are developing a simple, first model of the mass sliding
weight W of mass M and reaction force FR created by the on a lubricating fluid, we will ignore thermal effects. In
fluid pressure act in the y-direction. They have no compo- engineering jargon, the thermal effects on the viscosity of
nent force in the x-direction and, consequently, they do no the lubricant will be “unmodeled dynamics.”
work on the system. We can eliminate them from the free Deciding to use a constant viscosity µ and a uniform gap
body diagram of the model, yielding the simplified free body height h establishes the energetic model of shearing the lu-
diagram shown in Fig. 3.6. bricating film. The shear stress, τyx, is described by Eq. 3.6
Although we can dismiss the details of the results of fluid and is uniform over area A of the bottom of the mass, since h
mechanics, which contradict the geometric constraints of our is constant. The shear force, Fshear, is
simple model, we cannot be as cavalier with any energetic ef-
fects, since we are creating an energetic model. The fluid lu- Aµ dx Aµ
(3.7)
Fshear = = v
bricant in the model, shown in Fig. 3.2, has a single energetic h dt h
property, damping; it dissipates the mass’ energy of motion
as heat in the fluid. If the heat dissipated remains in the fluid where the velocity of mass v is also expressed as dx/dt.
and enough heat is dissipated to raise the fluid’s temperature The Newtonian formulation of the energetic model uses
and decrease the fluid’s viscosity, making it easier to shear, dx/dt. The formulation derived using linear graph method
then the damping property or coefficient b of the fluid would uses v.
change with motion of the mass. Energetic parameters, such Damping coefficient b is the parameter which relates the
as the fluid’s damping coefficient b, must be constant if the velocity difference across the damping element, the fluid
system model is to yield a linear differential equation. film in this system, and the force which creates that velocity
3.2  Introduction to the Linear Graph Method 121

difference. In this model, Fshear creates velocity difference v with the input variable, F( t), and the dependent or output
across the fluid film, variable, x, on opposite sides of the equation. Rewriting the
derivative terms in the order of decreasing order, and then,
Aµ Aµ for standard form, clearing the coefficient of the highest-order
Fshear =
(3.8) v = bv → b =
h h derivative of the output variable x yield the Newtonian form
of the system equation for the displacement of the mass:

d2x dx 1 d 2 x b dx
F (t ) = M 2
+b → F (t ) = 2 + System Equation
dt dt M dt M dt

Collecting the effect of viscous friction, which is distrib- This form of system equation is a second-order differen-
uted across the bottom surface of the mass, into damping tial equation. It can be rewritten in terms of velocity, using
coefficient b is an abstraction or simplification. A single the definition v = dx dt , to yield a first-order differential
parameter that represents a property distributed across a equation:
surface, such as viscous friction, or throughout a volume,
such as mass, is called a “lumped” parameter. This is a fa- 1 dv b
F (t ) = + v System Equation
miliar approximation, since free body diagrams also lump M dt M
or localize properties. Collecting or concentrating proper-
ties which are actually distributed simplifies the resulting In Sect. 3.5, we will introduce a special form for the first-
differential system equation by eliminating the spatial vari- order system equations known as the time constant form,
able, x. The system equation is then an ordinary differential where the output variable term is cleared of a coefficient,
equation with time derivatives, rather than a partial differ-
ential equation with derivatives of both time and position. M dv
bF (t ) = + v System Equation in Time Constant Form
b dt

Time
Constant

3.2.2 Newtonian Formulation of the


Force-Mass-Damper Model The coefficient M/b multiplied by dv/dt must have units of
time for the units to be consistent. It is known as the time
A Newtonian formulation of an energetic model is derived constant and given the symbol τ, Greek for t.
in terms of the dependent variables, translational displace-
ment x, and rotational displacement θ. A free body diagram is
drawn, showing the forces and torques acting on each mass, 3.2.3 Linear Graph Formulation of the
and each mass moment of inertia. The sum of the forces act- Force-Mass-Damper Model
ing on a mass accelerates the mass in translation. The sum
of the torques accelerates the mass’ mass moment of inertia. A “linear graph” is a network which represents the energetic
We have restricted the motion in the mass-damper model to properties of a dynamic system. A linear graph is drawn from
translate in the x-direction, so we do not need to sum torques an energetic model usually represented as a schematic. The
about any axis or forces in the y or z directions. Referring schematic indicates both graphically and with “call outs”
to the free body diagram, Fig. 3.6, which shows only the x- (text) every significant energetic property in the dynamic
direction forces which do the work, the sum of the forces is system. Each energetic property on the model is represented
expressed as as a single element in the network. The linear graph of a
mechanical system is drawn by identifying locations, called
d2x “nodes,” in the system with distinct velocities, and then add-
∑ Fx = F (t ) − Fshear = M dt 2
(3.9)
ing lines, called “branches,” between the nodes to represent
each energetic property. Each element in the circuit-like lin-
Substituting for the shear force yields ear graph represents the power being sourced (provided),
stored, or dissipated in that element, at any instant in time.
dx d2x We will begin with mass. The energetic property of a
F (t ) − b =M 2
dt dt mass is its storage of kinetic energy:

This is not in the proper form for a differential equation. Re-  1


EM = Mv 2 (1.34)
call from Chap. 2 that the standard form is to separate terms 2
122 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

x,v The third energetic element in our system is that which


applies force F( t) to the mass. The modeling decision was
made to depict the input force F( t) acting on a system as a
F(t) Mass force vector. The use of a vector indicates that the mecha-
Lubricating
M fluid film
nism, or component, applying force to the system is either
1 Damping b unknown or unimportant to the analysis.
Recall that mechanical power is
g
P�
= F·v
(1.17)
Fig. 3.7   Schematic of the mass-damper system showing the velocity
nodes, node 1 on the mass, and node g on the ground The velocity of the point of application of the force F( t) is
the velocity of the mass, v1. Consequently, a force source is a
An ideal mass is rigid and has a single velocity. As you recall, power source. A force source provides mechanical power to
the velocity of a mass must be measured relative to a non-ac- the system at a known or specified force. The velocity of the
celerating (inertial) reference. In practice, we make our mea- point of application is not known. The response of the sys-
surements relative to the surface of the earth, which we will tem to the applied force F( t) determines the velocity of the
call, not surprisingly, “ground.” We will likewise refer to the point of application of the force. Although force is a familiar
frame or base of a machine as a “ground” reference, if it is input to statics and dynamics problems, it is more common
stationary relative to the earth. We will designate distinct ve- in practice to know the velocity of a machine element, which
locities, such as the velocity of a mass and the velocity of its presses or pulls on a dynamic system, than to know the force
ground reference, as “nodes.” The linear graph method uses which is applied. A velocity source is a mechanical power
“nodal subscripts” to identify a velocity node. The subscripts source, in which we know or specify the velocity of the point
are usually numbers, except for reference nodes where ab- of application of an unknown force. The force which the ve-
breviations are used, that is, “g” for ground. The schematic locity source must provide is determined by the reaction of
of the energetic model with the velocity nodes of the mass the system to the velocity source.
and ground is provided in Fig. 3.7. All energetic elements in a linear graph require two veloci-
The kinetic energy of the mass is written using the nodal ty nodes, including the force source. It is convenient to draw a
subscript notation, to identify the velocity difference be- force source as a vector, showing only the point of application
tween the mass and its inertial (ground) reference, of the force. In reality, the force is applied by a mechanism
or mechanical element which needs to push or pull against a
1 suitable reaction in order to provide a force. Anyone who has
E M = Mv12g
(3.10)
2 pushed or pulled, while standing on a slippery surface, can
attest to this fact. We will use the convention of an input force
Notice that the subscript on the kinetic energy E is M for shown as a vector with its point of application indicated, but
mass. its tail floating in space, to indicate that it is reacting against
We now consider mechanical energy lost as heat in the ground, since it must react against something.
damping of the viscous fluid film due to shear. The strain
rate of the fluid was calculated as the difference between the 3.2.3.1 D’Alembert’s Force and Dynamic
velocities of two parallel surfaces divided by gap height h. Equilibrium
The velocities of two surfaces were identified generically So far, the only differences between the linear graph method
as v1 and v2 in Fig. 3.3. The surfaces are the bottom of the and a Newtonian formulation have been minor changes in
mass and the top of the ground, to which we have assigned notation and the use of velocity rather than position as the
the velocities, v1 and vg, respectively. The velocity difference dependent variable. The significant difference between the
between nodes 1 and g, written as v1g, allows us to write the two methods is the how force and torque equilibrium state-
force acting through the damping element, the lubricating ments are written. Recall that the Newtonian formulation
fluid film, in terms of damping coefficient b, and the velocity began by equating the sum of the forces acting on a mass
difference across the damping element: with the acceleration of the mass, Eq. 3.9. The sum of the
forces acting on a mass is often referred to as the “net force,”
Fb = bv1g
(3.11)
∑ Fx = FNet = max
(3.12)
Notice that the subscript on the force is the damping param-
eter, or coefficient b, rather than the subscript “shear” used The essence of a Newtonian formulation is that the unbal-
in Newtonian formulation. anced forces accelerate a mass. Another way of stating this
3.2  Introduction to the Linear Graph Method 123

is in the inverse. When the forces acting on a mass are in variables in a translational mechanical system are force and
equilibrium, the mass does not accelerate. The equilibrium velocity. Accordingly, we will express the acceleration in
requirement that the forces acting on a mass sum to zero, FM = Max as the derivative of velocity,

∑F x = FNet = 0 FM = M
(3.16)
dv
dt
is a “static” equilibrium.
A mass is fundamentally different from other energetic
mechanical elements. A single force can be applied to a mass. 3.2.3.2  Through and Across Variables
In contrast, a spring must have equal and opposite forces at Power is the product of two power variables in mechani-
both ends, in order to transmit or provide any force. You can- cal, fluid, and electrical systems. Only in thermal systems
not push on just one end of a spring and develop a force. is power a single variable. A set of analogies between the
There must be a reaction force pushing back against the other power variables in different types of systems will allow us
end of the spring. The “push back” of a mass develops at the to generalize and exploit the similarities between different
point of application of the force since an inherent property types of systems. There are two complementary sets of anal-
of mass is its resistance to acceleration. In 1743, D’Alembert ogies in common use, one used with linear graphs, which
introduced the concept of an “inertial” force, in the context we are developing, and the other used with “Bond Graphs.”
of extending the principle of virtual work from the static case Linear graphs resemble electric circuits, whereas the term
to a system of particles in motion. The D’Alembert force, “bond graph” describes graphs which look like diagrams of
also known as the inertial force, is the resistance of a mass to molecules with bonds between atoms.
acceleration. D’Alembert’s force is equal and opposite to the The linear graph method categorizes power variables into
net force, which accelerates a mass. An inertial force is the two classes, “through” variables and “across” variables. A
pseudo-reaction force of a mass. through variable acts or flows through an element. A through
A “dynamic” equilibrium includes D’Alembert’s force, in variable has the same magnitude at both ends of an element,
the sum of forces acting on the mass. A dynamic equilibri- with the exceptions of mass and mass moment of inertia and,
um is always zero, since, by definition, D’Alembert’s force, as we shall see later, fluid capacitance. An across variable
− max, equals the opposite of the ΣFx, varies or “drops” across an element. There is a single value
of an across variable at a node. With reference to Sect. 1.3,
∑ Fx − max = 0
(3.13) where continuity and compatibility equations were intro-
duced in the context of a piping system, volume flow rate
where D’Alembert’s force is is a through variable. It is the same at both ends of a pipe.
Through variables sum to zero at nodes where elements con-
FD ′Alembertx = − max
(3.14) nect. Pressure is an across variable. It decreases in the direc-
tion of fluid flow. Across variable differences sum to zero
We will invert the sign of Eq. 3.14 and use the opposite around a closed-loop in a linear graph.
of D’Alembert’s force as the force acting to accelerate the A useful method of identifying a power variable as a
mass, FM, through or an across variable is to consider how the quantity
is measured. For example, current must be routed through an
∑ Fx − FM = 0
(3.15) ammeter, or a fraction of the current routed through a parallel
shunt, for the current to be measured. Current is a through vari-
Mathematically, the force acting to accelerate the mass, FM, able. Conversely, voltage cannot be measured with a single
is the net force, ΣFx. Conceptually, FM = max. The distinction probe at a single location. Voltage is a relative measurement.
between the two previous statements is the difference be- We measure the difference in voltage between two nodes at
tween a force, ΣFx, and the effect of that force, max. either end of an element or between a node at the end of an el-
There are two reasons for including the D’Alembert ement and a reference node. In other words, we measure volt-
force (as FM) in the linear graph method. First, including age across an element between the nodes connected at either
D’Alembert’s force allows a dynamic equilibrium state- end or between (across) a node on the element and a reference
ment to be written for every force summation. Second, by voltage. To measure a force, the load cell (force meter) must
naming the force acting to accelerate a mass FM, we have a be in the load path. The force must act through the load cell.
force associated with the mass element. We shall see that two Force is a through variable. Conversely, the velocity of a com-
power variables per energetic element are needed to apply ponent can be measured visually. Observation of the surface is
the equivalent of Kirkchoff’s circuit laws to a mechanical sufficient. However, velocity is measured relative to a refer-
system represented as an energetic network. The power ence velocity, usually, but not necessarily zero. The difference
124 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Table 3.1   Linear graph and bond graph variable analogies reference node. Since the mass is rigid, the two nodes on the
System Type Power Variable Linear Graph Bond Graph mass, identified as node 1, are a single velocity. Likewise,
Variable Type Variable Type
two ground nodes identified as g are shown. They are also a
Fluid Volume flow rate Through Flow
Pressure Across Effort
single velocity.
Electrical Current Through Flow A linear graph is a network. It consists of a set of nodes
Voltage Across Effort that represent distinct values of the across variable, connect-
Translational Force Through Effort ed by lines known as “branches” which represent the ener-
Mechanical Translational velocity Across Flow getic elements. An arrowhead is drawn near the middle of
Rotational Torque Through Effort each branch, to define the positive direction for the through
Mechanical Rotational velocity Across Flow variable in that energetic element. The direction defined as
Thermal Heat flow rate Through Flow positive for an element is arbitrary, except for sources and
Temperature Across Effort a few special cases we will encounter later. Most elements
in a linear graph are identified by their parameters which
represent their energetic property. For example, a damper is
x,v identified with b and a spring with K. This contrasts with an
electrical circuit diagram, in which a resistor is identified by
both a resistor symbol and its parameter. Linear graph de-
F(t) Mass
Lubricating rives its name from the use of lines identified by parameters,
g 1 M
fluid film rather than unique symbols for each type of element. An ex-
1 Damping b ception to the no-special-symbols rule is an input element or
“source.” The symbol for a source is a branch with the input
g variable encircled in the middle of the branch.
Linear graphs are drawn by first drawing and identifying
Fig. 3.8   Schematic of the mass-damper system, showing a reaction (labeling) the nodes by a number or letter. In this example,
against ground for the input force source
we draw and identify the velocity nodes, v1 and vg.
Next, draw the elements between their respective nodes,
between the component’s velocity and the reference velocity which are identified on the schematic of the energetic sys-
is the value needed. Velocity is an across variable. tem. Note the mass elements’ dashed line after the arrow-
We will use the linear graph set of analogies to apply head. This convention indicates there is no reaction force
Kirchhoff’s current and voltage laws to the circuit-like net- from the ground node against the mass. Both the force source
work of our system. The linear graph and bond graph power and the damper must have a reaction force at the ground, in
variable analogies are listed in Table 3.1. order to apply a force at node 1.

3.2.3.3 Drawing a Linear Graph from a Mechanical 3.2.3.4  Power Flow in a Linear Graph
Schematic Sources power the systems, and they are classed as “ac-
The first and the most important step in drawing a linear tive” elements. The positive direction for the through vari-
graph is to find and label nodes of distinct values of the able of a source element is the direction of an increase in
across variable on the schematic of the energetic system. the across variable. In our system, the force source reacts
The across variable in mechanical systems is velocity. We against ground which has zero velocity. The point of applica-
cannot see most of the variables we work with, but we can tion of the force will move and, thus, has non-zero velocity.
see motion. “Distinct” values of velocity are those at either A positive force increases the velocity of node 1, relative to
end of the elements in the schematic of an energetic system, ground velocity, and a negative force decreases the velocity
with the exceptions of a mass, a mass moment of inertia, and of node 1, relative to ground. The arrow on the force source
a fluid capacitance. A mass and a mass moment of inertia points in the source’s positive direction, from the ground
are rigid. They have a single velocity. The second velocity node to node 1. Elements which store or dissipate energy are
node associated with them is the velocity of the inertial refer- classified as “passive” elements. The mass and damper are
ence. Any non-accelerating reference can be used as the iner- passive elements. The positive direction for the through vari-
tial reference velocity. We will refer to an inertial reference able of a passive element is the direction of a decrease of the
which is stationary in the local frame of reference as ground. across variable. Consequently, the arrows on the mass and
Figure 3.8 shows nodes on either side of the vector, which the damper elements point from node 1 to ground.
represents the force source, as well as nodes on either side of Each element in a linear graph represents a power flow. An
the fluid film, which contributes to the damping b. The latter element’s power flow is the product of that element’s through
two nodes also serve as a node on the mass and an inertial variable and the difference between the across variables on
3.2  Introduction to the Linear Graph Method 125

e­ ither end of the element. The positive direction, as defined by a 1 b 1


the arrowhead, specifies the order of the across variable differ-
ence, value at the first node minus the value at the second node.
The power flow of passive elements, such as masses, springs,
and dampers, is into the element, since the energy is being ei- F(t)
ther stored or dissipated. The power flow of a source is usually b M
out of the source and into the system, but ideal, linear power
sources can also “sink” or accept power from the system.
g g
3.2.3.5 Node Equations, Equilibrium,
and Continuity
Fig. 3.9  a Nodes of distinct values of the across variable velocity.
The network, or circuit-like linear graph representation of b Linear graph of a force source driving as mass-damper system
the system, allows two sets of equations to be written, as
developed in Sect. 1.4, using the example of a piping system.
One set of equations, analogous to Kirchhoff’s current law, a 1 b 1
is written by summing the through variables at nodes. The
sum of through variables at a node equals zero. We will use
the “bank account” sign convention, for summing through
variables at nodes. A through variable directed into a node or F(t) b M F(t) b M
control volume is positive. A through variable directed out of
a node or control volume is negative.
The through variable in a translational mechanical sys-
tem is force. Consequently, a node equation is a statement of g g
dynamic equilibrium, which is possible by including in the
summation the force acting to accelerate the mass, FM. Sum- Fig. 3.10  a Control volume cutting branches and through variables at
ming through variables at node 1, the force from the source, node 1. b Control volume at node g
F( t), is directed into node 1 and is positive, Fig. 3.10a. The
force acting through damping b, Fb, and the force acting to element itself. This may seem awkward at first, but it reduces
accelerate the mass, FM, are directed out of node 1 and are the number of equations we need to work with. Again, for
negative, passive elements, the across variable decreases or drops in
the positive direction of the through variable, as defined by
F (t ) − Fb − FM = 0
(3.17) the arrowhead on the branch. The difference in the across
variable is calculated in that order. Referring to Figs. 3.9 and
Notice that we have written one node equation less than we 3.10, the positive direction of the force acting through the
have nodes in the linear graph. We will always write one damper, Fb, is from node 1 to node g. Hence the velocity
node equation less than we have nodes. We will always omit drop across the damping element is as follows:
the node equation for the ground node. The omission of a
node equation for the ground node will prevent the inclusion v1 − vg ≡ v1g
(3.18)
of a dependent equation within the set of energetic equations
to be reduced to a differential system equation. It will also Note the use of the node number or letter as the subscript,
avoid the use of a fictitious reaction from ground for masses to indicate the velocity of a single node. The difference be-
and mass moment of inertias. tween the velocities of the nodes, 1 and g, is indicated by the
use of both subscripts in the order in which the difference
3.2.3.6  Path and Loop Equations, Compatibility was calculated.
The across variable in a translational mechanical system is The use of node subscripts to identify the across variable
velocity. The second set of equations, extracted from the lin- difference is contrary to the notation commonly used for the
ear graph, is written by summing across variable differences, voltage drop in an electric circuit, which is identified by the
either between two nodes along different paths or around a element’s subscript. Using the electrical engineering con-
closed loop. These equations are analogous to Kirchhoff’s vention, the velocity difference across the damping element
voltage law. is expressed as
We will identify the across variable difference between
the ends of an element by subscripts which denote the nodes v1 − vg ≡ vb
(3.19)
the element is connected between. We will not denote the
126 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.11   The mass-damper 1 positive and negative differences for the values to be “com-
loop is from node 1 across the
mass to node g, then back across
patible” and sum to zero. The loop equation for the mass-
the damper to node 1. The mass- damper loop of the system from node 1 to node 1 is
damper paths are from node 1
to node 2, across the mass and v1g − v1g = 0
across the damper
b M (3.20)
Equation 3.20 is written with the across variable differences
in the positive direction, as defined by the arrowheads on
g the branches which point from node 1 to node g. Reversing
the order of the node subscripts inverts the sign of the across
variable differences:
Although the electrical engineering style or element sub-
script notation is simpler, the advantage of the node subscript −v1g = −(v1 − vg ) = −v1 + vg = vg1
(3.21)
notation is clear, when there are elements in parallel, as in the
force source, mass, damper system, Fig. 3.10: This notation is not recommended. Reversing the subscripts
on the across variable “hides” the negative sign, and tends
Velocity Across Vari- Element Sub- Node Subscript to lead to subsequent sign errors, when the subscripts are
Difference able Drop script Notation Notation inadvertently written in the positive order. It is best to use
Across the v1 − vg vsource v1g a negative sign, rather than to reverse the order of the node
force source
from node
subscripts to invert the sign.
1 to node g A path equation equates the across variable differences
Across the vb along two different paths between the same two nodes and
v1 − vg v1g
damper does not need to contain negative values. The path equa-
from node tion from node 1 to node g, across the mass and across the
1 to node g
damper, is
Across the v1 − vg vM v1g
mass from node
1 to node g
v1g = v1g
(3.22)

Equation 3.22 is a “trivial” equation. It is correct but without


The node subscript notation, which we shall use, makes it any useful information.
obvious that all of the elements have the same velocity dif-
ference or drop across them. The element subscript notation 3.2.3.7  Elemental or Constitutive Equations
requires two equations to equate the velocity drops: The continuity and compatibility equations describe how the
elements of a linear graph are connected. The elemental or
vsource = vb and vb = vM “constitutive” equations relate the two power variables of an
element to each other. The elemental equations for our sys-
The maximum number of unique loop or path equations, tem describe the force–velocity relationships in the mass and
which can be formed in a circuit or a linear graph, equals the the damper.
number of “holes” between elements in the ­network. This The familiar equation, F = Ma, when written as an expres-
maximum is not guaranteed, however, because the number sion of the system’s power variable, velocity, becomes the
of elements which form the holes is a second factor. We have general form of the element equation for mass,
two holes in this network. In the general case, we could form
at most two unique loop or path equations. As it happens, dv
F=M
with this particular network, we can only form one unique dt
loop or path equation. Further, that lone equation is a trivial
equation, meaning it contains no useful information. We will From the general form, we write the element equation for the
write it for the sake of completeness, but it cannot be used in system’s specific mass element, by adding the identifying
the derivation of the differential system equation. subscripts to the force, and acting to accelerate the mass, FM,
Loop equations and path equations contain the same in- as the velocity drops across the mass, v1g,
formation, Fig. 3.11. They differ only in signs. A loop equa-
tion equates the sum of the across variable differences be- dv1g
(3.23)
FM = M
tween nodes around a loop to zero. Since a loop equation dt
begins and ends at the same node, the sum must contain both
3.2  Introduction to the Linear Graph Method 127

x 1 ,v1 x 2 ,v 2 lates one of the power variables in the system to the input
b variable, equals the number of “independent” energy stor-
Fb Fb
ages in the system. If there is only one energy storage ele-
Node 1 Node 2 ment (in this case, mass), that energy storage element must
be independent. Consequently, the differential system equa-
Fig. 3.12   A dashpot schematic symbol for translational damping.
Compression, or extension, of the piston–cylinder forces fluid through tions derived below will be of first-order.
a flow restriction, shearing the fluid The energy storage variable, v1g in this system, is called
the system’s “state” variable. If we know the value of sys-
We developed the relationship between the force, shearing tem’s state variable at any time t, then, with the value of
the viscous lubrication film, and the velocity of the mass, the input variable, we can calculate the value of every other
Eq. 3.11, when we derived the Newtonian formulation, power variable in the system at that time t. We will first make
use of state variables to determine the initial value of an out-
Fb = bv1g
(3.11) put variable in Sect. 3.4. In Chap. 7, we will introduce the
“state-space” which describes a dynamic system with multi-
Although Eq. 3.11 was developed for the specific case of ple independent energy storage elements by a set of simulta-
shearing a viscous film between two planar surfaces, it is neous first-order differential equations of the state variables,
the form of the elemental equation for all linear translational rather than by a single higher-order differential equation.
damping. A mechanical schematic symbol to represent trans-
lational damping is a cross section of a piston inside a cyl- 3.2.3.9 Derivation of a Differential System
inder, Fig. 3.12. The physical device is called a “dashpot” Equation
and dissipates mechanical power has heat by pumping a fluid We have extracted all information from the linear graph
through an orifice, a small diameter tube, or a small annular which describes how the energetic elements are connected to
space. The working fluid is typically oil but air is also used another. We have written element equations which describe
in some designs. how the energetic elements behave individually. Finally, we
Although we can write a functional description of the have written energy equations which can be used to estab-
input variable, F( t), we cannot write an elemental equation lish the initial condition of any power variable in the system,
that relates the through variable, F( t), to the across variable, other than the input force which we control. By systemati-
v1g( t), of the source. The velocity across the source will be cally extracting all of the information from the linear graph,
determined by the dynamic response of the system to the we can be assured that we have sufficient information to de-
force, F( t), imposed on the system. It is physically impos- rive a differential system equation to describe the energetic
sible to specify both power variables of the source. behavior of the system.
The verb “derive” connotes an expression of mathematic
3.2.3.8  Energy Storage Equations logic. Once we have extracted and collected all of the math-
The fourth category of energetic equations expresses the ematical statements of the energetic properties of a system,
amount of energy stored in individual energy storage ele- we need effective algebra, more than mathematical logic. We
ments, as well as in the system as a whole. Recall that we use the verb “reduce” to describe the process of elimination
are neglecting the energy dissipated as heat by shearing the by substitution, algebraic rearrangement, and occasional dif-
lubricating fluid. The energy stored in the mass-damping ferentiation to arrive at a differential system equation. We
system is stored in the mass as kinetic energy. The familiar begin the reduction of this set of equations to an input–output
equation for kinetic energy is written in terms of mass pa- relationship, relating the applied force to our output of inter-
rameter M and velocity relative to ground, v1g, est, by choosing any of the node, path, or element equations
except a trivial equation. Do not use the energy equations to
1 derive a differential system equation. The energy equations
E M = Mv12g
(3.24)
2 are used to determine the initial conditions needed to solve
the system equation.
Equating the energy in the system with the energy in the We will choose two output variables: the velocity of the
mass, mass, v1g, and the force acting to accelerate the mass, FM.
E sys = E M (3.25) Again, reduction of the energetic equations to a differential
 system equation is achieved primarily through elimination
These are very important equations, but they are not used to by substitution. Our objective is to eliminate all of the power
derive the system equation. They are used to establish the variables in the equation except for the input and the out-
initial condition needed to solve the equation, as we shall put variable. Once we have an equation written in terms of
see. The order of a differential system equation, which re- the input variable, the output variable, time, and the element
128 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.13  a Translational me-


chanical mass-damper system
a b 1
driven by a force source. x,v
b Linear graph of the mass-
damper system driven by a
force source
Mass F(t) b M
F(t) Lubricating
g M fluid film
1 Damping b
g g

parameters, we have completed the reduction. We will then 


F (t ) M dv1g F (t ) M dv1g
express the system equation in a “time constant” form, and = + v1g → = + v1g (3.26)
then check the units of the system equation in terms of the b b dt
 b b dt
τ
power variables, F and v, and time t.
It is always helpful to organize the information one works
with. The schematic of the energetic model and the linear Check the units of the system equation in terms of the power
graph drawn from it should be grouped with the set of ener- variables and time. The derivative operator and the element
getic equations. The set of energetic equations is used first, and node subscripts do not affect units. They are removed to
to formulate the differential system equation and, again, to simplify the notation:
establish the initial conditions needed to solve the system
equation. The order the energetic equations are presented  dv1g  v Ft
[ F ] = M  → [ F ] =  M t  → [ M ] = 
is not important. It is important to organize the energetic  v 
M
 dt   
equations by type. If an equation is missing, it may not be
immediately apparent during a reduction. Be systematic in F
extracting the information from the linear graph. Separate [ F ] = bv
b 1g
 → [ F ] = [bv ] → [b] =  v 
 
the set of energetic equations from any subsequent algebra.
Combining statements of physical truth with possibly er-  F   M dv1g  F  M v
roneous algebra is counter-productive. Finally, identify the  b  =  b dt  + v1g  →  b  =  b t  + [ v ]
 
input variable and the power variable to be the output vari-
able, that is, the objective of the reduction, at the top of the
 v   F t v v
reduction. F =  + [v] → [v] = [v] + [v]
 F   v F t
Energetic Equations
The units are in check. Unfortunately, this check can only
Continuity, Node Eq: F − Fb − FM = 0 identify equations in which the units are inconsistent be-
tween terms. The system equation can still be erroneous, if
Compatibility, Path Eq: v1g = v1g the units are consistent. However, it is definitely erroneous,
dv1g if the units are inconsistent.
Element Eqs: Fb = bv1g FM = M Reduction ii: Input F (t ), Output FM
dt
1 F (t ) − Fb − FM = 0 → F (t ) = Fb + FM → F (t ) = bv1g + FM
Energy Eqs: E sys = E M E M = Mv12g
2
dF (t ) d dF dF (t ) dv1g dFM
Reduction i: Input F (t ), Output v1g = (bv1g ) + M → =b +
dt dt dt dt dt dt

F (t ) − Fb − FM = 0 → F (t ) = Fb + FM dF (t )  F  dF dF (t ) b dF
= b M  + M → = FM + M
dt M dt dt M dt
dv1g dv1g
F (t ) = bv1g + M → F (t ) = M + bv1g M dF (t ) M dFM
dt dt (3.27) = + FM
b dt b dt
3.2  Introduction to the Linear Graph Method 129

Check the units: x 1 ,v1 x 2 ,v 2


FK K FK
Ft F
[ M ] =  [b] =  v 
 v    Node 1 Node 2

Fig. 3.14   A spring schematic symbol for translational damping


 M dF   M dFM  M F  M F 
 b dt  =  b dt  + [ FM ] →  b t  =  b t  + [ F ]
 
The product of the velocity difference across the spring and
 F t v F  F t v F spring constant K is the rate at which the force acting through
 =  + [F ] a spring changes.
 v F t  v F t
The two example systems shown in Fig. 3.15 differ in
[F ] = [F ] + [F ] the type of power source, and the arrangement of the spring
and damper. The system shown in Fig. 3.15a is driven by a
force source which acts against a “rigid and massless bar.” A
The units are consistent. mechanical element which is ideally rigid can neither store
strain energy nor dissipate energy. If the element is also ide-
ally massless, it cannot store kinetic energy either. Conse-
3.2.4 Examples Illustrating the Linear quently, a rigid, massless bar has no energetic properties.
Graph Method Its purpose in an energetic model is to transmit force and
power. In the example system, the rigid and massless bar
We will illustrate the linear graph method by deriving system connects the force source with the spring and damper. If the
equations for two more first-order translational mechani- physical system had the same configuration as the energetic
cal systems, and a second-order translational mechanical schematic, we would expect the bar to rotate during its tran-
system. The systems contain elastic strain energy storage, sient response. However, an energetic schematic represents
which is represented schematically as a spring. The famil- the energetic properties of a physical system and need not
iar schematic symbol for a translational spring is shown in resemble the physical system. Both the strain energy storage
Fig. 3.14. Note that there are nodes at either end of a trans- and the viscous damping in the schematic may, in fact, be at-
lational spring. The ends can displace independently of one tributes of a single machine component. Defining only trans-
another. The deformation of the spring is the difference in lational displacement and velocity on the energetic sche-
the displacements of the two ends. The rate of deformation is matic declares that we have restricted motion in the model
the difference in the velocities of the two ends. Also note that to transition. If there were rotation in the physical system
the magnitude of the spring force, FK, is the same at both the and a significant amount of energy was stored or dissipated
ends. The spring force acts through the spring. in rotation then we would add a rotational subsystem to the
In an energetic model, a spring represents strain energy model, using an interface element called a transducer, which
storage. The strain energy may be stored in a machine com- is introduced in Chap. 6.
ponent intended to act as spring but it need not be. Any ma- The system shown in Fig. 3.15b is driven by a velocity
chine loaded machine component will deform. It the defor- source. A translational velocity source is often a rod con-
mation stores a significant amount of elastic strain energy, nected to the energetic system it is driving. It could be the
then than energetic property of the physical system is rep- operating rod of a pneumatic or hydraulic piston under feed-
resented schematically by a spring. A spring is identified by back velocity control. Another type of translational velocity
its spring constant K. The machine design term “spring rate” source is a linkage connected to another machine or machine
is not used in system dynamics because “rate” is a misno- element which moves through a prescribed motion. Regard-
mer. The familiar expression relating force acting through a less of the details, the motion of the machine element acting
spring to the deformation of the spring, Eq. 1.7, as a velocity source is either controlled or known. What is
unknown is the force, which must be applied by the machine
FK = K x12
(1.7) element to impose the specified velocity at its point of ap-
plication. In short, we control or know the velocity of the
is differentiated with respect to time, so as to express the re- point of application. The dynamic response of the system
lationship in terms of the power variables, force and velocity determines the force required to impose that velocity on the
of a translational mechanical system: system.
Velocity sources are unfamiliar to most students. They
dF are not used in dynamics because, in the general case, it is
= Kv
(1.42)
dt impossible to apply a velocity source to a mass. A physi-
130 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.15  a Translational


mechanical spring-damper
a x,v b
system driven by a force source.
b Translational mechanical K
spring-damper system driven
by a velocity source x,v b
F(t) v(t) K
b

Rigid and massless bar

cally real source can supply only a finite amount of power. There are two velocity nodes associated with every ener-
We shall see in Sect. 3.4 that instantaneous application of getic element. Except for a mass, the velocity nodes are at
a step input would require an instantaneous change in the either end of every element. Again, masses are special. The
velocity of the mass. A finite change in the velocity of the second node associated with a mass is the inertial reference,
mass means a finite change in the kinetic energy of the mass. ground. Draw both nodes associated with each element on
Power is the flow rate of energy. An instantaneous change in the schematic, Fig. 3.16a, then eliminate redundant nodes
the kinetic energy of a mass requires infinite power: of the same value of the across variable, Fig. 3.16b. Rigid
objects have a single translational velocity, when motion is
∆E ∆E restricted to one axis or dimension. The three nodes on the

(1.28)≈ ∞ → P�
≈∞
dt 0 rigid and massless bar are the same velocity and, therefore,
the same node.
Likewise, in the general case, a force source cannot act di- Three ground nodes have been identified. The force
rectly on a spring. We would need infinite power, if we were source, represented as a vector, must react against some-
to dictate the force acting through a spring by imposing a thing. Since its reaction is not shown, we assume that it re-
step change on it, since the strain energy stored in a spring is acts against ground. Only one of two ground nodes on the
a function of the force acting through the spring. The argu- right end of spring K and damper b is needed. Note that
ment can also be made using deformation. If a spring were to ground is also indicated schematically by the cross-hatching.
deform a finite amount instantaneously, the end of the spring Draw the linear graph by first drawing and labeling nodes
would displace at infinite velocity. You may recall problem 1 and g. Then, add the elements by drawing lines (branch-
statements from previous courses, in which a force was ap- es) between the respective nodes. Draw arrowheads on the
plied directly to a spring. This is impossible in the general branches to define the positive direction of the through vari-
case, but not in two special cases. The first special case ap- able force for the element. Identify the branch with the pa-
plies to a force from the source, acting through the spring rameter of the element, Fig. 3.17.
which is increased gradually over time, limiting the power The energetic equations of the system are categorized as
flow from the force source. We will use for this special case, follows:
when we investigate the application of sinusoidally varying 1. Continuity or Node Equations: Sum through variables
forces to systems. The second special case occurs, when the (forces) at nodes. Write one continuity equation less than
transient response is neglected, and only the steady-state is there are nodes. Omit the ground node.
considered. Steady-state occurs after the force source and the 2. Compatibility or Path Equations: Sum across variable
spring have reached equilibrium. This special case is com- differences between two nodes along different paths.
monly used in machine design but is not used in system dy- Write no more compatibility equations than there are
namics, since we are interested in the transient response of holes in the linear graph.
systems. 3. Element Equations: Write one element equation relating
the two power variables of each passive element. There is
3.2.4.1  Force Source-Spring-Damper System element equation for a source.
We will now draw the linear graph of the force source- 4. Energy Equations: Equate the energy stored in the sys-
spring-damper system, shown in Fig. 3.15a. The first step tem with the energy stored in the energy stored elements.
is to identify the nodes of the across variable. The across Write an energy equation for each energy storage ele-
variable in a translational mechanical system is velocity. ment.
3.2  Introduction to the Linear Graph Method 131

Fig. 3.16  a The nodes at both


ends of each element of the sys-
a x,v b x,v
tem are identified. b Redundant
nodes removed and the nodes K K
labeled

F(t) F(t) g
b 1 b
g

Rigid and massless bar Rigid and massless bar

a 1 b 1 system equation. Start with a summation containing either


the input or output variable, if possible.
Reduction i: Input F( t), Output FK

F(t) F (t ) − Fb − FK = 0 → F (t ) = Fb + FK → F (t ) = bv1g + FK
b K
b dFK
F (t ) = + FK (3.29)
 K dt
g g
Check the units of the system equation in terms of the power
Fig. 3.17  a Velocity nodes of the system shown in Fig. 3.16. b Linear variables, F and v, and time t:
graph of that system
dFK F F 
= Kv1g → [ K ] =   Fb = bv1g → [b ] =  
Writing the set of energetic equations in a neat and orderly dt t v  v
manner saves time and reduces errors. Discipline yourself to
write the complete set of energetic equations, before begin-
 b dFK  b F
ning the reduction of the equations to a differential system  F (t ) =   + [ FK ] → [ F ] =  K + [F ]
equation.  K dt   t 

Energetic Equations  F t v F
[F ] =   + [F ] → [F ] = [F ] + [F ]
 v F t
Continuity or Node Eq: F (t ) − Fb − FK = 0
The units are consistent.
Compatibility or Path Eq: v1g = v1g Reduction ii: Input F( t), Output Fb

dFK F (t ) − Fb − FK = 0 → F (t ) = Fb + FK
Element Eqs: Fb = bv1g = Kv1g
dt
Differentiate both sides with respect to time, in order to sub-
F2 stitute for FK:
Energy Eqs: E sys = E K EK = K
2K
dF (t ) dFb dFK dF (t ) dFb
= + → = + Kv1g
We will reduce the set of energetic equations to the differ- dt dt dt dt dt
ential system equations, which relate the input force to the
force acting through the spring, FK, and the force acting dF (t ) dFb F
through the damper, Fb. Recall that the process utilizes elimi- = +K b
dt dt b
nation by substitution, differentiating if necessary. The en-
ergy equations are not used in the reduction to a d­ ifferential This is the system equation, since it is written in terms of
the input and output variable, time, and the element param-
132 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

eters. Express the system equation in a time constant form, x,v


by clearing the coefficient from the undifferentiated, or zero- b
order, output variable term: v(t) K

b dF (t ) b dFb
(3.30) = + Fb
K dt K dt
Fig. 3.18   Nodes of distinct values of the across variable velocity, iden-
Check the consistency of the units in terms of the power vari- tified on the energetic systematic
ables and time:

dFK F F  x,v


= Kv1g → [ K ] =   Fb = bv1g → [b ] =   b
dt t v  v K
v(t)
g
 b dF (t )   b dFb  b F b F
g 1 2
 =  + [ Fb ] →  = + [F ]
 K dt   K dt   K t   K t 
Fig. 3.19   Node labeled. A ground node is needed for the velocity
 F t v F  F t v F
 =  + [F ] → [F ] = [F ] + [F ] source to react against
 v F t  v F t
since they supply power to the system. The across variable
The units check. increases in the positive direction of the through variable in
a source element.
3.2.4.2 Velocity Source Spring-Damper System Write the complete set of energetic equations before be-
The linear graph of the velocity source spring-damper s­ ystem ginning the reduction to a system equation. Write one con-
shown in Fig. 3.15b is drawn following the same procedure. tinuity or node equation less than there are nodes. There are
The first step is to identify the nodes of the across variable three distinct velocity nodes in this system, since we drew a
translational velocity on the schematic of the energetic sys- redundant ground node for convenience. Therefore, we will
tem. Draw nodes at either end of each element on the sche- write two node equations. We will omit the continuity equa-
matic, Fig. 3.18. The velocity source is a power source. It tion for the ground node. There is force acting through the
will exert the force necessary at its point of application to im- velocity source. It is an unknown, which is established by the
pose the specified velocity. Velocity sources must have a re- response of the system.
action, or they cannot develop the force necessary to impose
the specified velocity. The velocity source is represented as a Energetic Equations
vector, indicating the nature of the source is either unknown
or unimportant. Place a ground node at the tail of the velocity Continuity or Node Eqs: Fsource = FK FK = Fb
source vector. Next, eliminate redundant nodes of the same
value of the across variable velocity, if necessary. In this sys- Compatibility or Path Eqs: v1g ≡ v (t ) = v12 + v2 g
tem, the only redundant nodes are ground nodes which we
are free to keep, Fig. 3.19. dFK
The linear graph is drawn by drawing and labeling the Element Eqs: Fb = bv2 g = Kv12
dt
nodes identified on the schematic, Fig. 3.20a, and then add-
ing branches which represent each energetic element be- FK2
tween the nodes. The velocity source symbol is a circle in Energy Eqs: E sys = E K EK =
2K
the middle of the branch with a “v” in it, to identify the
source’s power variable, which is known or controlled. We will reduce this set of energetic equations to the differ-
Draw arrowheads on the branches to define the positive ential system equations, which relate the input force to the
direction of the through variable force, Fig. 3.20b. In the force acting through the spring, FK, and the velocity drop
spring and the damper, the across variable decreases in the across the spring, v12. Again, start with a summation contain-
positive direction of the through variable. In this system, ing either the input or the output variable, if possible. Make
the spring is compressing, when the velocity of node 1 is successive substitutions to eliminate any variable, other than
greater than node 2. Likewise, the damper is in compres- the input and output variables and time. Element parameters
sion, when the velocity of node 2 is positive and, hence, are not variables, and they are not eliminated.
greater than ground velocity of zero. Sources are special
3.2  Introduction to the Linear Graph Method 133

Fig. 3.20  a The nodes identi-


fied on the energetic schematic, a b K
Figs. 3.18 and 3.19 are drawn 1 2 1 2
and labeled. b Branches repre-
senting the energetic elements
are drawn between the respective
nodes, labeled with the element
parameter, and oriented with b
an arrowhead pointing in the
v(t)
positive direction of the through
variable, force

g g

Reduction i: Input v( t), Output FK Check the consistency of the units in terms of the power vari-
ables and time:
1 dFK Fb 1 dFK FK
v (t ) = v12 + v2 g → v (t ) = + → v (t ) = +
K dt b K dt b dFK F F
dt
= Kv12 → [K ] = t v  Fb = bv2 g → [b] =  v 
   
This is the system equation. Clear the coefficient from the
zero-order (constant) term to express the system equation in  b dv (t )   b dv12   b v  b v
a time constant form:  =  + [ v12 ] →  K t  =  K t  + [ v ]
 K dt   K dt 
b dFK
bv (t ) =
(3.31) + FK  F t v v  F t v v
K dt  =  + [v] → [v ] = [v ] + [v ]
 v F t  v F t
Check the consistency of the units in terms of the power vari-
ables and time: The units check.

dFK F F  3.2.4.3  Force Source-Spring-Mass-Damper System


dt
= Kv12 → [K ] = t v  Fb = bv2 g → [b ] =  
Two equivalent energetic schematics of a spring-mass-damp-
  v
er system driven by a force source are shown in Fig. 3.21.
 b dFK  b F
bv (t ) =  K dt  + [ FK ] → [bv ] =  K + [F ] The schematic in Fig. 3.21a represents the damping b of the
   t  system in the shear of a viscous lubricating film. The sche-
matic shown in Fig. 3.21b shows mass M supported on ideal
F   F t v F frictionless rollers. The system’s damping b is represented
 v=   + [F ] → [F ] = [F ] + [F ]
v   v F t  by a dashpot. Both of these schematics yield the same lin-
ear graph, because they have the same elements connected
The units are consistent. between the same nodes of distinct values of the across vari-
Reduction ii: Input v( t), Output v12 able velocity, as we will demonstrate. The schematics can be
used interchangeably.
Fb F The first step in drawing a linear graph from an ener-
v (t ) = v12 + v2 g → v (t ) = v12 + → v (t ) = v12 + K
b b getic schematic is to find and label nodes of distinct values
of the across variable, translational velocity in this system,
Differentiate with respect to time to eliminate FK, using the Fig. 3.22. Sources, springs, and dampers have velocity nodes
element equation for the spring: at either end of their schematic symbol. In the case of a force
source represented as a vector, place a ground node on the
dv (t ) dv12 1 dFK dv (t ) dv12 1 vector’s tail. Force sources must have a reaction, or they can-
= + → = + Kv12
dt dt b dt dt dt b not apply a force. Masses are special. Since their only prop-
erty is storage of kinetic energy, they must be rigid. A mass
This is the system equation. Express it in a time constant has a single velocity node on the element. A mass’ other ve-
form: locity node is its inertial reference, ground.
Write the complete set of energetic equations, before be-
b dv (t ) b dv12 ginning the reduction to a system equation. Write one conti-
(3.32) = + v12
K dt K dt nuity or node equation less than there are nodes. Therefore,
134 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.21  a Schematic of a


force source spring-mass-damper
a b x,v
system, with the damping b
represented as a lubricating film x,v
supporting the mass. b Schematic
of a force source spring-mass- F(t)
damper system, with the damping F(t) K M
represented by a dashpot M
K

Lubricating fluid Frictionless


Damping b rollers

Fig. 3.22  a, b Energetic sche-


matics of a force source spring-
a b x,v
mass-damper system with nodes b
x,v
at either end of the force source, g
spring, and damper F(t) 1 1
F(t) K M
1 1 g g 1 g
M
g
1 K
g
Lubricating fluid Frictionless
Damping b rollers

we will write the node equation for node 1, and omit the substitute for either the input or output variable. Substitute to
ground node. There are three holes in the linear graph. If eliminate all other power variables:
the linear graph had two elements in each path from node 1
to ground, then three useful compatibility or path equations dv1g
F (t ) = Fb + FK + FM → F (t ) = bv1g + FK + M
could be written. Since there is only one element in each dt
path from node 1 to ground, only one unique equation can be
written, and it is a trivial equation. b dFK d  1 dFK 
F (t ) = + FK + M  
K dt dt  K dt 
Energetic Equations
b dFK M d 2 FK
F (t ) = + FK +
Continuity or Node Eq: F (t ) = Fb + FK + FM K dt K dt 2

Compatibility or Path Eq: v1g = v1g b dFK M d 2 FK


F (t ) = + FK +
K dt K dt 2
dFK dv1g
Element Eqs: Fb = bv1g = Kv1g FM = M
dt dt This is the system equation. Express it in standard form by
rearranging the derivatives into decreasing order, and clear-
FK2 1 ing the coefficient from the highest-order derivative:
Energy Eqs: E sys = E K + E M EK = EM = Mv12g
2K 2
M d 2 FK b dFK
F (t ) = + + FK
We will reduce the energetic equations to system equations K dt 2 K dt
for the two energy storage variables, FK and v1g. These are
the system’s state variables, because they, along with the K K M d 2 FK K b dFK K
F (t ) = + + FK
input variable, determine the energetic state of the system. M M K dt 2 M K dt M
Reduction i: Input F( t), Output FK
K d 2 FK b dFK K
We start with continuity equation, because it is a sum- F (t ) =
(3.33) 2
+ + FK
mation that includes the input and output variables. Do not M dt M dt M
3.2  Introduction to the Linear Graph Method 135

Check that the units of all terms are consistent. Substitute  1 dF (t )   d v1g   b dv1g   K
2

equivalent units for the element parameters:    2 +
=  +  v1g 
 M dt   dt   M dt   M 
F dFK F
Fb = bv1g → [b] =  v  = Kv1g → [K ] = t v   1 F   v   b v  K 
v
  dt    M t  =  t 2  +  M t  +  M 
dv1g Ft
FM = M → [ M ] =   v F  v F v v  F v 
dt  v    =  2+ + v
 F t t  t   v F t t  t v F t 

  d FK   b dFK   K
2
K  v v v v
 M F (t ) =  dt 2  +  M dt  +  M FK   t 2  =  t 2  +  t 2  +  t 2 
   

 K  F   b F   K 
 M F  =  t 2  +  M t  +  M F  The units check.

F v  F   F v F   F v  3.2.5 Summary of the Introduction to Linear


 F =  2  +  + F
 t v F t t
    v F t t   t v F t  Graphs
F  F  F  F  The most important step in developing a differential system
 t 2  =  t 2  +  t 2  +  t 2 
equation is modeling the energetic system. The modeling
process requires judgment, in order to retain the significant
The units check. energetic attributes of a system, while neglecting those which
Reduction ii: Input F( t), Output v1g can be considered insignificant. Understanding dynamic sys-
We again start with a continuity equation, because it is a tems is a prerequisite for modeling dynamics systems. To-
summation which includes the input variable: wards that end, energetic models are provided in this text as
schematics with the energetic attributes indicated.
dv1g Begin drawing a linear graph from the schematic of an
F (t ) = Fb + FK + FM → F (t ) = bv1g + FK + M
dt energetic model by identifying the nodes of distinct values of
the across variable. Mark and label them on the schematic to
dF (t ) dv1g dFK d  dv1g  prevent the omission of any nodes. An error in this step car-
=b + + M
dt dt dt dt  dt  ries forward through the subsequent effort. In translational
mechanical systems, the across variable is translational ve-
dF (t ) dv1g d 2 v1g locity. There must be an across variable difference associated
=b + Kv1g + M
dt dt dt 2 with every element.
The depiction of the source, acting on a system as a vec-
This is the system equation. Express the system equation in tor, is a modeling convention, which indicates that details
standard form by rearranging the derivative terms, in order of how the power is generated and applied to the system
of decreasing order of differentiation, and clearing the coef- are either unimportant or unknown. The input force’s point
ficient of the highest-order output variable term: of action has a distinct velocity. The tail of the force vector
must also have a distinct velocity, which is different from
dF (t ) d 2 v1g dv1g the head. Mechanical sources must react against some-
=M +b + Kv1g
dt dt 2
dt thing, in order to apply a force to a system. If no element is
shown for the reaction, then assume the source is reacting
2
1 dF (t ) d v1g b dv1g K against ground. A mass is rigid and, thus, can have only
(3.34) = + + v1g
M dt dt 2 M dt M one velocity. The velocity difference across the fluid film is
the velocity difference between the mass and ground. The
Check the units. velocity difference in a dashpot is between the piston and
the cylinder.
F dFK F
Fb = bv1g → [b] =  v  = Kv1g → [K ] = t v  Draw the nodes and label them. Then draw the branch-
  dt   es, which represent the energetic elements between the
nodes identified on the schematic. Energy storage and dis-
dv1g Ft
FM = M → [ M ] =  sipation elements are represented by a line with an arrow-
dt  v  head in the middle, to define the positive direction of the
136 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.23  a, b Energetic sche-


matics of a force source spring-
a b x,v
mass-damper system redundant b
nodes on the ideally rigid mass x,v
g
removed F(t) 1
F(t) 1 K M
M g g g
g
K
g
Lubricating fluid Frictionless
Damping b rollers

a 1 b 1 We will continue to develop the linear graph method in


Chaps. 4, 5, and 6, as we introduce additional energetic ele-
ments.
F(t) b K M
3.3  The Heaviside Unit Step Function

The Heaviside unit step function, us (t − ta ), has one of two


g g values, Fig. 3.25:
Fig. 3.24  a Nodes of distinct values of the across variable velocity of us (t − ta ) = 0 for t < ta
the force source spring-mass-damper schematics of Fig. 3.23. b Linear

(3.35)
graph of those schematics us (t − ta ) = 1 for t > ta

through variable, and identified by the element’s param- We choose what time to define as t = 0. Consequently, we
eter. Force is the through variable in a translational me- define time, t = 0, to be the beginning of the time period of
chanical system. Sources supply power to a system. Their interest to us. We may define t = 0 to be the time at which
symbol is a branch with a circle. The variable within the a step function “transitions” (or turns on). In order to turn
circle identifies the type of source. Only one of the two on a step input at an arbitrary time ta other than t = 0, we
power variables can be controlled by a source. The other subtract the constant ta from the independent variable t. If
power variable is determined by the power drawn from ta is positive, the Heaviside step function’s argument’s tran-
the source by the system. A unique aspect of a source is sition from negative to positive is delayed, or shifted right
that across variable increases in the positive direction of on the time axis, Fig. 3.26. Subtracting a positive constant
the through variable. Conversely, in energy storage and ta from time variable t keeps the argument of the unit step
dissipation elements, the across variable decreases in the function, (t − ta ), negative until time, t = ta . Conversely, if ta
positive direction of the through variable. is negative, subtracting the negative constant from the time
A linear graph is an energetic network. Continuity, or variable, or, equivalently, adding a positive constant, shifts
node, equations and compatibility, or path, equations de- the transition of the step function to that negative time ta.
scribe the configuration of the network. Node equations sum Graphically, adding a positive constant to the time variable
through variables at nodes. One less node equation is written causes the Heaviside step functions argument to become
than there are nodes. We will omit the node equation for the positive earlier, shifting the transition of the step function
ground node. Path equations that equate the sum of across left on the time axis.
variable drops between two nodes along different paths. The The Heaviside step function is an idealization used to
element, or constitutive, equations relate the two power vari- model step inputs. An event which is truly described by a
ables of an element to each other. The element equations are Heaviside step function cannot actually happen in physical
written for the assumed positive direction defined by the ar- systems, except on the scale of subatomic particle, because
rowhead on the element’s branch. Equations for the amount the transition of the Heaviside step function’s value from
of energy in an element and for the amount of energy in the zero to one is instantaneous, Fig. 3.27. On the macroscopic
system are used to establish the initial conditions needed to scale of everyday life, physical changes can be extremely
solve the differential system equation, not to derive the sys- rapid, but they cannot be instantaneous. The power variables
tem equation. in our models can change instantaneously, except for the
3.3  The Heaviside Unit Step Function 137

us(t) us(t)
u s(t) = 1
1 1

u s(t) = 0
0 time time
t = 0- t = 0+
Fig. 3.25   The Heaviside unit step function. The unit step function t=0
transitions from 0 to 1, when its argument transitions from negative
to positive
Fig. 3.27   Ideal Heaviside unit step function
us(t-t a ) us(t-t a ) = 1
1
u(t) u(t) = 1
1
us(t-t a ) = 0
0 t-t a < 0 ta t-t a > 0 time
u(t) = 0
t<0 0 ∆t ∆t < t time
Fig. 3.26   Time shifted Heaviside unit step function. Subtracting ta
from t shifts the time at which the function transitions from time, t = 0
to time, t = ta Fig. 3.28   Realistic unit step transition

of the transient response, which occurs during transition of


state or energy storage variables, which require time. The the input is small. When almost all of the system’s transient
value of a state (energy storage) variable cannot change in- response occurs after the actual transition of the input vari-
stantaneously without infinite power, since there would be a able, the Heaviside step function is an excellent model for
finite change in the energy in the storage element over zero the application of the input. On the other hand, when the
time. Energetic models are simplified representations of real duration ∆t of the transition is a substantial fraction of the
physical systems. An actual physical machine component system’s timescale, either the time constant, τ, or the period
used to apply an input to a system possesses all of the en- of oscillation, T, then the system will have begun to respond
ergetic attributes, energy dissipation and both modes of en- to the input, before the transition is complete, and modeling
ergy, to some extent. We have neglected all but the most sig- the input as a Heaviside step function would be inaccurate.
nificant attributes to create a simplified model. In practice, The input model would need a transitional period for the sys-
we cannot generate ideal step inputs because of the energy tem’s response to be accurately predicted, Fig. 3.28.
storage attributes of our actual power sources. The closer we
examine power sources, the clearer it is that they are, in fact,
dynamics systems. 3.3.1 Differentiation of the Heaviside
Fortunately, if the input variable changes from zero to Unit Step Function
a constant value over a time which is small relative to the
timescale of the transient response of the system it is driv- Many of our differential system equations have differen-
ing, then the input “looks” like a step change to the system. tiation of the input function. Consequently, we must ad-
There are many circumstances, when an input variable can dress the question of how to handle differentiation of the
be reasonably modeled as an instantaneous step change. A Heaviside step function. The derivative of the Heaviside
Heaviside step function can be used to model an input vari- unit step function with respect to time is easy to evaluate
able’s transition, when the duration of the transition of an for t ≠ ta , since the step function’s value is a constant, ei-
actual physical input power variable ∆t is brief, relative ther zero or one:
to the timescale of the system’s response measured, by ei-
ther its time constant, τ, or period of oscillation, T. When  dus (t − ta )
= 0 for t ≠ ta (3.36)
duration ∆t of the transition of a physical input variable is dt
short relative to the system’s timescale, then the proportion
138 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

∞ Heaviside unit step. The differential system equation for a


power variable within the system has the derivative of the
δ(t) Heaviside step as an input term, dus (t )/ dt. The initial condi-
tions of the system are determined by the energy in the sys-
tem prior to the application of the Heaviside unit step func-
tion’s transition and the value of the input the instant after it
δ(t) = 0 δ(t) = 0
has transitioned, as we will see in Sect. 3.4. Recall that the
t<0 t>0 time input terms were only used in the method of undetermined
t = 0- t=0+ coefficients to establish the particular solution, which is the
t=0 steady-state solution. The roll of the derivative of the any
input term is to establish its contribution to steady-state re-
sponse (the particular solution). As a result, we ignore its de-
Fig. 3.29   Unit impulse or Dirac delta function, δ( t)
rivative at time t = 0. At first, this statement may be surpris-
ing. How could an infinite value of an input term not have an
The instantaneous transition of the Heaviside step func- immediate effect on the response of the output variable? An
tion is a discontinuity at which its derivative with respect to input term to a system equation which is the derivative of the
time becomes infinite. Mathematically, the derivative of the Heaviside step function does not hammer a system with an
unit step function is the unit impulse function, also known impulse of infinite magnitude and zero duration. The input to
as the Dirac delta function, δ( t), which has infinite ampli- the system is a step of finite magnitude.
tude and zero width (duration), Fig. 3.29. Paul Dirac was a We shall see in Chap. 8 that the value of the derivative of
British physicist influential in the development of quantum Heaviside’s step function at the instant of transition, when its
­mechanics. argument equals zero at t = ta , is problematic in numerical
solutions. Heaviside’s original definition of the step func-
 δ (t ) = 0 for t ≠ 0 tion, Eq. 3.35, has no value at the instant of transition. In
(3.37)
δ (t ) = ∞ for t = 1 other words, the function and its derivative were undefined at
time t = ta . Heaviside’s original definition of the step func-
The unit impulse function is a truly bizarre concept invented tion contains two different types of discontinuities. Heavi-
for mathematical convenience. The term “unit” refers to the side’s step function has a “jump” discontinuity, which means
fact that the value of the integral of the unit impulse equals one: there is one value as t = ta is approached from positive time,
∞ and a different value as t = ta is approached from negative
∫ δ (t ) dt = 1 (3.38) time. Heaviside’s definition also has a “point” discontinuity,
 −∞
meaning the function is missing a value at a point. Specifi-
By itself, this makes sense, since the unit impulse is the de- cally, the function has no value at time t = ta . A number of
rivative of the unit step and integration is the inverse opera- modern definitions of Heaviside’s step function which have
tion of differentiation. added a value at time t = ta or values in the vicinity of t = ta
Infinity as the value of the derivative of a function which to eliminate the point and jump discontinuities. If you have
transitions instantaneously also makes sense mathematically. worked with Mathcad, you may have noticed that Mathcad’s
It is a “singularity,” where we are dividing finite change by Heaviside step function adds two points in the transition in-
zero time: terval. Eliminating the point discontinuity of the step func-
tion is important for machine computation, since numerical
dus (t ) computations need functions to return values.
(3.39) ≡ δ ( 0) = ∞
dt t = 0 Although we will not do so, a finite pulse can be expressed
as a continuum of unit impulses in the time-domain. We will
Although we can calculate a value of infinity, we cannot have use for the unit impulse as an input in the Laplace-­
physically create the time derivative of a power variable with domain, as we shall see in Chap. 10. It serves to initiate the
infinite magnitude. The bizarre aspect of a unit impulse is homogeneous response of a transfer function.
that its unit area, expressed by the integral of Eq. 3.38, is the
product of infinity times zero.
The unit impulse is created by differentiating the Heavi- 3.4  Initial Conditions
side unit step function. What is physical effect on a system
of an input term which differentiates the Heaviside step In a textbook problem, the initial conditions needed to solve
function? Fortunately, the physical effect is straightforward. a differential equation are given. In system dynamics, the
There is none. The input to the actual physical system is a energetic state of the system at the instant before the input
3.4  Initial Conditions 139

Fig. 3.13  a Translational me-


chanical mass-damper system
a b 1
driven by a force source. b Linear x,v
graph of the mass-damper system
driven by a force source
Mass F(t) b M
F(t) Lubricating
g M fluid film
1 Damping b
g g

F(t) Let us say that the system is de-energized before the step
N F0 us(t)
input, F (t ) = F0 us (t ), Fig. 3.30, is applied.
F0 The initial conditions needed to solve a system equation
are the values of the output variable, and its time derivatives
at time, t = 0+ , at the instant after the input is applied. The
number of initial conditions needed to solve a differential
system equation is equal to the order of the system equation.
0 t, sec The order of a differential system equation equals the num-
ber of independent energy storage elements in the system.
Fig. 3.30   Step input driving the mass-damper system of Fig. 3.13 and
the spring-mass-damper system of Fig. 3.31 The mass-damper system has a single energy storage ele-
ment, the mass. It is a first-order system. The initial condi-
tion needed is the value of the output variable at the instant
is applied, and the values of input are used to determine the after the input is applied at time t = 0+ .
initial conditions needed to solve the system equation. The system is described as de-energized before the input
is applied. The mass is the only energy storage element in the
system, and it stores kinetic energy. If there is no energy in
3.4.1  Initial Condition, First-Order System the system, then the velocity of the mass is zero. Expressing
these statements mathematically,
We shall use the mass-damper system of Fig. 3.13 to illustrate
how to determine the initial conditions needed to solve the
E sys ( 0− ) = E M ( 0− )
system equations derived for the velocity of the mass, v1g, and
the force acting to accelerate the mass, FM, Eqs. 3.26 and 3.27. and
Figure 3.13, the energetic equations, and the system equations
1
are reproduced below from Sect. 3.2.3.9 for reference. E M (0− ) = ( )
Mv12g 0 − ( )
→ v1g 0 − = 0
2
Energetic Equations
The energy storage variable, v1g, is the state variable of the
Continuity or Node Eq: F − Fb − FM = 0 system. State variables have two important attributes. First, if
we know the value of the state variable at any time t, then we
Compatibility or Path Eq: v1g = v1g can calculate any other power variable in the system at that
time by using its value and the value of the input value, which
dv1g is always known. Second, state variables are the only variables
Element Eqs: Fb = bv1g FM = M
dt in a system which cannot change instantaneously. State vari-
ables represent the amount of energy stored in energy storage
1 elements. Energy flows at a finite rate, which is power. A finite
Energy Eqs: E sys = E M EM = Mv12g
2 change in the amount of energy stored in an element requires a
finite amount of time for the energy flow to occur. An instan-
System Equations: Eqs. 3.26 and 3.27 taneous change in the amount of energy in an element would
require infinite power, which is physically impossible:
F (t ) M dv1g M dF (t ) M dFM
= + v1g = + FM
b b dt b dt b dt ∆E ∆E

(1.40)≈∞ → P ≈∞
dt 0
140 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Since a state variable cannot change instantaneously, if we Evaluating this equation at time t = 0+ and rearranging the
know the value of the state variable the instant before a step expression yields
input is applied to a system at time t = 0 −, then we also know
its value the instant after the step input is applied at time ( ) ( )
F 0+ − Fb 0+ = FM 0+ ( )
t = 0+. In this example, the system is de-energized at the in-
stant before the step input of force, F( t), is applied. The force We know F (0+ ) = F0 , but Fb (0+ ) is unknown. We will
source can only provide finite power. The energy transferred eliminate it by substitution. Reviewing the energetic equa-
to the system is zero over the infinitesimal time between tions, we see our only choice is the use of the elemental
t = 0 − and t = 0+ when the step transition is on. The mass equation for the damping Fb = bv1g, expressed in terms of
does not change energy. Consequently, the state variable, the the known value of the state variable,
velocity of the mass, v1g, does not change between t = 0 −
and t = 0+ either: ( ) ( )
F 0+ − bv1g 0+ = FM 0+ ( )
E sys ( 0− ) = 0 = E sys ( 0+ ) → E M ( 0− ) = 0 = E M ( 0+ ) Substituting for F (0+ ) and v1g (0+ ) ,

1 1 
2
( )
Mv12g 0 − = 0 = Mv12g 0+
2
( ) ( )
→ v1g 0 − = 0 = v1g 0+ ( ) ( )
F0 − b v1g 0+
0
( )
= FM 0+ → ( )
FM 0+ = F0 (3.43)

As it happens, the state variable, v1g, is an output variable of we have the initial value of the force acting to accelerate
interest, for which we have established an initial value: the mass. Our result makes sense. If the mass is motionless,
then the fluid film is not in shear, and there is no damping
(3.41)
v1g 0+ = 0 ( ) force acting on the mass. All of the applied force acts to
accelerate the mass at the instant after the step input is ap-
Regardless of whether or not the state variable is the output plied. The division of the applied force, F( t), between the
variable, the first step in calculating the initial condition of force acting to accelerate mass FM and damping force Fb is
any output variable is always to determine the value of the dynamic. The division of F( t) between Fb and FM changes
state variable at time t = 0 −, the instant before the input is ap- with time.
plied. Again, a state variable cannot change instantaneously.
The value of a state variable at t = 0 − is its value at t = 0+.
Having established the initial value of the energy storage 3.4.2  Initial Conditions: Second-Order System
(state) variable, v1g (0+ ) = 0, and knowing the value of the
input at time t = 0+ , We shall use the force source-spring-mass-damper system in
Fig.  3.31 as a second-order system to determine the initial
F 0+ = F0 us 0+ = F0
(3.42) ( ) ( ) conditions needed to solve the system equations for the force
acting through spring FK, Eq. 3.33, and the velocity of mass
We can now establish the initial value of every other power v1g, Eq. 3.34. The energetic schematic in Fig. 3.23a, its linear
variable in the system by using algebraic energetic equa- graph in Fig. 3.24b, and energetic equations are reproduced
tions. We need to establish the value FM (0+ ). below for convenience. The system is driven by the step
We cannot use the differential elemental equation for the input F (t ) = F0 us (t ) , shown in Fig. 3.30.
mass, because we do not know the value of the derivative
of the velocity of the mass at time t = 0+ . We can calculate Energetic Equations
the derivative, but we do not know it. Avoid the sophomoric
mistake of confusing knowledge of the value of a variable at Continuity or Node Eq: F (t ) = Fb + FK + FM
an instant in time with the variable being constant and con-
cluding that the derivative of the variable is zero. We know Compatibility or Path Eq: v1g = v1g
the value of the velocity of the mass at time t = 0+ is zero but
will not stay zero. dFK dv1g
Element Eqs: Fb = bv1g = Kv1g FM = M
Perusing the energetic equations for FM, we see that the dt dt
only other appearance of FM is in the continuity or node
equation: FK2 1
Energy Eqs: E sys = E K + E M EK = EM = Mv12g
2K 2
F − Fb − FM = 0
3.4  Initial Conditions 141

Fig. 3.31  a Force source


spring-mass-damper system of a b 1
Fig. 3.23. b Linear graph of the
force source spring-mass-damper x,v
system

1 F(t) b K M
F(t) K
M g
g

g g
Lubricating fluid
Damping b

What are the initial conditions needed to solve the system Likewise,
equation, Eq. 3.33?
v1g (0 − ) = 0 and v1g (0 − ) = v1g (0+ ) → v1g (0+ ) = 0
2
 K d FK b dFK K
F (t ) = + + FK (3.33)
M dt 2
M dt M The only other variable in a system we know at time
t = 0+ is the value of the input, since we control it,
We need the value of the output variable, FK (0+ ) , and its F (0+ ) = F0 us (0+ ) = F0 .
derivative, dFK (0+ )/dt , the instant after step input is ap- The initial conditions are determined by using these
plied to the system at time t = 0+ . In order to derive the known values, the energetic equations, and algebra. The out-
initial conditions, we must know the amount of energy put variable of system equation, FK, is a state (energy stor-
in the system, and the location of that energy the instant age) variable. We assumed the system to be de-energized at
before step input is applied to the system at time t = 0 − . time t = 0 − and have established the values of the state vari-
We will assume that the system is de-energized before the ables to be zero then and at the next instant, time t = 0+ . We
step input is applied, E sys (0 − ) = 0. Since E sys = E K + E M , have established one of the two initial conditions we need,
and strain energy and kinetic energy must be positive due namely,
to the square of their power variables; we know that if
E sys (0− ) = 0 , then E K (0− ) = 0 and E M (0− ) = 0 . Thus, we ( )
FK 0+ = 0
have established the values of the energy storage, or state
variables, at time t = 0 − : We now peruse the energetic equations and see that dFK /dt is
a term of the spring’s element equation:

E K (0− ) =
( )=0
FK2 0 −
2K
( )
→ FK 0 − = 0 dFK
= Kv1g
dt
and
The velocity of mass v1g is a state (energy storage) variable.
1
E M (0 −
) ( ) ( )
= Mv12g 0 − = 0 → v1g 0 − = 0 We have established at time t = 0+ , v1g (0+ ) = 0. Substituting,
2
dFK (0+ )
= Kv1g (0+ )
It is impossible for the amount of energy stored to change dt
instantaneously during the transition of the Heaviside step
input, since any finite change over infinitesimal time would We will now find the initial conditions needed to solve the
require infinite power. The energy storage, or state, variables system equation, Eq. 3.34, again assuming that the system is
are the only variables in a dynamic system which cannot de-energized prior to the application of the step input.
change instantaneously. If we know their values prior to the
2
application of a step input to a system, we know their values  1 dF (t ) d v1g b dv1g K
= + + v1g (3.34)
immediately after the application of the step, because they M dt dt 2 M dt M
cannot change instantaneously:
We have already established the values of the state (energy
( ) ( )
FK 0 − = 0 and FK 0 − = FK 0+ ( ) ( )
→ FK 0+ = 0 storage) variables, FK (0+ ) = 0 and v1g (0+ ) = 0 . The system
142 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

equation’s output variable is v1g, so we have one of the two the input and output variables (the system equation) is a first-
initial conditions we need: order differential equation. If the equations which describe
the elements of the system are linear equations, then the sys-
( )
v1g 0+ = 0 tem equation will be a linear differential equation, and can be
solved by determining the unknown coefficients of a general
The second initial condition needed is the value of solution. The homogenous solution is always an exponential
dv1g (0+ )/dt . Checking the energetic equations, we find it in function of the form
the element equation for the mass: 
X h (t ) = Ae st (3.44)
dv1g ( )=
dv1g 0+ 1
FM = M
dt

dt M
FM 0+ ( ) If the system is of first order, then the characteristic equation
will be a first-order polynomial
We must now establish FM (0+ ) in terms of the known
values of input, F (0+ ) = F0 , and the state variables, s + a = 0 → s = −a
FK (0+ ) = 0 and v1g (0+ ) = 0 . Begin by substituting for FM:
yielding the homogeneous solution
dv1g 0+ ( )= 1
dt M
FM 0+ ( ) (3.45)
X h (t ) = Ae − at

dv1g 0+( )=
dt
1
M
( ( )
F 0+ − Fb 0+ − FK 0+ ( ) ( )) where the exponent is known and is a function of the param-
eters of the dynamic system. Coefficient A is unknown and
dependent on both the initial conditions and the input to the
Eliminate Fb. Checking the energetic equations, we find that system.
the damper’s element equation is written in terms of the state It is convenient to express the homogeneous response in
variable, v1g, Fb = bv1g : terms of the inherent or characteristic timescale of the first-
order system, known as the system’s time constant, τ The
time constant, τ, of a first-order system equals the absolute
dv1g 0+( )=
dt
1
M
( ( )
F 0+ − bv1g 0+ − FK 0+ ( ) ( )) value of the inverse of its eigenvalue, λ, which, again, is its
characteristic value, s

( )=
dv1g 0+ 1 b 1
−t
1
X h (t ) = e st = e τ where τ = =
1
dt M
F 0+ − ( )
M
v1g 0+ ( ) 0

M
( )
FK 0+
0
(3.46)
s λ

( )=
dv1g 0+ 1
F0
The particular solution has the same functional form as the
dt M input driving or forcing the system. When the input is the
product of a constant and the Heaviside step function, which
Our result tells us that at the instant after the step input is is a constant after its transitions from zero to one,
applied, time t = 0+ , all of the applied force is used to ac-
celerate the mass. This makes sense, because no force can be Input (t ) = F0 us (t )
transferred to the spring or the damper, until the mass moves.
then the particular solution is a constant,

3.5  First-Order Step Responses X p (t ) = C


(3.47)

The response of a system to a step input reveals the underly- Physically, the particular solution is the response of the sys-
ing nature of the dynamic system. Dynamic systems with tem after the effects of the initial conditions and the homoge-
one independent energy storage element can only have neous response have decayed to zero. In the ideal case, time
two basic responses to a step input. The system can either must approach infinity for the effects of the initial conditions
monotonically gain energy from the source or lose energy to decay to zero. In practice, the “steady-state” response of a
to the source. There can be no oscillations. Oscillatory step system is defined as when there is no more discernible effect
responses are caused by energy flow between energy storage of the initial conditions and all that remains is the response to
elements within the system. the input. In the particular case of a response of a system to a
Systems with only one independent energy storage ele- step input, the system is declared to be in steady-state when
ment are called “first-order,” because the equation relating there is no more discernible change with time of any of the
3.5  First-Order Step Responses 143

Fig. 3.32   The four possible unit


step responses of a first-order
1 1
system
-t
__ Decay -t
__
eτ 1- eτ Growth

0 0
0 τ 2τ 3 τ 4 τ 5 τ 0 τ 2 τ 3τ 4τ 5τ
time, Time Constants time, Time Constants

1+C 1+C
-t
__ Offset Decay -t
__
eτ + C (1- eτ ) +C Offset Growth
C C
0 0
0 τ 2 τ 3τ 4τ 5τ 0 τ 2 τ 3τ 4τ 5τ
time, Time Constants time, Time Constants

power variables in the system. The steady-state response of the system have the same time constant, τ. There is only one
all power variables in any system to a step input is to reach time constant in a first-order system. However, responses of
a constant value. In practice, steady-state is defined to be individual power variables in a system can be in opposite di-
reached in time equal to a finite multiple of the time constant rections. That is, a variable may grow, while another variable
of the system, t = 4τ, 5τ, or 6τ. The multiples of τ used de- may decay, as the system reaches its new equilibrium.
pends on the linearity of the system, the quality of the signal, All first-order step responses are based on the same ex-
t
and the precision of the measurement. − 

There is no further change with time in a system when ponential function, e  τ  , where e is Euler’s number,
e = 2.7183 ≈ 2.72, the base of the natural logarithm, and τ is
it reaches steady-state in response to a step input. This is
the time constant. The time constant must have units of time
both an observed physical truth and a mathematical fact.
so that the exponent is dimensionless. The time constant, τ,
The particular solution, which is the steady-state response
provides the timescale of the response. Since time in the ex-
of the system, has the same functional form as the input.
ponent is divided by time constant τ, slow systems have large
The Heaviside step input remains constant after its transi-
time constants. Small time constants lead to fast responses.
tion. Hence, the particular solution or steady-state response
A “fast” system is not necessarily a system which reaches a
of a linear system to a step input is a constant. Any power
high velocity. A fast system reaches steady-state quickly, and
variable in the system may be chosen as the output vari-
a slow system reaches steady-state slowly.
able. All power variables in the system will be constant in
There are only four solutions for the step response of a
steady state.
first-order system, Fig. 3.32: two growth equations and two
The general solution to the differential equation describ-
decay equations. The output variable can grow to a constant
ing the dynamic response of a first-order system is construct-
value from either zero or a non-zero initial value C. The out-
ed by adding the homogeneous and particular solutions,
put variable can either decay from an initial value to zero
or to a non-zero constant value C. The constant C is a verti-
X (t ) = X h (t ) + X p (t ) = Ae − at + C
(3.48)
cal shift or offset added to the exponential decay or growth
­function.
There is no need to use this formal solution method to deter-
mine the response of a first-order system to a step input. Re- One formulation of the step response of a first-order
stricting the system to the first order and the input to a Heavi- ­system is
side step function limits the possible responses of a system’s
 −t
output variable. The step response of all power variables in
a first-order system is either a stable exponential growth or ( ( ) )
X (t ) = X 0+ − X ( ss ) e τ + X ( ss ) (3.49)
decay. In a given first-order system, some variables grow
while others decay. Importantly, all changes occur at the same where X (0+ ) is the value of the output variable at the instant
rate and at the same time. All power variables in the system after the step input is applied and X ( ss ) is the steady-state
change simultaneously. The step responses of all variables in value.
144 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

There are two disadvantages of using the formula,


Eq. 3.49. The first is that this formula is, well, formulaic. V(0+)
It is not a functional relationship in the sense that it does
63.2%
not describe the response of a system. It is impossible to V(t) ∆V
visualize the shape of this function until X (0+ ) and X ( ss )
are known. In other words, because Eq. 3.49 describes all 36.8%
possible solutions, it has no specific meaning until the val- 0
ues have been determined. Second, and more importantly, 0 τ 2τ 3τ 4τ 5τ
reliance on this formula obscures the essential point that time
the change of any variable within a first-order dynamic sys- −
t

tems is caused by the flow of energy into or out of the sys- Fig. 3.33   Decay to zero. V (t ) = ∆V e τ , where ∆V = V (0+ )
tem. For example, consider the step response of a grinding
wheel. Apply a step change increasing power to the grinding
wheel, and the wheel speeds up. There is a growth in speed
V(ss)
and, hence, in the kinetic energy of the wheel. We can hear 36.8%
the change in speed. Hence, we can hear the wheel gain-
ing energy. When the wheel gains energy, we know that we V(t) ∆V
can describe the step response of its energy storage variable, 63.2%

the wheel’s angular velocity, by a growth function. When it


loses energy, we can describe the wheel’s angular velocity 0
0 τ 2τ 3τ 4τ 5τ
by a decay function. If the wheel is rotating in the negative time
direction and gains energy, it makes more physical sense to
describe its step response as a growth to a negative steady- Fig. 3.34    Stable growth from zero. V (t ) = ∆V 1 − e ( −t
τ ), where
state value than as decay attributed to a negative steady- ∆V = V ( ss )
state value.
Understanding dynamic responses of physical systems
must be dimensionless (or unitless, if you will). Coefficient
requires moving beyond a purely formulaic approach. De-
A carries physical units.
veloping the ability to visualize response functions of a sys-
tem to step input allows one to make productive use of one’s Having railed on formulaic thinking, we will now pres-
understanding of dynamic system behavior, even when there ent the four possible types step responses of first-order
is insufficient information to formulate a mathematical re- systems formulaically, by normalizing the first-order step
sponse. responses. The output variable is V( t), which can be thought
of as ­either translational velocity or electrical voltage. The
vertical axis (ordinate) is normalized by dividing the out-
3.5.1 Categorization of First-Order Step put variable, V( t), by the range by the dynamic range, ∆V.
Responses The time axis is normalized by dividing time by the time
constant, τ. The four forms of the first-order step responses
Scaling and “normalizing” are inverse mathematical opera- are summarized below and shown in Figs. 3.33, 3.34, 3.35,
tions. To formulate a response function, we multiply an ex- 3.36, and 3.37,
ponential decay or stable growth function, which has a range −t

of zero to one, by a constant factor which scales the out-  V (t ) = ∆V e τ where ∆V = V (0+ ) (3.50)
put variable and gives it physical units. The constant is the
dynamic range of the output variable. Normalization is ac- (
V (t ) = ∆V 1 − e τ where ∆V = V ( ss )
(3.51)
−t
)
complished by dividing observed data by the same constant
factor to reduce its maximum value to unity and remove the The growth and decay curves are mirror images of each
physical units. The scaling and normalizing by multiplying other, as shown in Fig. 3.35, where the dependent variable,
or dividing by the dynamic range affects only the ordinate V( t), has been normalized by dividing it by its maximum
(vertical axis). The abscissa (time axis) is normalized the value, and time is expressed in time constants. Time equal to
time constant, τ, as shown in Eqs. 3.50 and 3.51, and then one time constant τ reduces the difference between the cur-
by measuring or expressing time in units of time constants. rent value and the steady-state value by 63.2 %.
Division of the time variable in the exponent by the time The addition of a constant to the growth and decay func-
constant is not an additional step. It is an essential
−t
part of the tions, Eqs. 3.50 and 3.51, offsets the growth and decay
homogeneous solution, vh (t ) = Ae st = Ae τ . The exponent curves vertically, Eqs. 3.52 and 3.53,
3.5  First-Order Step Responses 145

Fig. 3.35  a Normalized first-


order growth. b Normalized
a b
first-order decay v(ss)
____ v(0)
___
=1 =1
v(ss) v(0)

0.632
v(t)
____ v(t)
____ 0.865 0.950 0.982 0.993
0.950 0.982 0.993
v(ss) 0.865 v(0)
0.632

0 τ 2τ 3τ 4τ 5τ 0 τ 2τ 3τ 4τ 5τ
time time

V(0+) V(0+)
63.2%
63.2%
V(t) ∆V
36.8%
V(t) 0 ∆V
V(ss)
36.8%

0 V(ss)
0 τ 2τ 3τ 4τ 5τ
time 0 τ 2τ 3τ 4τ 5τ
−t
Fig. 3.36   Decay to a non-zero steady-state value. V (t ) = ∆V e τ + V ( ss ),
time
where ∆V = V (0+ ) − V ( ss ) Fig. 3.38   Stable exponential growth from a positive initial value to a
negative steady-state value

V(ss)
36.8%
i­ncreases the energy stored in an element as an increase,
V(t) ∆V and a change that decreases the amount of energy stored as
63.2%
a decrease. Consequently, we will define growth and decay
V(0+) relative to the magnitude of a variable. A variable grows if
its magnitude increases. Likewise, a variable decays if its
0 magnitude decreases. Using this definition, it is possible to
0 τ 2τ 3τ 4τ 5τ
time grow from a positive initial value to a negative steady-state
value, and to decay from an initial negative value to a posi-
Fig. 3.37   Stable exponential growth from a non-zero initial value. tive steady-state value.
(
V (t ) = ∆V 1 − e
−t
τ ) + V (0 ), where ∆V = V (ss) − V (0 )
+ +

−t 3.5.2 Step Response of a First-Order Mechanical


 V (t ) = ∆Ve + V( ss ) where ∆V = V(0 ) − V( ss )
τ +
(3.52) System

 (
V (t ) = ∆V 1 − e
−t
τ ) + V(0 )
+
(3.53)
We will use the force source-mass-damper system, Fig. 3.13,
as the example and solve the system equations for the veloc-
ity of the mass, Eq. 3.26, and force acting to accelerate the
Signs can create semantic problems when describing a dy- mass, Eq. 3.27, repeated below,
namic response. For example, the plot of V( t) in Fig. 3.38 is
concave upward, the shape of a decay from a positive value. F0 us (t ) M dv1g
(3.26) = + v1g
In the caption, the trace is described as a growth from an b b dt
initial positive value to a negative steady-state value. Either
description is acceptable. If the variable, V( t), was veloc- M dF0 us (t ) M dFM (3.27)
= + FM
ity and we were to observe the response of the system, we b dt b dt
would likely describe the response as the system reversing 
direction, since the initial and final magnitudes of V( t) are for the step input, F (t ) = F0 us (t ), plotted in Fig. 3.26.
the same. It is more consistent to describe a change that
146 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

The Heaviside step input transitions from zero to one be-


E sys ( 0− ) = 0 = E sys ( 0+ ) → E M ( 0− ) = 0 = E M ( 0+ )
fore the response can begin. Thus, we can remove the step
function, us( t), and leave just its scaling factor, F0,
1 1
F0 M dv1g M dF0 M dFM 2
( )
Mv12g 0 − = 0 = Mv12g 0+
2
( ) ( )
→ v1g 0 − = 0 = v1g 0+ ( )
= + v1g and = + FM
b b dt b dt b dt
If we know the value of the state variable at any time t, then,
Our solution method is to determine the initial and final val- also using the value of the input which we always know, we can
ues of the response. These values are sufficient to identify calculate any other power variable in the system at that time.
which of the four possible first-order step responses is the We can now identify the form of the step response. The
correct response function. initial value of the output variable is v1g (0+ ) = 0, and the
The time constant and the final, or steady-state, value are steady-state value is v1g ( ss ) = F0 / b . The step response is a
determined from the system equation. The time constant is stable exponential growth from zero of the form of Eq. 3.51,
found by unit analysis after expressing the system equation
in time constant form by clearing the output variable of its V (t ) = ∆V 1 − e( −t
τ )
coefficient. The factor multiplying the derivative of the out-
put variable with respect to time must have units of time for where
that term to have the units of the output variable,
 −t
M

F0  b 
∆V = VSS → v1g (t ) = 1− e
F0 M dv1g M b  
= + v1g → τ =  
b  b dt b
τ
Improper ratios are difficult to read and lead to errors.
The steady-state value is determined using the same logic Eliminating the improper ratio in the exponent yields the
and procedure as used in the method of undetermined coef- following:
ficients. It is a fact that in steady-state, all power variables
within a linear system will have the same functional form F0  − t
b
v1g (t ) =  1 − e M
 (3.54)
as the input. A Heaviside step input has a single transition.  b
After it has transitioned from zero to one, it remains con-
stant. Consequently, the steady state of all power variables in The solution for force FM, acting to accelerate the mass, fol-
a linear system driven by a step input is constant. Remember lows the same procedure as the solution for the response v1g.
that zero is a constant. Substitute v1g ( ss ) = C into the system The time constant and the steady-state value are determined
equation, and evaluate the derivative: from the system equation,

F0 M dv1g ( ss ) M dF0 M dFM


= + FM → τ=
M
= + v1g ( ss ) b dt b dt b
b b dt 
τ
F0 M dC F0
= +C → = C = v1g ( ss )
b b dt 0 b There is only one time constant for a first-order system. If the
time constant of Eq. 3.25 were not the same as the time con-
Calculation of the initial value of the output variables was stant of Eq. 3.26, then one or both of them would be in error.
presented in Sect. 3.4 and is only reviewed here. The system The steady-state value of FM is determined by substitut-
was described as de-energized before the input is applied. ing FM ( ss ) = C into the system equation and evaluating the
The mass is the only energy storage element in the system, derivative, as below:
and it stores kinetic energy. If the system has no energy, then
the velocity of the mass is zero. M dF0 M dFM ( ss )
= + FM ( ss )
1 b dt b dt
( ) ( )
v1g 0 − = 0 E sys 0 − = E M 0 − ( ) ( )
E M 0 − = Mv12g 0−
2
( )
M dF0 M dC
The energy storage variables are known as state variables. = +C → 0=C → FM ( ss ) = 0
b dt 0 b dt 0
We see that v1g is the state variable of the mass-damper sys-
tem. State variables are the only variables in a system which
cannot change instantaneously, The rate of change of energy stored in an element equals the
net power flow into and out of that element. If the energy
3.6  Time Shift 147

1 u s(t-t a )
No power flows into
the mass in the steady- 1
state of a step response.
All power flows into the
F(t) b M damper and is dissipated
as heat.
0 ta time

g Fig. 3.40   Time shifted heaviside unit step function. Subtracting ta


from t shifts the time the function transitions from 0 to 1 to time ta by
Fig. 3.39   Linear graph of the mass-damping system showing the keeping the argument negative
power flows in steady-state

tioned from zero to one, we first determined the value of the


stored in the mass is changing, then the system has yet to system’s state (energy storage) variable at time t = 0+ and
reach steady-state in response to the step input, found that v1g (0+ ) = 0, because the system was de-energized
at time t = 0 −. Then, using v1g (0+ ) = 0 and the value of the
dE dv1g dE input at t = 0+ , F (0+ ) = F0 us (0+ ) = F0 . We found
= v1g M =0 → = v1g FM ≡ P = 0 → FM = 0
dt dt dt
( )
F 0+ = b v1g 0+ ( ) 0
( )
+ FM 0+ → ( )
F 0+ = F0 = FM 0+ ( )
We can reach the same conclusion by considering the power
flows within the linear graph. If there is no change in the We can now identify which of the four step responses is the
amount of energy stored in the mass when the system has correct response. The initial value is F0 = FM (0+ ), and the
reached steady-state, then the power flowing into or out of steady-state value is FM ( ss ) = 0. The step response is an ex-
the mass equals zero. The power flowing from the source ponential decay to zero in the form of Eq. 3.43,
into the system, Psource = F (t )v1g , must all be dissipated by
shearing the lubricating fluid film, Pb = Fb v1g . In the steady- −t −t

state of a step response, all power that flows into node 1 from V (t ) = ∆V e τ where ∆V = V0 → FM (t ) = F0 e τ
the source also flows out into the damping element, b. No b
power flows into the mass in steady-state. (3.55) − t
FM (t ) = F0 e M
Nodes represent the value of the across variable velocity
and transmit force and power between elements. Nodes in Physically, the mass is at rest at the instant after the Heavi-
mechanical systems are ideally rigid and massless. Energy side step input is applied, since its kinetic energy cannot
is sourced, stored, or dissipated in the elements between the change instantaneously without infinite power. There is no
nodes, not at nodes. We can express conservation of energy shear of the lubricating fluid film, since there is no motion.
in terms of power if we sum energy flow rates at nodes. The Initially, all of the applied force acts to accelerate the mass.
assumed positive direction of power flows is in the direc- Eventually, the velocity of the mass will create a shear force
tion indicated by the arrowhead on each element, Fig. 3.39. equal to the applied force. The system will then be in steady-
­Summing power flowing into and out of node 1, where flow state. Since none of the applied force will act to accelerate
in is positive and flow out is negative, the mass, there will be no further change. All of the power
variables will have constant values.
Psource = Pb + PM → F0 us (t ) v1g = Fb v1g + FM v1g

When the system reaches steady-state, velocity v1g ( ss ) is 3.6  Time Shift
not zero. Therefore,
Time t = 0 is arbitrary and defined by the engineer typically
PM = FM v1g ( ss ) = 0 → FM = 0 at the beginning of a time period of interest. We need to ad-
dress the general case when a single step input is applied to
and a system but at a time other than t = 0. The step input and
the corresponding step response are formulated using a “time
F0 us (t ) v1g = Fb v1g + FM 0 v1g → F0 us ( ss ) = F0 = Fb shift,” in which the time at which the input is applied, say
time t = ta , is subtracted from the time variable, t, so that
argument of the function is t − ta , Fig. 3.40. Nothing more
Again referring to Sect. 3.4, in order to determine the initial needs to be done to time shift the Heaviside unit step func-
value of FM (0+ ) at the instant after the Heaviside step transi- tion, since us (t − ta ) is equal to zero when (t − ta ) ≤ 0.
148 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

v(ss) 3.7 Superposition of Heaviside Step


Functions

Combining two or more step functions allows construction


v(t) 0 This portion of the of system inputs that approximate any arbitrary function. A
computed response “pulse” input is created by summing, or superposing, a step
violates cause and
effect. It is before function of positive amplitude with a step function of nega-
the input acts on tive amplitude,
the system.
0 ta time Unit Pulse(t , ta ) = us (t ) − us (t − a )
(3.58)
Fig. 3.41   Time shifted unit step response, Eq. 3.56. Subtracting ta from
t shifts the time the function turns on to time ta. However, it is still The amplitude of a unit pulse is scaled by multiplying it by
possible to evaluate the function for times before the input is applied, a coefficient. An arbitrarily shaped function of time can be
violating cause and effect approximated as a sum of appropriately scaled rectangular
pulses, as illustrated in Figs. 3.44 and 3.45.
We often drive or force a system with pulses of varying
v(ss) amplitude or varying width, because we need the system to
respond to events which are separated in time, or we wish
to approximate the effect of a continuous input of some
v(t) 0 amplitude. The first case is described as a “pulse train,” in
Mutliplying the response
function by the Heaviside
which the pulses may or may not be of equal amplitude,
unit step with the same duration, and time-spacing, Fig. 3.46. The second case is
time shift zeros out this usually accomplished by “pulse-width-modulation,” in
portion of the response.
which the duration of the “on” portion of a pulse is scaled
0 ta by the desired equivalent DC or constant amplitude. Pulse-
time
width-modulation is useful because the energy storage ele-
Fig. 3.42   Time shifted step response multiplied by a time shifted ments in our systems accumulate or integrate the power
Heaviside unit step function, Eq. 3.57
pulses from the input source, and, as a result, smooth out
variation in the system’s power variables. Practical pulse-
Time shifting a response function is only slightly com- width-modulated amplifiers operate at pulse frequencies
plicated by the fact that a time shifted response function has which are at least an order of magnitude greater than the
a non-zero value if evaluated for time before the input that upper limit of human hearing.
created the response acted on the system. The result of such Pulses can also be used to approximate continuous func-
a calculation is clearly nonsense, since it violates cause and tions, such as sinusoids. Applying a concept familiar from
effect. However, it may not be immediately apparent from calculus, decreasing the duration (or width) of the pulses
the response plot alone that the result is nonsense, because increases the precision of the approximation, Fig. 3.47.
the plot will be a continuous exponential function. For ex- Rather than reducing the width to infinitesimal, we stop at
ample, the stable exponential growth, Eq. 3.56, plotted in a finite value to yield a discontinuous input function with a
Fig. 3.41, stair step appearance. A trapezoidal approximation is more
accurate but limited in its applicability. Real time systems,
F0  −   ( t − ta ) 
 b

v1g (t ) =  m such as digital controllers and “inverters,” which approxi-


1 − e  (3.56)
 b   mate the sinusoids of AC current from DC power, use a
“zero-order-hold” which holds the variable constant until
was shifted forward to time t = ta, but the function was plot- the next pulse and creates rectangular pulses. Pulses are
ted from time t = 0. The only way to spot the error would the basis of digital control with the zero-order-hold using a
be to know the correct initial value of the response function. pulse of unit amplitude constructed as shown in Fig. 3.43,
The step response function must be “zeroed out” for Eq. 3.58.
(t − ta ) ≤ 0, by multiplying it by a time shifted Heaviside Although we can superpose pulse responses to approxi-
unit step function, Eq. 3.57, Fig. 3.42, mate the response of a linear model to an arbitrary input,
as shown in Fig. 3.47, and then construct the response by
F  −   ( t − ta )  
 b
superposing the corresponding step pulse responses, it is not
v1g (t ) = us (t − ta )  1 − e  m 
  0
 (3.57)
  b    worth the computational effort. We will develop a simpler
and more powerful approach to calculate the response of
3.8  Superposition of First-Order Step Responses 149

Fig. 3.43  a Unit step functions,


us( t) and us( t − ta). b Sum or
a us(t) b us(t)
superposition of the two unit step
1 1
functions create a unit pulse of
width ta
us(t) - us(t-ta)

0 ta time 0 ta time

-1 -1
-us(t-ta) -us(t-ta)

Fig. 3.44  a Force pulse,


F(t) = 10N us(t) + 10N us(t − 1) a b
20 20
− 20N us(t − 2).
b Force steps superposed to create 10 us(t)
the pulse 10 10
10 us(t-1)
F(t) 0 F(t) 0
N 0 1 2 N 0 1 2 3
t, sec t, sec
-10 -10
-20 us(t-2)
-20 -20

Fig. 3.45  a Force pulse, a b 10 N u s(t)


F(t) = 10N us(t) − 15N us(t − 1) + 5N us(t − 3).
10 10
b Force steps superposed to create the
pulse 5 N u s(t-3)
5 5

F(t) 0 F(t) 0
N 1 2 3 N 1 2 3 4
t, sec t, sec
-5 -5

-10 -10
15 N us(t-1)
-15 -15

system equations to arbitrary inputs, after we introduce the


“state-space” representation of systems as sets of coupled
first-order differential equations. 1

Pulse(t)
Period T
3.8 Superposition of First-Order Step
Responses
0
0 0.5T T 1.5T 2T 2.5T
Linear equations can be viewed as “linear operators” since time
they “operate” on the input variable to yield the output vari-
able. Our differential system equations are linear operators. Fig. 3.46   Unit pulse train with 50 % duty cycle. Duty cycle is the per-
Linear operators have two valuable properties, scaling and centage of a period, T, during which the pulse has non-zero amplitude
150 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.47  a Arbitrary input


approximated by one second
a b
duration pulses. b Input more 40 40
accurately approximated by half F(t) F(t)
second duration pulses N 20 N 20

0 0
0 2 4 6 8 0 2 4 6 8
t, sec t, sec

Fig. 3.48  a Example mass- a b


damper system driven by a force x,v
source. b Pulse input of force F(t), N
F0
F(t) Mass Lubricating
M fluid film
Damping b
0 ta t, sec

time shifted step input is Eq. 3.57. Both equations are re-


F(t), N F0 us(t)
peated below,
F0
(3.54) F  − t
b
v1g (t ) = 0 1 − e M 
b  
0 ta t, sec F  −   ( t − ta )  
 b

v1g (t ) = us (t − ta )  1 − e  m 
  0
 (3.57)
-F0   b   
-F0 us(t-ta)
It must be emphasized that the output or response function
has the same time shift as the input function which created it,
Fig. 3.49   Pulse input of force plotted in Fig. 3.48b created by super-
posing two step inputs, Eq. 3.59 Fig. 3.50. Hence, time in the exponent of the step response
created by the step input at time t = ta must be shifted by sub-
tracting time ta from the time variable t to synchronize the
superposition. We can calculate the response of a system to response function with the input function. The time shifted
an input of unit amplitude, and then scale the unit input re- response function must be multiplied by a Heaviside step
sponse, to calculate the response of the system to an input of function with the same time shift, in order to zero-out the
any magnitude. We can use superposition to calculate the re- result of calculating the response, prior to the time the cor-
sponse of a system to combined inputs or inputs constructed responding input acts on the system.
from simpler functions, specifically step functions, by sum- The analytical expression of the superposition of step re-
ming the responses to each individual input. sponses is a straightforward sum, Figs.  3.51 and 3.52. It is
Example of Superposition, Response to a Pulse Input. important to remember that time shifted step response func-
Calculate the response of the mass-damper system of the tions must be multiplied by time shifted unit step functions,
previous example to a pulse input of force of magnitude F0, in order to zero-out (or null) the functions for time, t < ta,
as shown in Fig. 3.48.
F  −   t 
 b
We first construct the pulse input from the superposition
v(t ) = us (t )  0  1 − e  m  
of two step inputs, Fig. 3.49, Eq. 3.59,   b  

F (t ) = F0 us (t ) − F0 us (t − ta ) = F0 us (t ) − us (t − ta )  −F   −    (t − ta ) 


 b
(3.59) + us (t − ta )  0  1 − e  m 
 b   (3.60)

We then superpose (sum) the response to each of the inputs,
which we have already calculated. The system to a step input A Heaviside unit step function is used to zero-out the re-
of force, F0, is Eq. 3.54 and the response of the system to a sponse function term, until its corresponding input is ap-
plied. Again, if the Heaviside unit step function is not used as
3.8  Superposition of First-Order Step Responses 151

Response to positive stant of a linear system is independent of the amplitudes of


force step input
F0 the input and output variables. If this seems counter-intuitive,
b it is because the timescales of the responses of most systems
F0 we are familiar with in everyday life depend on the amplitude
F(t), N vb(t), m of the output. A system with a “time constant” (which is not
sec
ta constant) is non-linear. For example, drop a coin on a hard
0 0
time surface, and listen to it, as it rattles to rest. The frequency in-
creases, as the rattle amplitude decreases. If the rattling coin
-F0 were described by a linear system equation, it would have a
-F0 constant frequency. In contrast, a non-linear system does not
have a single oscillation frequency. It has a range of frequen-
Response to negative b
force step input cies. The timescaling of a non-linear system’s step response is
dependent on the amplitude of the output variable.
Fig. 3.50   Force step inputs (left axis) and the step responses (right
axis)
3.8.1 Scaling, Time shifting, and Superposing
Superposition (sum) of Unit Step Responses
Response to positive
the two step responses force step input
F0
A particularly efficient method of using superposition to cal-
v1g(t) b
culate the response of a first-order linear system to an arbi-
m trary pulse input is as follows:
sec 1. Construct the input function by scaling, time shifting, and
ta
superposing (summing) Heaviside step functions.
0 2. Assume the system is de-energized, and determine the
t, sec
response of the output variable to an input step of unit
magnitude.
3. Scale, time shift, and superpose (sum) the unit step re-
-F0
sponse, using the same time shifting and scaling, as used
b
Response to negative to construct the input pulse from Heaviside step func-
force step input
tions. Multiply each term in the response function by the
same time shifted Heaviside step function used, to create
Fig. 3.51   Superposition of the positive and negative step responses
showing the summation the input pulse function, Fig. 3.53.
Scaling and superposition of the Heaviside step inputs to
construct the input pulse and its corresponding unit step re-
v1g(t) sponses, in order to construct the response function, are pos-
m
___ sible only if we restrict models of energetic attributes of the
sec system to linear models, as we have done.
We will illustrate the technique using the force source-
spring-damper system of Sect. 3.2.4.1, Fig. 3.54. The en-
0
0 ta t, sec ergetic equations and the system equations for the force
through the spring, FK, Eq. 3.29, and the force through the
Fig. 3.52   Pulse response. Growth during the pulse from t = 0 to t = ta, damper, Fb, Eq. 3.30, are reproduced below for convenience.
followed by decay after time, t = ta
Energetic Equations

a multiplicative factor with the corresponding step response Continuity or Node Eq: F (t ) − Fb − FK = 0
function term of the same time shift, then the result includes
the evaluation of the step response for the entire period of Compatibility or Path Eq: v1g = v1g
calculation, rendering said result “non-causal” for its viola-
tion of “cause and effect.” dFK
Element Eqs: Fb = bv1g = Kv1g
Both step responses have the same time scaling, because dt
the time constant of a system depends on the system’s param-
eters, and how the elements are interconnected. The time con- FK2
Energy Eqs: E sys = E K EK =
2K
152 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.53   Block diagram F(t)


showing the pulse input–pulse
v1g(t)
response, as an input–output
relationship

0 T time 0 T time
Linear
Pulse Input System Pulse Response

Fig. 3.54  a Energetic schematic


of a force source acting on an
a x,v b
ideal rigid and massless bar, 1
which is connected to a spring,
K, and dashpot, b. b Linear graph
K
of the system shown schemati-
cally in a
F(t) F(t) b K
1 b g
g

g
Rigid and massless bar

Fig. 3.55  a Input force pulse.


b Scaled and time shifted
a b
20 20
Heaviside step functions which
superpose to create the pulse in a 10 us(t)
10 10
10N us(t-2)
F(t) F(t) 0
0
N 0 1 2 3 N 0 1 2 3 t, sec
t, sec
-10 -10

-20 -20
-20N us(t-1)

b dFK
 F (t ) = + FK (3.29) from earlier inputs will be accounted for by response func-
K dt
tions time shifted to those earlier times.
b dF (t ) b dFb We begin by finding the unit step response of the force
(3.30) = + Fb acting through the spring, FK. We can establish the time con-
K dt K dt
stant and the steady-state value of the output variable from
the system equation in time constant form,
We will solve these system equations for the force input in
the force pulse, shown in Fig. 3.55a, which is created by b dFK b
F (t ) = + FK → τ =
scaling and superposing Heaviside unit step functions, as K dt
 K
shown in Fig. 3.55b. τ
The input function is The system is in steady-state when all of the power variables
in the system have the functional form of the input, which for
F (t ) = 10 N us (t ) − 20 N us (t − 1) + 10 N us (t − 2) (3.61) a step input is constant in steady-state. Hence, the derivative

of FK is zero in steady-state, since the spring force is constant:
To use superposition, the unit step response must be derived
by assuming that the system is de-energized, even if the sys- b dFK ( ss )
1 N u s ( ss ) = + FK ( ss ) → 1 N = FK ( ss )
tem is clearly not de-energized. The energy in the system 1 K dt 0
3.8  Superposition of First-Order Step Responses 153

This result tells us that the spring carries all of the applied 10
force in steady-state.
We now establish the initial value of the output variable, 5
at the instant after the step of 1 N has been applied to the sys-
tem. Again, we must assume that the system is de-energized
before the unit step is applied to the system. If a system is de- FK(t), N 0
energized, then its energy storage (state) variables equal zero,

E sys ( 0− ) = 0 = E K ( 0 − ) → E K ( 0− ) =
( )→F
FK2 0 − -5

2K
K (0 ) = 0

-10
0 1 2 3 4 5
A state (energy storage) variable cannot change instanta- t, sec
neously without an infinite power flow to transfer a finite
Fig. 3.56   Response function, FK( t), evaluated using K = 5,000 N/m
amount of power during an infinitesimal time. Consequently, and b = 2,500 N · sec/m
the state variable, FK, has the same value at time, t = 0 − , be-
fore the input step is applied as it does at time, t = 0+, at the
instant after the step is applied, There is only one time constant in a first-order system. If
we had arrived at a different time constant than that for the
( )
FK 0 − = 0 = FK 0+ ( ) spring force in the same system, one of the two (or both)
of the time constants would be erroneous. In steady-state,
We identify the form of the first-order step response from the output variable must have the form of the input variable,
FK (0+ ) = 0 and FK ( ss ) = 1 N as a stable exponential growth which, in steady-state, is a constant,
from zero with the form of Eq. 3.53 and Fig. 3.34,
b d1N us ( ss ) b dFb ( ss )
= + Fb ( ss ) → Fb ( ss ) = 0
(
FK (t ) = 1N 1 − e τ
−t
) → FK (t ) = 1N 1 − e ( −
K
b
t
) K dt 0 K dt 0

u .s. u .s.

The initial value is determined, using the values of the unit


We construct the pulse response function by weighting (scal- input and the state (energy storage) variable, FK, at time
ing) the unit step response function with the same scaling t = 0+ and the energetic equations. We know F (0+ ) = 1N and
factors and time shifts used to construct the input function. FK (0+ ) = 0. We need Fb (0+ ). Using continuity,
Both the exponents of the exponentials and the Heaviside
unit step function must be time shifted. Each term in the ( ) ( ) ( )
F 0+ − Fb 0+ − FK 0+ = 0 → F 0+ − FK 0+ = Fb 0+ ( ) ( ) ( )
response function is multiplied by the corresponding term
in the input function. The units must be removed from the
input function because they already appear in the unit step ( )
Fb 0+ = 1 N
response function,
Comparing Fb (0+ ) = 1 N with Fb ( ss ) = 0, we identify the
(
FK (t ) = 1N 1 − e )10 u (t ) − 1N (1 − e ) 20 u (t − 1)
K
− t
b
s
K
− (t −1)
b
s
unit step response function as a decay to zero in the form, as
expressed in Eq. 3.50 and Fig. 3.33,

+ 1N (1 − e

)10 u (t − 2)
K
b
(t − 2 )
s
Fb (t ) = 1N e τ
−t
→ Fb (t ) = 1N e

K
b
t

u .s. u .s.

We will use the parameter values K = 5, 000 N/m and


b = 2,500 N · sec/m to evaluate the response, plotted in We construct the pulse response function by weighting (scal-
Fig. 3.56. ing) the unit step response function with the same scaling
We will now repeat the process to determine the response factors and time shifts used to construct the input function,
of the damper force, Fb, to the force pulse input. We find the Eq. 3.61, F (t ) = 10 N us (t ) − 20 N us (t − 1) + 10 N us (t − 2),
time constant and the steady-state response to a step input of
K K
1 N from the system equation, − t − (t −1)
 Fb (t ) = 1Ne b
10us (t ) − 1Ne b
20us (t − 1)
(3.62)
b dF (t ) b dFb b −
K
(t − 2 )
= + Fb → time constant τ = + 1Ne b
10us (t − 2)
K dt K
 dt K
τ
154 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

10 response is obtained. First-order systems have only four


possible step responses. The correct unit step response can
be identified on the basis on the initial and steady-state
0 values. Second-order step responses are a continuum of re-
sponses, from those which resemble the step response of a
Fb (t), N first-order system, to those which are almost sustained os-
cillations. The increased variation in potential responses is
-10
due to the possibility of internal energy transfer between
the two independent energy storage elements, the spring
and the mass. Whether or not a step response is oscillatory
-20 depends on the relative energy storage capacity of the two
0 1 2 3 4 5
t, sec storage elements, and on the rate at which energy is dis-
sipated as heat, during the flow of energy between the two
Fig. 3.57   Response function, Fb( t), evaluated using K = 5,000 N/m storage elements. If the eigenvalues (characteristic values)
and b = 2,500 N · sec/m of the system are purely real, then the step response is over-
damped or non-oscillatory. Conversely, if the eigenvalues
are complex conjugates, then the step response is under-
The damper force picks up the instantaneous changes of the damped or “oscillatory.” Oscillatory is in quotation marks,
input force, which are then transferred to the spring as the because one normally thinks of an oscillation as consisting
spring deforms (Fig. 3.57). of at least one complete cycle. An underdamped system’s
step response will overshoot or undershoot the final steady-
state value, but it may not complete an entire cycle.
3.9 Superposition of Second-Order Step We will solve the system equation for the force acting
Responses through the spring, Eq. 3.33, using the method of undeter-
mined coefficients first for parameter values that yield an
We will demonstrate the superposition of second-order step overdamped or non-oscillatory response. We shall then do
responses using the force source-spring-mass-damper system likewise for parameter values which yield an underdamped
of Sect. 3.2.4.3, shown as an energetic schematic and a linear or oscillatory response.
graph in Fig. 3.31, reproduced below for reference, with the
energetic equations and the system equations for the spring
force, FK, Eq. 3.33, and the velocity of the mass, v1g, Eq. 3.34. 3.9.1 Overdamped or Non-Oscillatory Step
Response
Energetic Equations
We will use the parameter values K = 5, 000 N/m,
Continuity or Node Eq: F (t ) = Fb + FK + FM b = 2,500 N · sec/m, and M = 200 kg for the overdamped,
non-oscillatory case. Substituting these values into the sys-
Compatibility or Path Eq: v1g = v1g tem equation for the force acting through the spring, Eq. 3.33,

dFK dv1g K d 2 FK b dFK K


Element Eqs: Fb = bv1g = Kv1g FM = M F (t ) = + + FK
dt dt M dt 2
M dt M
FK2 1 N N · sec N
Energy Eqs: E sys = E K + E M EK = EM = Mv12g 5, 000 2,500 5, 000
2K 2 m F (t ) = d 2
F m dF mF
K
+ K
+ K
200 kg dt 2 200 kg dt 200 kg
K d 2 FK b dFK K
F (t ) = + + FK (3.33) 25 d 2 FK 12.5 dFK 25
M dt 2
M dt M 2
F ( t ) = 2
+ + FK
sec dt sec dt sec 2
 2
1 dF (t ) d v1g b dv1g K (3.34) Particular Solution We will determine the particular so-
= + + v1g
 M dt dt 2 M dt M lution for a unit step input of force, F (t ) = 1N us (t ). The
particular solution is the steady-state solution, after the ho-
The methodology differs from the superposition of first- mogeneous response has decayed to zero. The steady-state
order step responses only in the manner that the unit step response for all power variables in a system is the functional
3.9  Superposition of Second-Order Step Responses 155

Fig. 3.31  a Force source spring-


a b 1
mass-damper system of Fig. 3.23.
b Fig. 3.24b
x,v

1 K F(t) b K M
F(t) g
M
g

g g
Lubricating fluid
Damping b

form of the input variable. If the input is a Heaviside step, General Solution Sum the particular and homogeneous so-
then in steady-state all power variables will be constants, lutions to form the general solution:

25 d 2 FK ( ss ) 12.5 dFK ( ss ) 25 FK (t ) = FK P ( ss ) + FK H (t ) → FK (t ) = 1N+A1e s1t + A2 e s2t


1N u s( ss ) = + + FK ( ss )
P P

sec 2 dt 2 0 sec dt 0 sec 2 P

Use the initial conditions to determine the coefficients A1


FK P ( ss ) = 1 N. and A2. In order to use superposition, we must assume that
the ­system is de-energized prior to the a­ pplication of the first
Homogeneous Solution Form the homogeneous equation input. The initial conditions needed to solve this equation are
by setting the input to zero: the values of the output variable, FK (0+ ), and its derivative,
dFK (0+ ) /dt , at time t = 0+ , the instant after the step input
d 2 FK 12.5 dFK 25 is applied to the system. The output variable, FK, is a state
0= + + FK
dt 2
sec dt sec 2 (energy storage) variable. If the system is de-energized at
time t = 0 −, the energy storage variables equal zero. The state
Perform the Laplace transformation (neglecting initial con- variables cannot change instantaneously, since time is need-
dition terms) to form the characteristic equation. We will ed to move energy into or out of the energy storage elements.
now drop the units of sec. Therefore, the state variables remain zero at time t = 0+ , and

 d 2 FK dFK 
( )
FK 0+ = 0
L {0} = L  2 + 12.5 + 25 FK 
 dt dt 
In Sect. 3.4.2, we established
0 = s 2 FK ( s ) + 12.5sFK ( s ) + 25 FK ( s )
dFK 0+ ( ) = Kv
(0 )+

(
0 = s + 12.5s + 25 FK ( s )
2
) dt
1g

0 = s 2 + 12.5s + 25 The velocity of the mass, v1g, is the other state variable of the
system. Hence, v1g (0+ ) = 0 and
Solve the characteristic equation to determine eigenvalues
of the system: ( ) = Kv
dFK 0+
dt
1g (0 ) = 0
+

−12.5 ± 12.52 − 4.25


s1 , s2 =
2 Apply the initial condition, FK (0+ ) = 0 , to the general
solution:
s1 , s2 = −6.25 ± 3.75 → s1 = −2.5, s2 = −10
( )
FK 0+ = 0 → 0 = 1N + A1 e s1 ⋅ 0 1 + A 2 e s2 ⋅ 0 1
The eigenvalues are real. The system is overdamped, and the
step response is non-oscillatory. 0 = 1N + A1 + A 2
Form the homogeneous solution

FK H (t ) = A1e s1t + A2 e s2t


156 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

dFK (t )
Apply the initial condition, = 0, to the general
dt 1.0
solution:
dFK (t ) d
dt
=
dt
(
1 N+A1e s1t + A 2 e s2t )
FK(t), N 0.5
dFK (t ) d1N de s1t de s2t
= + A1 + A2
dt dt 0 dt dt

dFK (t ) ds1t ds t 0
= A1e s t 1
+ A2 es t 2
2
0 1 2 3
dt dt dt t, sec

dFK (t ) dt dt Fig. 3.58   Unit step response function, FK( t), evaluated using
= A1e s t s1
1
+ A 2 e s t s2
2

dt dt 0 dt 0
K = 5,000 N/m, b = 2,500 N · sec/m and M = 200 kg

dFK (t )
= A1e s1t s1 + A 2 e s2t s2 This is the unit step response and is plotted in Fig. 3.58.
dt
The overdamped second-order unit step response,
Evaluate for t = 0+ : Fig. 3.58, strongly resembles the first-order stable exponen-
tial growth, shown in Fig. 3.34. A second-order step response
dFK (t ) differs from a first-order step response by the “reverse”
=0 → 0 = A1 e s1 ⋅0 1 s1 + A 2 e s2 ⋅ 0 1 s2 curve at the base or beginning, seen in both the unit step
dt
response and the pulse response of the overdamped second-
0 = A1 s1 + A 2 s2 order spring-mass-damper system.
We now construct the pulse response function by weight-
Form a system of two equations in the unknowns A1 and A2. ing (scaling) the unit step response function,
Solve the system for A1 and A2. We will use direct elimination:
FK (t ) = 1N − 1.33Ne −2.5t + 0.333Ne −10t
0 = s1 A1 + s2 A2 s s
→ A1 = − 2 A2 → − 1N = − 2 A2 + A2
−1N = A1 + A2 s1 s1 with the same scaling factors and time shifts used to con-
struct the input function, Eq. 3.61,
 s   s s
−1N =  − 2 + 1 A2 → −1N =  − 2 + 1  A2 F (t ) = 10N us (t ) − 20N us (t − 1) + 10N us (t − 2)
 s1   s1 s1 

The most common error is to omit the time shift in one or


s1 − s2 − s1
−1N = A2 → A2 = 1N more of the exponential terms when constructing the re-
s1 s1 − s2
sponse function, Fig. 3.59:
s2 s − s1 s2
A1 = −
s1
A2 → A1 = − 2
s1 s1 − s2
1N → A1 =
s1 − s2
1N (
FK (t ) = 1N − 1.33Ne −2.5t + 0.333Ne −10t 10us (t ) )
(
− 1N − 1.33Ne −2.5(t −1)
)
+ 0.333Ne −10(t −1) 20us (t − 1)
Substitute for the eigenvalues, s1 = −2.5, s2 = −10: + (1N − 1.33Ne −2.5(t − 2)
+ 0.333Ne −10 (t − 2)
)10u (t − 2)
s

s2 −10
A1 = 1N → A1 = 1N → A1 = −1.33 N 3.9.2 Underdamped or Oscillatory Step
s1 − s2 −2.5 − ( −10)
Response
− s1 − ( −2.5) N
A2 = 1N → A2 = 1N → A2 = 0.333N We will use the parameter values, K = 5, 000 ,
s1 − s2 −2.5 − ( −10) N · sec m
b = 2,500 , and M = 400 kg, for the underdamped,
m
Substitute into the general solution: oscillatory case. Substitute these values into the system
equation for the force acting through the spring, Eq. 3.33,
FK (t ) = 1N+A1e s1t + A2 e s2t
K d 2 FK b dFK K
F (t ) = + + FK
FK (t ) = 1N − 1.33Ne −2.5t + 0.333Ne −10t M dt 2 M dt M
3.9  Superposition of Second-Order Step Responses 157

10 Increasing the mass from 200 kg to 400 kg changed the ei-


genvalues from a real pair to complex conjugates. When we
5 express the quadratic equation in terms of the system’s pa-
rameters,

FK(t), N 0 K d 2 FK b dFK K
F (t ) = 2
+ + FK
M dt M dt M
-5
2
b  b K
-10 − ±   − 4·
0 1 2 3 4 M M M
s1 , s2 =
t, sec 2
Fig. 3.59   Response FK( t) due to the pulse input evaluated using
K = 5,000 N/m, b = 2,500 N · sec/m and M = 200 kg we see that increasing mass M reduced both the ratio of the
damping coefficient b to mass M and ratio of the spring con-
stant K to mass M.
N N · sec N The ratio between damping and mass is due to the physical
5, 000 d 2
F 2,500 dF 5, 000 mechanisms of energy dissipation and kinetic energy storage.
m F (t ) = K
+ m K
+ mF
400 kg dt 2 400 kg dt 400 kg
K Energy is dissipated as heat due to shear motion. Increasing
the mass of the system reduces the velocity v1g needed to store
12.5 d 2 FK 6.25 dFK 12.5 an amount of energy, decreasing the rate of kinetic energy lost
2
F (t ) = + + FK as heat. For an oscillation to occur, the energy stored in one
sec dt 2 sec dt sec 2
independent energy storage mode must flow into to the other
Particular Solution Determine the particular solution for a mode and back again. The ratio of the spring constant K over
unit step input of force, F (t ) = 1N us (t ) : mass M governs the relative magnitudes of the energy stor-
age (state) variables of the system and the system’s ideal, un-
damped, natural frequency, ωn. Increasing the spring constant
12.5 d 2 FK P ( ss ) 6.25 dFK P ( ss ) 12.5
2
1N u s ( ss ) = 2
+ + FK ( ss ) K allows the system’s energy to be transferred from kinetic
sec dt sec dt sec 2 P
0 0
to elastic strain energy with less displacement of the mass,
decreasing the time needed to transfer kinetic to strain energy,
FK P ( ss ) = 1N
and, thereby, increasing the frequency of the oscillation.
We now form the homogeneous solution,
Homogeneous Solution Form the homogeneous equation
by setting the input to zero: FK H (t ) = A1e s1t + A2 e s2t

d 2 FK 6.25 dFK 12.5 The symbolic form of the homogeneous solution is the same
0= + + FK
dt 2 sec dt sec 2 for all second-order systems. The solutions differ in the val-
ues of the eigenvalues and the undetermined coefficients.
Perform the Laplace transformation (neglecting the initial General Solution Sum the particular and homogeneous
condition terms) and form the characteristic equation. We solutions for the general solution:
will now drop the units of sec.
FK (t ) = FK P ( ss ) + FK H (t ) → FK (t ) = 1N+A1e s1t + A2 e s2t
 d 2 FK dFK 
L {0} = L  2 + 6.25 + 12.5 FK 
 dt dt  The general solution in symbolic form is also identical to
0 = s FK ( s ) + 6.25sFK ( s ) + 12.5 FK ( s )
2 that of the overdamped case. We established the initial con-
ditions in Sect. 3.9.1 to be the following:
(
0 = s 2 + 6.25s + 12.5 FK ( s ) ) dFK (t )
FK (t ) = 0 and =0
2
0 = s + 6.25s + 12.5 dt

Solve the characteristic equation to determine the eigenval- We applied the initial conditions to the general solution of
ues of the system: the overdamped case and found the following:

−6.25 ± 6.252 − 4 ·12.5 s2 − s1


s1 , s2 = → s1 , s2 = −3.13 ± j1.65 A1 = 1N and A2 = 1N
2 s1 − s2 s1 − s2
158 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

These results for A1 and A2 apply for all eigenvalues s1 jω s-plane


and s2. We shall see that the overdamped and the under- s1 = -3.13 + j1.65
j1.65
damped solutions are quite different because their eigen-
values differ. The eigenvalues of this underdamped case are |s |
1 =3 j
s1 , s2 = −3.13 ± j1.65. .54
We will defer substituting in numerical values as long
φ1= 2.66 rad
as possible. Our first substitutions will be s1 = σ + jw and
s2 = σ − jw , -4 -3.13 -2 -1 σ
s2 σ − jw
φ2
A1 = 1N → A1 = 1N
s1 − s2 σ + jw − (σ − jw ) -j

σ − jw
A1 = 1N -j1.65
j 2w s 2 = -3.13 - j1.65
-j2

− s1 − (σ + jw ) Fig. 3.60   Eigenvalues, s1 = −3.13 + j1.65 and s2 = −3.13 − j1.65, plot-


A2 = 1N → A2 = 1N
s1 − s2 σ + jw − (σ − jw ) ted as vectors in the s-plane, s = σ + jw

− (σ + jw )
A2 = 1N Express the eigenvalues as complex exponentials by calcu-
j 2w lating their angle (arguments) and magnitude (modulus). Al-
ways sketch complex numbers to ensure that their angles are
Substitute these expressions for A1 and A2 and the symbolic calculated correctly, Fig. 3.60,
form of the eigenvalues into the general solution:
s1 = −3.13 + j1.65 = 3.54 e j 2.66
FK (t ) = 1N + A1e + A2 e
s1t s2 t

and
σ − jw − (σ + jw ) s1 = −3.13 − j1.65 = 3.54 e − j 2.66
FK (t ) = 1N + 1N e(σ + jw )t + 1N e(σ − jw )t
j 2w j 2w
1Ne −3.13t  3.54e − j 2.66 j1.65t 3.54e j 2.66 − j1.65t 
FK (t ) = 1N + e − e
Distribute time t into the exponents: 1.65  j2 j2 

σ − jw − (σ + jw ) Factor out a magnitude of 3.54. Place the remaining terms


FK (t ) = 1N + 1N eσ t + jwt + 1N eσ t − jwt
j 2w j 2w over the common denominator, j2:

Use the property of exponentials, e a + b = e a eb : 3.54 N e −3.13t  e − j 2.66 e j1.65t − e j 2.66 e − j1.65t 
FK (t ) = 1N +  
1.65 j2
σ − jw −(σ + jw )
FK (t ) = 1N + 1Neσ t e jwt + 1Neσ t e − jwt
j 2w j 2w Use the property of exponentials, e a eb = e a + b:
σt
Factor out 1Ne :  e j1.65t − j 2.66 − e − j1.65t + j 2.66 
w FK (t ) = 1N + 2.15 N e −3.13t  
 j2
1Neσ t  σ − jw jwt σ + jw − jwt  Factor the imaginary number j out of the exponents:
FK (t ) = 1N + e − e 
w  j 2 j2 
 e j (1.65t − 2.66) − e − j (1.65t − 2.66) 
FK (t ) = 1N + 2.15 N e −3.13t  
The parenthetical quantity resembles the form we need to  j2
use Euler’s sine formula, Eq. 2.54. We now substitute in val-
ues for the eigenvalues, s1 = σ + jw , and s2 = σ − jw , Use Euler’s sine formula, Eq. 2.54, to eliminate the complex
numbers,
s1 = −3.13 + j1.65 and s2 = −3.13 − j1.65
e jθ − e − jθ
(2.54) = sin (θ )
FK (t ) = 1N j2
1N e −3.13t  −3.13 − j1.65 j1.65t −3.13 + j1.65 − j1.65t 
+ e − e
1.65  j2 j2  FK (t ) = 1N + 2.15 N e −3.13t sin (1.65t − 2.66)
3.9  Superposition of Second-Order Step Responses 159

Particular Solution Determine the particular solution for


1.0
a unit step input, where force, F (t ) = 1N us (t ), remains un-
changed, since it is the steady-state response:

12.5
1N us ( ss ) =
FK(t), N 0.5 sec 2
d 2 FK P ( ss ) 6.25 dFK P ( ss ) 12.5
2
+ + FK ( ss )
dt 0 sec dt 0 sec 2 P
0
0 1 2 3 FK P ( ss ) = 1N
t, sec
Homogeneous Solution Form the homogeneous equation.
Fig. 3.61   Underdamped unit step response. The eigenvalues are Drop the units of sec. Perform the Laplace transformation,
s1 , s2 = −3.13 ± j1.65. The response overshoots approximately 0.1 % of
the steady-state value
neglecting initial condition terms. Set the characteristic func-
tion to zero to yield the characteristic equation:

This is the unit step response, plotted in Fig. 3.61. This un- 0 = s 2 + 1.25s + 12.5
derdamped response appears overdamped. The response
overshoots the steady-state value by a mere 0.1 %. It does Solve the characteristic equation to yield the eigenvalues:
not oscillate; yet the unit step response has a sine factor with
a frequency of 1.65 rad/sec. −1.25 ± 1.252 − 4 ·12.5
The lack of oscillation is due to the relative magnitudes of s1 , s2 = → s1 , s2 = −0.625 ± j 3.48
2
the real and imaginary components of the eigenvalues. The
decay rate is the real component sigma, σ = −3.13. Sigma is The time constant of the decay envelop described by these
approximately twice the magnitude of the angular frequency, eigenvalues is
w = 1.65 rad/sec. In the time necessary to complete an oscil-
lation, the period as T, 1 1
= τ deacy → = 1.6 sec
σ −0.625


T= = = 3.81sec
w 1.65 rad and the period of the oscillation is
sec
2π 2π
=T → = 1.8 sec
σt
the real exponential decay, e , has progressed to w rad
3.48
e −3.13T = e( −3.13)(3.81) = 6.62 × 10 −6 . The energy in the system is sec
dissipated too quickly for the system to complete a cycle of
oscillation. which will allow the oscillation to persist for five cycles.
We will reduce the damping in the system from Again, the homogeneous solution of all second-order sys-
b = 2,500 N · sec/m to b = 500 N · sec/m , while keeping tems is the following:
K = 5, 000 N/m and M = 400 kg, in order to create a more
oscillatory underdamped response. Substitute these values FK H (t ) = A1e s1t + A2 e s2t
into the system equation for the force acting through the
spring, Eq. 3.33, as shown below: General Solution The general solution is also identical to
the previous one:
K d 2 FK b dFK K
F (t ) = 2
+ + FK FK (t ) = 1N + A1e s1t + A 2 e s2t
M dt M dt M

Apply the Initial Conditions. We find the same form of the


N N · sec N
5, 000 2 500 5, 000 undetermined coefficients, A1 and A2, expressed in terms of
m F (t ) = d F m dF mF
K
+ K
+ K the eigenvalues written symbolically as s1 , s2 = σ ± jw ,
400 kg dt 2 400 kg dt 400 kg
σ − jw − (σ + jw )
A1 = 1N and A2 = 1N
j 2w j 2w
160 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

jω s-plane
j4
s 1 = -0.625+ j3.48 1.5
j3.48

j3
1.0
FK(t), N
|s1| = 3.54
j2
0.5

j
0
φ1= 1.77 rad 0 2 4 6 8 10
t, sec
-2 -1 -0.625 1 2 σ Fig. 3.63   Underdamped unit step response. The eigenvalues are
s1 , s2 = −0.625 ± j 3.48
Fig. 3.62   The eigenvalue, s1 = −0.625 + j 3.48, plotted as a vector in
the s-plane, s = σ + jw

Substitute into the unit step response:


Substitute these expressions for A1 and A2, as well as the
symbolic form of the eigenvalues into the general solution: 1N e −0.625t 3.54 e − j1.75 e j 3.48t − 3.54 e j1.75 e − j 3.48t
FK (t ) = 1N +
3.48 j2
σ − jw − (σ + jw )
FK (t ) = 1N + 1N e(σ + jw )t + 1N e(σ − jw )t
j 2w j 2w Factor out the magnitude 3.54 and use the property of expo-
nentials, e a + eb = e a + b:
Distribute time, t, into the exponents, using the property of
1Neσ t 3.54 N e −0.625t e j 3.48t − j1.75 − e − j 3.48t + j1.75
exponentials, e a + b = e a eb . Factor out , and place over FK (t ) = 1 N +
w 3.48 j2
a common denominator to yield the following:
e j (3.48t −1.75) − e − ( j 3.48t − j1.75)
1N e σt
(σ − jw ) e jw t
− (σ + jw ) e − jw t FK (t ) = 1N + 1.02 N e −0.625t
FK (t ) = 1N + j2
w j2
We now have the proper form to use Euler’s sine formula,
We now substitute in values for the eigenvalues s1 = σ + jw
and s2 = σ − jw : e jθ − e − jθ
sin(θ ) =
j2
s1 = −0.625 + j 3.48 and s2 = −0.625 − j 3.48
e j (3.48t −1.75) − e − ( j 3.48t − j1.75)
= sin(3.48t − 1.75)
j2

1N e −0.625t ( −0.625 − j 3.48) e − ( −0.625 + j 3.48) e − j 3.48t


j 3.48t

FK (t ) = 1N +
3.48 j2

Express the eigenvalues as complex exponentials. Calculate yielding the purely real unit step response function, FK (t ),
the magnitude (modulus) and angle (argument), Fig. 3.62: below, and plotted in Fig. 3.63, u .s.

s1 = s2 = −0.625 + j 3.48 = ( −0.625)2 + 3.482 = 3.54 FK (t ) = 1N + 1.02 N e −0.625t sin (3.48t − 1.75)
u .s.

 3.48 
The angle (argument) is π − tan −1  = 1.75 rad, Construct the pulse response function by superposing unit
 0.625 
step responses which are scaled and time shifted with the
s1 = −0.625 + j 3.48 = 3.54e j1.75 same scaling and time shifting as the input function. The
input function, F( t), Eq. 3.61, shown in Fig. 3.55, is
and
s2 = −0.625 − j 3.48 = 3.54e − j1.75 F (t ) = 10N us (t ) − 20N us (t − 1) + 10N us (t − 2)
3.10  Initial Condition of Energized Systems 161

20
F(t), N
10 F2 = 250

0
FK(t), N F1 = 100
-10

-20 0
time
-30 Fig. 3.65   Representation of a previously applied step input. The zig-
0 2 4 6 8 10
t, sec zags indicate discontinuities in the time axis

Fig. 3.64    Response of the spring force, FK( t), in the under-


damped spring-mass-damper system with the eigenvalues, the use of zigzags on the horizontal lines, representing F1 (t )
s1, s2 = −0.625 ± j 3.48 , to the pulse input shown in Fig. 3.55 and F2 (t ). This graphic device implies extensions of the time
axis, further into the past and future, respectively, with no
change in their values.
All arguments in each term of the pulse response function We can calculate the response of the system to the step
must be time shifted by the same amount, input applied at time t = 0 using superposition. The input

( ) ( )
FK (t ) = 1N + 1.02 N e −0.625t sin (3.48t − 1.75) 10 us (t ) − 1N + 1.02 N e −0.625(t −1) sin (3.48 (t − 1) − 1.75) 20 us (t − 1)

( )
+ 1N + 1.02 N e −0.625(t − 2) sin (3.48 (t − 2) − 1.75) 10 us (t − 2)

The pulse response is plotted in Fig. 3.64.


function, F( t), can be constructed by superposing (sum-
ming) two step inputs but we must first estimate the mini-
3.10  Initial Condition of Energized Systems mum duration before time t = 0 the first input could have
been applied such that the system reached steady-state by
Our previous calculations assumed that the system was de- time t = 0.
energized, prior to the application of a step input. This was The largest time constant in a system, τmax, or, equiva-
also true, when we used the superposition of multiple step lently, the smallest magnitude real eigenvalue component,
inputs to create pulses, or to approximate an arbitrary input. σ min, in the system determines the duration of the transient
We assumed that the system was de-energized, prior to the period of a step response. Although mathematically, an
application of the first step input, and we assumed that all exponential decay never reaches zero, in reality, physical
corresponding superposed step responses started from a de- systems reach steady-state once their current value is indis-
energized state. Otherwise, superposition cannot be used. tinguishable from the final value. When a system reaches
The remaining case to consider is that of a system energized steady-state depends on the precision of the measurement,
before a step input is applied to it at time t = 0. It is com- as well as the noise and ripple in the response signal, which
mon for a system to be energized and in already steady-state are unknown before a measurement is made. When the du-
under a step input applied previously, at an unknown time, ration of the transient period is estimated, it is calculated in
when a new step input is applied to the system. For example, units of the largest time constant, τmax. Typically, the tran-
a DC electric motor may be running in steady-state under an sient period of a response is estimated as five time con-
applied voltage of 12 VDC, when a step change increases the stants, which leaves only 0.7 % of the initial range of the
voltage to 36 VDC. response remaining, Fig. 3.66.
We will use the force source-mass-damper system shown Since the mass-damper system is described as in steady-
in Figs. 3.13 and 3.48 as the example system. The system is state at time t = 0 under the input force of 100 N, we will
in steady-state under a step input of 100 N when we increase estimate minimum duration ∆t prior to t = 0 that the first
the applied force to 250 N, as shown in Fig. 3.65. We define step input was applied as ∆t = 5τ max , ta ≤ −5τ max. Having set
time, t = 0, for our convenience, as the time when the second a limit, or a constraint, on time ta we can now construct the
step input of force, ∆ F us (t ) = 150 N us (t ), is applied. Note input function by superposing (summing) two steps. Recall
162 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

∆v(0)
____ F(t), N F2(t) = 150 N us(t)
∆v(0) =1
F1(t) = 100 N us(t-t a ) Step Input 2
36.8%
Step Input 1
∆v(t)
_____ t = 0- t = 0+
13.5%
∆v(0) 5.0% 1.8% 0.7%
ta 0 t, sec

Fig. 3.67   Input function F( t), Eq. 3.64


0 τ max 2τmax 3τmax 4τmax 5τmax
time
1.0
Fig. 3.66   Normalized change remaining before steady-state. Time is
measured in terms of the largest time constant in the system 0.8

0.6
that the Heaviside unit step function transitions when its ar- v1g(t)
gument equals zero. Be careful with signs since ta is nega- 0.4
m
___
tive. The input force F( t) is the superposition of a 100 N step sec 0.2
which transitions at time t = ta and a 150 N step which transi-
tions at time t = 0, Eq. 3.63, 0

F (t ) = 100 N us (t − ta ) + 150 N us (t )
(3.63) -0.2
-1 -0.5 0 0.5 1
t, sec
We will use the parameter values b = 300 N · sec/m and
M = 40 kg and solve the system equation for the velocity of Fig. 3.68   Response function v1g( t), Eq. 3.65
the mass, Eq. 3.26, repeated below,

F0 us (t ) M dv1g with the same time shifting and scaling using in the input
(3.26) = + v1g
b b dt function, Eq. 3.64,

There is only one time constant in a first-order system. 100 N  − ( t + 0.91)



v1g (t ) =  1 − e 0.13
 us (t + 0.91sec )
Hence, N ⋅ sec 
300
m
M 40 kg
τ max = τ = → τ max = → τ max = 0.13sec 150 N  −t

N · sec
 us (t )
b + 1 − e 0.13
300 
N ⋅ sec 
m 300
m
We shall set the time of the previous step input to seven times
 m  − ( t + 0.91)

the time constant prior to t = 0, v1g (t ) = 0.33 1 − e  us (t + 0.91sec )
0.13

sec 
ta = −7τ max → ta = ( −7 )( 0.13) sec = −0.91sec
m  −t

 us (t )
+ 0.5 1 − e 0.13 (3.65)

sec 
F (t ) = 100N us (t + 0.91sec) + 150N us (t )
(3.64)

The input force F( t), Eq. 3.64, is plotted in Fig. 3.67. The response function v1g( t), Eq. 3.65, is plotted in Fig. 3.68.
We have already found the step response v1g( t), Eq. 3.54,
repeated below,
3.10.1 Example of Second-Order Pulse
F  − t
b
Response of an Energized System
v1g (t ) = 0 1 − e M 
(3.54)
b  
Formulate a transfer function for the force source-spring-
The response function v1g( t) is constructed by superposing mass-damper system equation for the velocity of the mass,
the step response function, Eq. 3.54, time shifted and scaled Eq. 3.34, reproduced below, and find the velocity of the
3.10  Initial Condition of Energized Systems 163

mass for the input force, F( t), shown in Fig. 3.67a, using Form the transfer function
the parameter values b = 500 N · sec/m, K = 5, 000 N/m, and
 d v1g 
2
M = 400 kg ,  1 dF (t )   b dv1g  K 
L  =L  2  +L   + L  v1g 
2  M dt   dt   M dt  M 
1 dF (t ) d v1g b dv1g K
(3.34) = + + v1g
M dt dt 2 M dt M
 d v1g  b
2
 dv1g  K
1
L  dF (t )  = L  2 + L   + L v1g { }
This system equation allows us to clarify a confusing aspect M  dt   dt  M  dt  M
of Laplace transformations for the solution of differential
equations. The two initial conditions required to solve a sec- 1 b K
sF ( s ) = s 2V1g ( s ) + sV1g ( s ) + V1g ( s )
ond-order differential equation are the values of the output M M M
variable and its first derivative. In Sect. 3.4.2, we established
that if the spring-mass-damper system was de-energized 1  b K
sF ( s ) =  s 2 + s +  V1g ( s )
prior to the application of a step input of force, F (t ) = F0 us (t ), M  M M
the initial conditions were the following:
1
V1g ( s ) s
( )=
dv1g 0 +
1 = M
( )
v1g 0+ = 0 and
dt M
F0 F (s)
s2 +
b
s+
K
M M
Notice that the initial condition of the first derivative of ve- 1
locity with respect to time is non-zero, though the system s
V1g ( s ) 400 kg
was de-energized at time t = 0 −. How is a non-zero initial =
F (s) N ⋅ sec N
condition to be incorporated into a transfer function method 500 5, 000
s2 + m s+ m
solution? Must we include the initial condition terms when 400 kg 400 kg
performing the Laplace transformation to create the trans-
fer function? No. If the system is initially de-energized, the V1g ( s ) 0.0025s
initial condition terms are neglected when using transfer = 2
F (s) s + 1.25s + 12.5
functions. The Laplace transform pair incorporates the con-
tribution of the non-zero initial condition. Again, the system
must be initially de-energized to neglect the initial condi- The denominator of the transfer function is the character-
tions when using transfer functions. Conveniently, superpo- istic function of the system. Set the characteristic function
sition presumes that the system is de-energized before the equal to zero to form the characteristic equation and solve
first input is applied. We will assume that the previous step it for the system’s eigenvalues. If the eigenvalues are real,
input of 25 N was applied at time ta = −6τ . then the unit step response is overdamped or non-oscillato-
The procedure is as follows: ry. If the eigenvalues are complex conjugates, then the unit
1. Construct the input function from superposed time shifted step response is underdamped or oscillatory:
and scaled Heaviside step inputs.
2. Form the transfer function from the system equation. −1.25 ± 1.252 − 4 · 12.5
s 2 + 1.25s + 12.5 = 0 → s1 , s2 =
3. Multiply the transfer function by the transform of a unit step, 2
to form the Laplace transform of the unit step response. s1 , s2 = −0.625 ± j 3.48.
4. Identify the correct Laplace transform pair. Perform the
algebra needed to create an exact match with the trans-
form pair to find the time-domain unit step response. The eigenvalues are complex conjugates. Therefore, the sys-
5. Construct the pulse response by time shifting and scaling tem is underdamped.
the unit step response with the same time shifts and scal- The largest time constant in the system is also the sys-
ing factors used to construct the input function. tem’s only time constant. It is the time constant of the decay
Construct the input function. Superpose scaled and time envelop of the decaying oscillation. We chose to apply the
shifted Heaviside step functions. The amplitude of each step preexisting step input six times constants before t = 0, at
input is the change in the input function at the time that step ta = −6τ .
function transitions,
1 1
F (t ) = 25 N us (t + 9.6) + 75 N us (t ) τ max = τ = = → τ = 1.6 sec → ta = −9.6
σ −0.625
− 150 N us (t − 5) + 100 N us (t − 10)
164 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.69  a Force input, F( t).


b Step inputs of force scaled a b 100 N u s (t-10)
in amplitude and shifted in time
which superpose to create 75 N us (t)
the force input, F( t) 100 100
F(t), N F(t), N
50 50
25 N us (t+6τ)

0 5 10 15 -6τ 0 5 10 15
t, sec t, sec
-50 -50

-100
-150 N us (t-5)

-150

The input function shown in Fig. 3.69b is Perform the algebra needed to create an exact match with the
form of the transform pair. Multiply the response signal by the
F (t ) = 25 N us (t + 9.6) + 75 N us (t ) unity ratio, 12.5/12.5. Factor in 12.5 and factor out 0.0025,
− 150 N us (t − 5) + 100 N us (t − 10)
12.5  0.0025 
V1g ( s ) =  2 
Multiply the transfer function by the transform of a unit u .s. 12.5 s + 1.25s + 12.5 
step. The result will be the Laplace transform of the unit step
response, 0.0025  12.5 
V1g ( s ) =  2 
1N u .s. 12.5 s + 1.25s + 12.5 
L {F u .s. (t )} =L {1N us (t )} → Fu .s. ( s ) = 1NL {us (t )} =
s
 12.5 
V1g ( s ) = 0.0002  2
V1g ( s )1 0.0025s  u .s.  s + 1.25s + 12.5 
Fu .s. ( s ) =  2 
F ( s ) s  s + 1.25s + 12.5 
Calculate the damping ratio ζ and ideal undamped natural
0.0025 s frequency ωn by equating the coefficients of like powers of
V1g ( s ) =
u .s. (
s s 2 + 1.25s + 12.5 ) the Laplace variable, s, in the characteristic functions of the
response signal and the Laplace transform pair,
0.0025
V1g ( s ) = s 2 + 2ζw n s + w n2 = s 2 + 1.25s + 12.5
u .s. s 2 + 1.25s + 12.5
2ζw n = 1.25 and w n2 = 12.5
Identify the correct Laplace transform pair to perform the
inverse transformation. The eigenvalues are complex con- rad 1.25 1.25
jugates, so the correct Laplace transform pair is written in w n = 12.5 = 3.54 and ζ = = = 0.177
sec 2w n 2 · 3.54
terms of the damping ratio, ζ, and the ideal undamped natural
frequency, ωn. The correct transform pair is the following:     12.5 
L −1
V 1g ( s ) = L
−1
0.0002  2
  
( )
wn  u .s.   s + 1.25 s + 12.5 
f (t ) = e −ζw n t sin w n 1 − ζ 2 t
1−ζ 2
 12.5 
v1g (t ) = 0.0002L −1  2 
w n2 u .s  s + 1.25s + 12.5 
F (s) = 2
s + 2ζw n s + w n2


v1g (t ) = 0.0002 
u .s.
3.54
 1 − 0.177 2
( 
e − (0.177)(3.54)t sin 3.54 1 − 0.177 2 t 

)
3.10  Initial Condition of Energized Systems 165

0.6 6

0.4
3
v1g(t) 0.2
mm v1g(t) 0
___
0 cm
___
sec
sec -3
-0.2
-0.4 -6
0 2 4 6 8 10
t, sec -9
-10 -5 0 5 10 15 20
Fig. 3.70   Unit step response of the mass velocity, v1g( t), to a step input t, sec
of 1 N
Fig. 3.71   Response of the mass velocity, v1g( t), in the spring-mass-
damper system to the force pulse input, F( t), as shown in Fig. 3.69

The unit step response, below, is plotted in Fig. 3.70,

v1g (t ) = 0.000719e −0.627 t sin(3.48t ) accelerate the mass, FM, and (ii) the force acting through the
u .s.
dashpot, Fb. The parameter values are b = 500 N · sec / m and
Construct the pulse response by time shifting and scaling M = 200 kg.
the unit step response function with the same time shifting
and scaling used to construct the input function, F( t), Energetic Equations

F (t ) = 25 N us (t + 9.6) + 75 N us (t ) Continuity or Node Eq : F (t ) − Fb − FM = 0


− 150 N us (t − 5) + 100 N us (t − 10)
Compatibility or Path Eq : v1g = v1g

The response function is given below and plotted in dv1g


Element Eqs : Fb = bv1g FM = M
Fig. 3.71, dt
1
v1g (t ) = 0.000719e −0.627 (t + 9.6) sin (3.48(t + 9.6)) 25N us (t + 9.6) Energy Eqs : E sys = E M EM =
2
Mv12g
+ 0.000719e −0.627 t sin (3.48t ) 75 us (t )
− 0.000719e −0.627(t − 5) sin (3.48(t − 5))150 us (t − 5) Reduction i: Input F( t), Output FM

+ 0.000719e −0.627(t −10) sin (3.48(t − 10))100 us (t − 10)


F (t ) − Fb − FM = 0 → F (t ) = Fb + FM

3.10.2  Initial Value Method F (t ) = bv1g + FM

Superposition is the most efficient method to find the re- Differentiate with respect to time to eliminate v1g:
sponse of a system energized prior to the application of
an input. However, it is possible to view the response to dF (t ) dv1g dFM dF (t ) b dF
=b + → = FM + M
a step change in the input as creating a new initial value dt dt dt dt M dt
problem. Thus, we may also create a piecewise continu-
ous response. This is the system equation. Reorder terms and clear the
We will illustrate the technique with the use of the force zero-order output variable term of its coefficient, in order to
source-mass-damping system of Fig. 3.2, but with the damp- express the system equation in time constant form:
ing represented as the alternative energetic schematic symbol
of a dashpot, as given in Fig. 3.72. The energetic equations M dF (t ) M dFM
= + FM
are reproduced below from Sect. 3.2.3.9 for reference. We b dt b dt
will choose as the output variables, (i), the force acting to
166 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.72  a Translational


mechanical mass-damper system a b 1
driven by a force source.
b Linear graph of the mass-
x,v
damper system driven by a
force source b F(t)
F(t) b M
M

Check the consistency of the units of the system equation, in F(t), N 600 N us(t)
terms of the power variables and time: F0+

F
Fb = bv1g → [b] =  v  t = 0- t = 0+
 
and 0 t, sec
dv1g Ft
FM = M
dt
→ [ M ] =  v  F0- = -400 N
  Previous
step input

 M dF (t )   M dFM 
 + [ FM ]
Fig. 3.73   Input force, F( t). The previous step input was applied at an
 = unknown time, ta. The system is in steady-state at time t = 0−
 b dt   b dt 

M F  M F 
 b t  =  b t  + [ F ] Check the consistency of the units in terms of the power vari-
ables and time:

 v F t F  v F t F F 
Fb = bv1g → [b ] =  
 =  + [F ] → [F ] = [F ] + [F ] v
F v t F v t
and
dv1g Ft 
The units are consistent. FM = M → [M ] =  
dt  v 
Reduction ii: Input F( t), Output Fb
 M dFb  M F 
F (t ) − Fb − FM = 0 → F (t ) = Fb + FM  F (t ) =   + [ Fb ] → F =  b t  + [ F ]
 b dt   
dv1g
F (t ) = Fb + M  F t v F
dt [F ] =   + [F ] → [F ] = [F ] + [F ]
 v F t
Use the element equation for the damper, Fb = bv1g , to elim-
inate v1g:
The units check.
d Fb M dFb We will now determine the responses of FM and Fb to
F (t ) = Fb + M → F (t ) = Fb +
dt b b dt the input function, F( t), shown in Fig. 3.73. The system is
in steady-state under a step input of force of −400 N, when
This is the system equation. Express it in time constant form the input force undergoes a step increase of 1,000 N at time
by reordering the output variable terms in decreasing order t = 0.
of differentiation: The initial value method views the previous input of
− 400 N as establishing the initial value of the output vari-
M dFb able at t = 0+ .
F (t ) =
(3.66) + Fb
b dt
3.10  Initial Condition of Energized Systems 167

Response FM(t) We begin with the response of the force, FM (t) Transient Steady-State
FM, which acts to accelerate the mass. In order to establish
the value of a power variable in a system at time, t = 0+ , at FM(0+) t = 0+
the instant after a step input is applied, we must first estab-
lish the value of the energy storage (state) variable of the
system at time, t = 0 − , at the instant before the step input is
applied. The state (energy storage) variable of this system is ~5τ time
the velocity of the mass, v1g, t = 0-

1 System in steady-state
E sys = E M and E M = Mv12g → v1g ≡ E storage variable
2 under previous step input

Because the value of an energy storage variable cannot Fig. 3.74   Response of the force, FM , acting to accelerate the mass
change instantaneously without infinite power, once we have in the force source-mass-damper system to the input force, plotted in
Fig. 3.73
established the state variable’s value at t = 0 − , we also know
its value at time t = 0+ ,
step input is applied to the system, as given in Fig. 3.74.
1 1
E M (0 −
) = E (0 )
M
+
→ ( )
Mv12g 0 − = Mv12g 0+ ( ) The new input initiates a new dynamic response, which will
2 2 reach a new steady-state in five time constants. Time t = 0+
is the beginning of the new transient period. All power vari-
( )
v1g 0 − = v1g 0+ ( ) ables in the system can change immediately in response to
the new input, except the state (energy storage) variables.
We then use the state variable’s and input variable’s values They remain unchanged because no time has elapsed be-
at t = 0+ in the energetic equations to algebraically deter- tween t = 0 − and t = 0+. Consequently, there has been no
mine the value of any power variable in the system at time time for an energy transfer into or out of the energy storage
t = 0+ . element. The transfer of a finite amount of energy requires
To determine FM (0+ ) , we must first determine the a finite amount of time.
value of the state variable, v1g, at time t = 0 − in response The state variable of this system is the velocity of the
to step input, F0 us (t + ta ) = −400 N. We do not know ta, but mass, v1g. We must establish the value of the state variable
we do know that ta occurred earlier enough, that the sys- at the instant before the step input is applied, v1g (0 − ). We
tem reached steady-state in response, by time t = 0 −. The know the input, F (0 − ) = −400 N, and we have established
steady-state response of all power variables in the system FM (0 − ) = 0. We use these values in the energetic equations
to a step input is to reach a constant value. If the state vari- to calculate v1g (0 − ),
able, v1g, has reached a constant value, then its derivative
equals zero. Hence, the force acting to accelerate the mass
in steady-state is zero,
F (t ) − Fb − FM = 0 → ( ) ( )
F 0 − = Fb 0 − + FM 0 − ( ) 0

( )
dv1g 0 − Use the element equation of the damper to eliminate Fb:
( )
v1g ( ss ) = v1g 0 − = constant → ( )
FM 0 − = M
dt 0
F (0 − )
F (0 − ) = bv1g (0 − ) → = v1g (0 − )
( )=0
FM 0 −
b

We arrive at the critical step in the process. It may appear −400 N m


= −0.80 = v1g (0 − )
that we have the initial condition we need, since FM is the N · sec sec
500
output variable, and we have established FM (0 − ). However, m
the force acting to accelerate a mass can change instanta-
neously, We can now establish the value of the force acting to acceler-
ate the mass at time t = 0+ , FM (0+ ) . We know the value of
FM 0 − ≠ FM 0+
(3.67) ( ) ( ) the input at time t = 0+ , F (0+ ) = 600 N. We know that the en-
ergy storage (state) variable cannot change from time t = 0 −
It is essential to recognize that a steady-state condition ex- to time t = 0+ ,
ists at time t = 0 −, but not at t = 0+ , the instant after the new m
v1g (0 − ) = −0.80 = v1g (0+ )
sec
168 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

1,000 Response Fb(t) The response of the force acting through


the damper to the input plotted in Fig. 3.71 follows the
800 same logic. The initial value method uses the same prem-
ise as the method of undetermined coefficients, which is
600 that the power variables in the system have the functional
FM (t), N
form of the input when the system has reached steady-
400
state. We found
200
m
v1g (0 − ) = −0.80 = v1g (0+ )
0
sec
0 0.5 1.0 1.5 2.0 2.5 3.0
t, sec Due to the simplicity of this system, the velocity drop
across the damper equals the velocity of the mass. Since
Fig. 3.75   Response of the force, FM (t ) , Eq. 3.68, acting to accelerate the force acting through the damper is proportional to the
the mass in the force source-mass-damper system to the input force,
plotted in Fig. 3.73 velocity difference across it, the force in the damper cannot
change instantaneously. However, in general, the force act-
ing through a damping element can change instantaneously.
We use the values of the input and the state variable at time In general,
t = 0+ in the energetic equations to determine FM (0+ ). Start-
ing with the continuity equation, ( )
Fb 0 − ≠ Fb 0+ ( )
F (t ) − Fb − FM = 0 → F 0+ = Fb 0+ + FM 0+( ) ( ) ( ) The velocity drop, in this case, across the damper and the ve-
locity of the mass are equal, so a conceptual error of equating
( ) ( )
F 0+ − Fb 0+ = FM 0+ ( ) ( )
→ F 0+ − bv1g 0+ = FM 0+ ( ) ( ) Fb (0 − ) and Fb (0+ ) would have had no consequence. Howev-
er, this is not true in general since a damper force can change
N · sec  m instantaneously:
600 N − 500
m
 −0.80  = FM 0
sec
+
( )
Pdamper ≡ Pb = Fb v1g
(3.69)
FM 0 ( ) = 1, 000 N
+

The rate of energy dissipation can change instantaneously,


The next step is to find the steady-state value of the out- because it is not an integral quantity. The accumulation of
put variable, FM( ss), and the time constant using the system energy requires time. Power is the flow rate of energy. The
equation in a time constant standard form: power flow into an energy storage element is the rate at which
that element accumulates energy. Power can change instanta-
dF0 us ( ss ) M dFM ( ss ) neously in our simple, abstract models.
= + FM ( ss ) → FM ( ss ) = 0 The beginning of the new transient response is the mo-
dt 0 b
 dt 0
τ ment when we need the initial value of the output vari-
able in order to solve the system equation. Knowing the
where input variable’s value at time t = 0+ and the output vari-
able’s value at time t = 0 − is insufficient, in the general
M 200 kg
τ= → τ= = 0.40 sec case, since all power variables in a system can have values
b N · sec at t = 0+ , which differ from those at time t = 0 − , except the
500
m energy storage variables. Therefore, we exploit the spe-
cial case of the steady-state of a step response, in which
Comparing the initial and final values of the output variable, all power variables have constant values, and all deriva-
FM (0+ ) = 0 < FM ( ss ), we see that the step response is a decay tives of power variables are zero. We calculate the value of
to zero from initial value of FM (0+ ) = 1, 000 N with the form the energy storage variable in that steady-state condition,
−t
FM (t ) = FM (0+ )e τ . which extends up to time t = 0 −. Knowing the input vari-
able’s and the energy storage variable’s values at any time,
The step response is FM(t), Eq. 3.68 below, Fig. 3.75,
t, is always sufficient to calculate values of every power
variable in a system.
−t
 FM (t ) = 1, 000 N e 0.40 (3.68)
3.11  Solved Problems 169

Knowing the values of the input and the state variable at 800
time t = 0+, we can now determine the initial value of the 600
output variable, Fb (0+ ):
400
Fb 0 ( ) = bv (0 )
+
1g
+

Fb(t), N
200
0
 N · sec   m
( )
Fb 0+ =  500
 m 
  −0.80 
sec  -200
-400
( )
Fb 0+ = −400N
-600
0 0.5 1.0 1.5 2.0 2.5 3.0
Having the initial value of the output variable, Fb (0 ), we + t, sec
use the system equation in time constant form to find the
Fig. 3.76   Response of the force, Fb, acting through the damper in
steady-state value, in response to the step input applied at the force source mass-damper system to the input force plotted in
time t = 0 and identify the time constant: Fig. 3.73

M dFb ( ss ) determine the response, rather than the initial value method.
F0 us ( ss ) = + Fb ( ss ) We will find the step responses of second-order systems, by
b
 dt 0
first calculating the eigenvalues to determine whether the
τ
system is overdamped and non-oscillatory, or underdamped
Fb ( ss ) = F0 = 600 N for t > 0 and oscillatory. Overdamped systems are easily solved with
either the method of undetermined coefficients or that of
where transfer functions. Underdamped systems are most easily
solved using transfer functions.
M 200 kg
τ= → τ= = 0.40 sec
b N · sec
500
m 3.11.1 Step Responses of Initially De-energized
System
Comparing the initial and final values of the output vari-
able we see Fb (0+ ) < Fb ( ss ). Hence, the step response is a The example system is the velocity source-spring-damper
growth from a non-zero initial value over the range system of Sect. 3.2.4.2. The energetic schematic and linear
graph of that system are shown in Fig. 3.77. The system’s
∆ Fb = Fb ( ss ) − Fb ( 0+ ) → ∆ Fb = 600 N − ( −400 N ) energetic equations and the system equations for the force
acting through the spring, FK, Eq. 3.31, and the velocity
∆Fb = 1, 000 N drop across the damper, v1g, Eq. 3.32, are reproduced below
for reference. The parameter values are K = 5, 000 N/m and
The step response Fb( t) is plotted in Fig. 3.76: b = 2,500 N · sec/m.

 −t
 Energetic Equations
Fb (t ) = ∆ Fb 1 − e τ  + Fb 0+
 
( )
Continuity or Node Eqs: Fsource = FK FK = Fb
 −t

Fb (t ) = 1, 000 N 1 − e 0.40
 − 400 N Compatibility or Path Eq: v1g ≡ v (t ) = v12 + v2 g

dFK
Element Eqs: Fb = bv2 g = Kv12
dt
F2
3.11  Solved Problems Energy Eqs: E sys = E K EK = K
2K
Rather than using the method of undetermined coefficients,
b dFK
we will find first-order system step responses by finding the bv (t ) = + FK (3.31)
initial and steady-state values of the output variable and time K dt
constant. We will then identify which is the correct step re-
 b dv (t ) b dv12
sponse function of the four possible options. If the system is = + v12 (3.32)
energized prior to the step input, we will use superposition to K dt K dt

170 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.77  a Energetic schematic


of a velocity source spring-damp-
a b K
ing system. b Linear graph of the 1 2
system shown schematically in
Fig. 3.77a x,v b
v(t) K
v(t) b

b dFK b dFK ( ss )
10 bv (t ) = + FK → bv0 us (t ) = + FK ( ss )
v(t) K dt 1 K dt 0
cm
___
sec 5 bv0 = FK ( ss )

0 The initial value of the output variable, FK (0+ ), is found


0 5 10 15
t, sec from the energetic equations. The system is described as de-
energized, prior to the application of the step input,
Fig. 3.78   Step input of velocity acting on the velocity source-spring-
damper system of Fig. 3.75. Note that input units are metric, not SI E sys ( 0− ) = 0 → E sys ( 0− ) = E K ( 0− ) = 0

Determine the spring force, FK, and the velocity difference The amount of energy stored in the system, and in any en-
across the spring, v12, when the de-energized system is acted ergy storage element within the system, cannot change in-
on by the step input velocity of 10 cm/sec, as shown in stantaneously, since it would require an infinite power flow
Fig. 3.78. to create a finite change over the infinitesimal duration from
Centimeter is not an SI unit. Express the step input in SI t = 0 − to t = 0+ . Hence,
by converting from centimeters to meters:

E K ( 0− ) = 0 = E K ( 0+ ) → E K ( 0+ ) =
( )→F
FK2 0+
v (t ) = v0 us (t ) → v (t ) = 0.1
m
us (t ) 2K
K (0 ) = 0
+

sec
We can summarize that systems which are de-energized,
Response Function, FK(t) Identifying the correct step re- prior to the application of an input at time t = 0 −, remain de-
sponse requires the establishment of initial and steady-state energized at t = 0+ , at the instant after the application of the
values, as well as the time constant, τ. The time constant input. Consequently, the state (energy storage) variables of
is found by unit analysis. The system equation is expressed those systems equal zero at time t = 0+.
in time constant form, when the zero-order output variable Comparing FK (0+ ) = 0 with bv0 = FK ( ss ), we identify the
term has no coefficient. The coefficient multiplying the first first-order step response as a stable exponential growth from
derivative of the output variable must have units of time, zero of the form, Eq. 3.51,

b dFK b  −t
  −t

bv (t ) = + FK → τ= FK (t ) = FK ( ss ) 1 − e τ  → FK (t ) = bv0 1 − e τ 
K dt
 K    
τ
 N · sec   m
2,500
N ⋅ sec FK (t ) =  2,500
 m 
 
sec 
(
0.1  1 − e − 2t )
τ= m = 0.5 sec
N
5, 000
m
(
FK (t ) = 250 N 1 − e − 2t )
The steady-state value of the output variable is found by im- Response Function, v12(t) The velocity difference across
posing the steady-state condition of a step response that all the spring, v12 (t ), is found using the same method. Find the
power variables in the system reach constant values, initial and final values of the output variable and the time
constant. Next identify which of the four possible first-order
3.11  Solved Problems 171

step responses is the correct response. There is only one time 300 0.1
constant in a first-order system. All system equations must FK(t)
0.08
have the same time constant, τ = K/b, in the example system.
200
Similarly, first-order systems have only one independent en- FK(t) 0.06 v12(t)
ergy storage element and, consequently, one state variable, v12(t) m
__
N 0.04 sec
which is the force acting through the spring, FK, in the ex- 100
ample system. If we were to find otherwise, then one or both 0.02
of the analyses would be erroneous.
The time constant, t, and the steady-state value of the out- 0 0
0 1 2 3 4
put variable are found from the system equation, t, sec
b dv (t ) b dv12 b
= + v12 → τ= = 0.5 sec Fig. 3.79   Step responses, FK( t) and v12(t)
K dt K dt
 K
τ The response functions FK( t) and v12( t) are plotted in
The time constant is the same as found in the previous analy- Fig. 3.79.
sis. We now use the steady-state condition of a step response
that all power variables in the system reach constant values
to determine the steady-state value of the output variable, 3.11.2 Step Responses of Initially Energized
System
b dv (t ) b dv12
= + v12
K dt K dt We will find the response of a force source-mass-damper
which is energized, when a step change in force is applied
b dv0 us (t ) 1 b dv12 ( ss ) to the system at t = 0. The key to using superposition to de-
= + v12 ( ss )
K dt K dt rive the response of a system is to formulate the response to
every input assuming that the system is de-energized when
b dv0 b dv12 ( ss ) that input is applied, even if it is clearly not. One assumes
= + v12 ( ss ) → 0 = v12 ( ss )
K dt 0 K dt 0 that the system is de-energized, because the effect of the non-
zero initial conditions will be accounted for, as the response
The initial value of the output variable is found from the en- of the prior input.
ergetic equations. The system is de-energized, prior to the We will rework the example of Sect. 3.10.2. The force
application of the step input in velocity. Consequently, the source-mass-damper system is shown in Fig. 3.80a, and
state (energy storage) variable equals zero, at the instant after the input is shown in Fig. 3.80b. The output variables are
the input is applied, FK (0+ ) = 0 . We always know the value (i) the force acting to accelerate the mass, FM, and (ii) the
of the input. In this case, v(0+ ) = 0.1m/ sec. Begin with the force acting through the dashpot, Fb. The energetic equa-
compatibility or path equation, as shown: tions and the system equations are reproduced below from
Sects. 3.2.3.9 and 3.10.2 for reference. The parameter values
( )
v (t ) = v12 + v2 g → v 0+ = v12 0+ + v2 g 0+( ) ( ) are b = 500 N·sec / m and M = 200 kg.

Eliminate the velocity difference across the damper by ex- Energetic Equations
pressing, v2g in terms of the damper force, Fb,
Continuity or Node Eq: F (t ) − Fb − FM = 0
( )
Fb 0 +
FK 0 ( ) +

( ) ( )
v 0+ = v12 0+ +
b
( )
→ v 0+ = v12 0+ + ( ) b
Compatibility or Path Eq: v1g = v1g
0
dv1g
Element Eqs: Fb = bv1g FM = M
m dt
( )
v 0+ = v12 0+ ( ) → 0.1
sec
= v12 0+( )
1
Energy Eqs: E sys = E M EM = Mv12g
2
m
Comparing v12 (0+ ) = 0.1 with v12 ( ss ) = 0, we see that the
sec M dF (t ) M dFM
response is a decay to zero of the form of Eq. 3.50, = + FM (3.26)
b dt b dt
−t 
m −2t
( )
v12 (t ) = v 0+ e τ → v12 (t ) = 0.1
sec
e
F (t ) =
M dFb
+ Fb (3.59)
 b dt
172 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. 3.80  a Energetic schematic


of a mass-damping system acted a b F(t), N 600 N us(t)
on by a force source. b The F0+
system was in steady-state, prior x,v
to step change of the input from
F (0− ) = −400 N to F (0+ ) = 600 N b t = 0- t = 0+
F(t)
M 0 t, sec
F0 - = -400 N
Previous
step input

Response Function, FM(t) We will first determine the unit 1


E sys ( 0− ) = E M ( 0− ) = 0 → E M ( 0 − ) = ( )
Mv12g 0 − = 0
step response for the output variable, FM, assuming that the 2
system is initially de-energized. Then we shall scale, time
shift, and sum these responses to correspond to the scaling The amount of energy stored cannot change instantaneously.
and time shifting of Heaviside step functions superposed to Therefore, it follows that
create the input function.
Determine the unit step response assuming that the sys- 1
E M (0− ) = E M (0+ ) = Mv12g (0+ ) = 0 → v1g (0+ ) = 0
tem is initially de-energized. Express the time constant and 2
the steady-state value of the step response from the system
equation in time constant form: Use the values of the unit input and the state variable at time
t = 0+ in the energetic equations to establish the value of the
M dF (t ) M dFM (t ) M output variable at time t = 0+ , F M (0+ ):
= + FM (t ) → τ=
b dt b
 dt b u .s.

τ
200 kg ( )
F − Fb − FM = 0 → FM 0+ = F 0+ − Fb 0+ ( ) ( )
τ= = 0.40 sec
N ⋅ sec
500
m
( ) ( )
FM 0+ = F 0+ − b v1g 0+ ( ) 0
→ ( )
FM 0+ = 1N
u .s.
Determine the steady-state unit step response value of the
output variable, FM. The steady-state response of power vari-
ables in any system to a step is to reach a constant value: Compare the initial and steady-state values of the output
variable to identify the type of step response:
M d M dFM (t )
1 N u s (t ) = + FM (t ) FM (0+ ) = 1N and FM ( ss ) = 0
b dt b dt

M d M dFM ( ss ) The unit step response is a decay to zero of the form


1 N us ( ss ) = + FM ( ss )
b dt 1 b dt −t  b

( ) ( )
− t
FM (t ) = FM 0+ e τ → FM (t ) = FM 0+ e M

u .s. u .s
M d 1N M dFM ( ss )
= + FM ( ss ) → FM ( ss ) = 0
b dt 0 b dt 0
FM (t ) = 1N e − 2.5 t
u .s

Use the algebraic equations of the equation list to determine Construct the input function, F( t), by superposing (sum-
the initial value of the output variable, FM (0+ ), assuming the ming) scaled and time shifted Heaviside step functions,
system is de-energized. The first step is always to establish F (t ) = F1 (t ) + F2 (t ), Fig. 3.81,
the value of the state (energy storage) variable at time t = 0+ ,
the instant after the step input is applied.  F (t ) = −400 Nus (t + ta ) + 1, 000 Nus (t ) where ta ≥ 5τ
If a system is de-energized, then all of its energy storage (3.70)
(state) variables equal zero. First-order systems have one in-
dependent energy storage element. The energy storage ele- Construct the response function, FM (t ) = FM1 (t ) + FM 2 (t ) ,
ment in this system is mass, by superposing (summing) scaled and time shifted unit step
3.11  Solved Problems 173

F(t), N F2(t) = 1,000 N us(t) 1,000


1,000 FM(t)
Step Input 2 800
600
FM(t), N 400
t = 0- t = 0+
ta Fb(t), N 200
0 t, sec 0
F1(t) = -400 N us(t+t a ) Fb(t)
-200
Step Input 1
-400
0 0.5 1.0 1.5 2.0 2.5 3.0
Fig. 3.81   Input step function, F (t ), created by superposing (summing)
scaled and time shifted heaviside step functions t, sec

Fig. 3.82   Response functions of the forces acting to accelerate the


functions, using the same scaling and time shifting used to mass, FM( t), and acting through the damper, Fb( t)
create the input function
variables equal zero. State variables cannot change instanta-
 b  b
−   ( t + ta )
M
−  t
M
neously, since that would require infinite power. Therefore,
FM (t ) = −400us (t + ta )1Ne + 1, 000 N us (t )e

−2.5(t + ta )
( )
v1g 0 − = 0 = v1g 0+ ( )
FM (t ) = −400N us (t + 2) e + 1, 000N us (t ) e −2.5t
We now use the values of the input and the state variable at
Response Function, Fb(t) We will again use superposition time t = 0+ to determine the initial value of the output variable,
to determine the response function, by applying the preexist- Fb (0+ ). As it happens, we only need the value of the state vari-
ing step input at a time ta such that the system has reached able, because the force acting through the damper is propor-
steady-state by time t = 0, and then superpose (sum) the re- tional to the velocity difference between nodes one and ground,
sponses due to the inputs applied at time t = ta and time t = 0
to construct the response function with the same time shifts ( )
Fb 0+ = b v1g 0+ ( ) → ( )
Fb 0+ = 0
and scaling as the input function. 0

Determine unit step response. The time constant and


the steady-state values of the output variable are determined Identify the correct first-order step response from the ini-
from the system equation. The time constant is shown here: tial and steady-state values of the output variable, Fb (0+ ) = 0
and Fb ( ss ) = 1 N . The unit step response is one of stable ex-
M dFb M
F (t ) = + Fb → τ= = 0.4 sec ponential growth,
b dt
 b
τ  −t
  −b
t
Fb (t ) = F0 1 − e τ  → Fb (t ) = 1N 1 − e M 
If we had found a different time constant than that from the   u .s.  
system equation for FM, one or both constants would be er-
roneous. The steady-state value of the output variable’s (
Fb (t ) = 1N 1 − e −2.5 t )
unit step response is found by the condition that it must be u .s.

constant,
Construct the response function scaling and time shifting
M dFb the unit step response function, by means of the same scaling
F (t ) = + Fb
b dt and time shifting used in the input function. We constructed
the input function, F( t), Eq. 3.70, for our calculation of the
M dFb ( ss )
1N us (t ) = + Fb ( ss ) → 1N = Fb ( ss ) response, FM. We employ Eq. 3.70, but remove the units of
1 b dt 0 newtons, since the unit step response function carries units,
yielding
To determine the value of the output variable, Fb, at the in-
 −b
( t + ta ) 
 400 us (t + ta )
stant after the unit step is applied to the system, we must Fb (t ) = −1N 1 − e M
first determine the value of the state variable at that instant. 
This is particularly easy, since we must assume that the sys-  −b
t
tem is de-energized to use superposition. If the system is + 1N 1 − e M  1, 000 us (t )
 
de-­energized, then we know that the state (energy storage)
174 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

x,v m
___ 0.8
v(t), sec
b 0.6
v(t) K 0.4
M
0.2

2 4 6 8 10 12 14
t, sec
Frictionless rollers
-0.4
Fig. 3.83   Energetic schematic of a velocity source-damper-mass-
spring system Fig. 3.84   The system was in steady-state under a step input of
− 0.25 m/sec, when the pulse input shown was applied at time, t = 0+

and, substituting for the time constant,


x,v
(
Fb (t ) = −1N 1 − e
−2.5 (t + ta )
) 400 u (t + 2)
s
b
( )
+ 1N 1 − e −2.5 t 1, 000 us (t ) g v(t) 1 2 K
M g
3.11.3 Second-Order Step Responses and Pulse
Responses
Frictionless rollers
We will use the velocity source-damper-mass-spring system
shown in Fig. 3.83 as the example system. The velocity source Fig. 3.85   Nodes of distinct values of the across variable velocity
shown on the energetic schematic of a velocity source-damper-mass-
is a machine or machine element which moves with a known
spring system
(and perhaps controlled) velocity v( t). A velocity source is a
power source. It is capable of exerting the force necessary to
impose the velocity, v( t), at its point of action. Using the linear respective nodes to represent energetic elements. Define
graph method, we will derive system equations for (i) the force the positive direction of the through variable force in each
acting to accelerate the mass and (ii) the force acting through branch with an arrowhead. Sources increase the across vari-
the spring. able in the positive direction of the through variable. In all
The response of a system to any arbitrary pulse, such as other elements, the across variable decreases in the positive
the input function, v( t), shown in Fig. 3.84, is constructed direction of the through variable.
using superposition. We will solve the system equations for Write one continuity equation less than there are nodes.
two sets of parameters, in order to illustrate solution meth- There are three nodes. The two ground nodes are of the same
ods for both the overdamped, non-oscillatory and the under- velocity and represent the same node. Write continuity equa-
damped, oscillatory cases. tions for nodes one and two. Omit the ground node. Notice
The key step in drawing a linear graph from an ener- that force applied by the velocity source equals force act-
getic schematic is to locate the nodes of distinct values of ing through the damper, Fb, whereas the force Fb divides be-
the across variable, Fig. 3.85. The across variable in a me- tween the force, FM, acting to accelerate the mass, and the
chanical system is velocity. The only energetic attribute of force, FK, acting through the spring.
an ideal mass is kinetic energy storage. It cannot deform, There can be no more useful compatibility equations than
or it would either store strain energy or dissipate mechani- there are holes in the linear graph. There are two holes. In
cal energy as heat through the shear of plastic deformation. this system there is only one useful compatibility equation.
Therefore, a mass is perfectly rigid with a single velocity. The second is either a restatement of the first or a trivial
The inertial reference velocity for the mass is ground. The equation.
velocity source must react against something to produce
the force needed to impose the prescribed velocity at its Energetic Equations
point of action. Since nothing is shown in the schematic,
one assumes that it reacts against ground. The velocity Continuity or Node Eqs: Fsource = Fb Fb = FM − FK
source, damper, and spring have nodes at either end, since
their ends can move independently. Compatibility or Path Eqs: v1g ≡ v (t ) = v12 + v2 g
Draw the linear graph by first drawing and labeling the
velocity nodes, Fig. 3.86. Then add branches between the v2 g = v2 g
3.11  Solved Problems 175

Fig. 3.86  a Nodes of distinct


values of the across variable
a b b
velocity. b Linear graph of the 1 2 1 2
velocity source damper-mass-
spring system

v(t) M K

g g

dv2 g
dFK
Element Eqs: Fb = bv12 FM = M = Kv2 g Check the consistency of the units in terms of the power vari-
dt dt
ables and time:
1 F2
Energy Eqs: E sys = E M + E K E M = Mv22g E K = K
2 2K F  dv F t
F = bv → [b ] =   F=M → [M ] =  
v dt  v 
Reduction i: Input v( t), Output FM
dF F
= Kv → [ K ] =  
1 dt t v 
v (t ) = v12 + v2 g → v (t ) = Fb + v2 g
b
dv (t ) 1 dFb dv2 g dv (t )1 d 1  d 2 v (t )   d 2 FM   b dFM   K 
= + → = ( FM + FK ) + FM b 2 
= 2 +
M dt  +  M FM 
dt b dt dt dt b dt M  dt   dt     

dv (t ) 1 dFK 1 dFM 1  v  F   b F   K 
= + + FM b t 2  =  t 2  +  M t  +  M F 
dt b dt b dt M
dv (t ) K 1 dFM 1
= v1g + + FM F v  F   F v F   v F 
dt b b dt M  2 
=  2+ + F
 v t  t   v F t t   F t t v 
d 2 v (t ) K dv1g 1 d 2 FM 1 dFM
2
= + + F  F  F  F 
dt b dt b dt 2 M dt  t 2  =  t 2  +  t 2  +  t 2 
d 2 v (t ) K 1 d 2 FM 1 dFM
= FM + +
dt 2 bM b dt 2
M dt The units are consistent.
Reduction ii: Input v( t), Output FK
This is the system equation, since the only power variables
1 1 dFK
which appear are the input and output variables. Rearrange v (t ) = v12 + v2 g → v (t ) = Fb +
the output variable terms in the decreasing order of differen- b K dt
tiation. Clear the coefficient from the second-order term to 1 1 dFK
express the system equation in standard form: v (t ) =
b
( FM + FK ) +
K dt
d 2 v (t ) 1 d 2 FM 1 dFM K 1 1 1 dFK
2
= + + FM v (t ) = FM + FK +
dt b dt 2 M dt bM b b K dt
d 2 v (t ) d 2 FM b dFM K M dv2 g 1 1 dFK
b = + + FM (3.71) v (t ) = + FK +
 dt 2
dt 2
M dt M b dt b K dt

M d 2 FK 1 1 dFK
v(t ) = 2
+ FK +
bK dt b K dt
176 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

0.6 duration before t = 0 of five times the maximum constant of


m
v(t), ___ the system, ta = −5τ max,
sec 0.4
0.2
t, sec m m m
v(t ) = −0.2 us (t + 5τ max ) + 0.4 us (t ) + 0.6 us (t − 2)
-5τmax 2 4 6 8 10 12 14 16
sec sec sec
m m
 − 1.2 us (t − 5) + 0.4 us (t − 10)
-0.4 sec sec
-0.6
(3.73)
-0.8
-1.0 We will now calculate the responses of the system to the
-1.2
pulse input, beginning with the overdamped case. The pulse
responses are constructed by superposing scaled and time
Fig. 3.87   The input function, v( t), is constructed of five scaled and
shifted unit step responses. We will determine the step re-
time shifted heaviside step functions sponses via transfer functions.

3.11.3.1 Pulse Response of the Overdamped


This is the system equation for the force acting through System
the spring. Express it in standard form by rearranging the The parameter values b = 600 N · sec / m, M = 200 kg, and
­derivatives in decreasing order. Remove the coefficient from K = 400 N / m yield an overdamped system.
the second-order output variable term: Pulse response of the force accelerating the mass, FM(t)
Create the transfer function. Perform the Laplace trans-
M d 2 FK 1 dFK 1 formation on the system equation, Eq. 3.71, neglecting the
v(t ) = + + FK
bK dt 2 K dt b initial condition terms:

bK
(3.72) d 2 FK b dFK K  d v (t ) 
2
d 2F b dFM K 
v(t ) = + + FK L b = L  2M + + FM 
M dt 2 M dt M  2 
 dt   dt M dt M 

Check the units of the system equation for consistency: b K


bs 2V ( s ) = s 2 FM ( s ) + sFM ( s ) + FM ( s )
M M
F dv Ft
F = bv → [b] =  v  F=M → [ M ] = 
  dt  v  
bs 2V ( s ) =  s 2 +
b K
s +  FM ( s )
 M M
dF F
dt
= Kv → [K ] = t v 
  Form the ratio of the output variable over the input variable:

 bK   d FK   b dFK   K
2
 Output ( s ) FM ( s ) bs 2
 M v ( t )  =  2 
+  +  M FK  Input ( s )
(3.74)
=
V (s)
=
b K
 dt   M dt    s2 + s+
M M
 bK  F   b F   K 
 M v =  2  +  + F
  t   M t   M  Calculate the eigenvalues to establish whether the system is
underdamped or overdamped. The denominator of the trans-
F F v  F   F v F   F v  fer function is the characteristic function. Set it equal to zero
 v =  2+ + F to form the characteristic equation:
 v t v F t  t   v F t t  t v F t 

F  F  F  F  b K
s2 + s+ =0
 t 2  =  t 2  +  t 2  +  t 2  M M
600 400
The units are consistent. s2 + s+ = 0 → s 2 + 3s + 2 = 0
Construct the input function by scaling and time shift- 200 200
ing Heaviside step function. The input, v( t), Fig. 3.84, is the
superposition of five step functions, Fig. 3.87. The system is −3 ± (3)2 − 4 · 2
s1 , s2 = → s1 = −1 and s2 = −2
described as in steady-state at time t = 0 in response to a step 2
input of −2.5 m/sec , applied previously. The time at which
this input was applied to the system is unknown, but the fact Real eigenvalues indicate the system is overdamped or non-
the system has reached steady-state provides the minimum oscillatory.
3.11  Solved Problems 177

The maximum time constant of the system equals the ab- 600
solute value of the smallest real component of the eigenval-
ues of the system,
400
1 1 FM(t), N
τ max = → τ max = = 1sec u.s.
σ min −1
200

Find the Laplace transformation of the input using the


Laplace transform pairs of Table 2.3. The input is a unit step, 0
v(t )=us (t ) . The relevant Laplace transform pair is
-100
-1 0 2 4 5
 us (t ) = 0 for t < 0 t, sec
 u (t ) = 1 for t > 0
s
 Fig. 3.88   The unit step response of the force acting to accelerate the
F s = 1
 ( ) s
mass, FM( t)

The Laplace transform of the unit step input is Factor the denominator of FM( s):
1
v (t ) = us (t ) → L {v (t )} = L {u (t )}
s → V (s) = s2 +
b
s+
K
= s 2 + 3s + 2 → ( s + a ) ( s + b ) = ( s + 1) ( s + 2)
s M M
Multiply the transfer function by the Laplace transfor-
mation of the input to yield the Laplace transformation of 600 s
FM ( s ) =
the output: u .s. ( s + 1)( s + 2)
FM ( s ) 1 bs 2    600 s 
V (s) = L −1

 FM ( s ) = L
−1

 
V (s) s b K
s2 + s+  u .s.   ( s + 1)( ) 
s + 2
M M
 s 
bs 2 FM (t ) = 600L
−1

FM ( s ) =  
 b K u .s  ( s + 1)( s + 2) 
s  s2 + s+ 
 M M 600N
FM (t ) =
u .s. b−a
(
be − bt − ae − at )
bs 600 s
FM ( s ) = → FM ( s ) = 2
b K s + 3s + 2 600 N
s2 +
M
s+
M
FM (t ) =
u .s. 2 −1
(
2e −2t − 1e −1t )
Perform the inverse Laplace transformation by using the The unit step response is below and plotted in Fig. 3.88:
Laplace transform pair to return to the time-domain. Exam-
ine Table 2.3 to seek a Laplace transform pair with the same FM (t ) = 1, 200Ne −2t − 600Ne − t
(3.75)
u .s.
form. The only signals with second-order denominators in
polynomial form correspond to oscillatory time-domain re- The unit step response of the force acting to accelerate the
sponses. The overdamped second-order pairs are written in mass, Eq. 3.75, undershoots its steady-state value of zero. Is
factored form. The relevant Laplace transform pair for the the system oscillatory, even though the eigenvalues are real?
inverse transformation unit step response is below, No, the system is not oscillatory. The power variable plot-
ted, FM, is the force acting to accelerate the mass. It reverses
 1
(
 f (t ) = b − a be − ae
− bt − at
) signs, because the initially de-energized (at rest) mass must
move to compress the spring and damper to force transfer
 s
F (s) = to it. What we see in the plot is the mass accelerating and
 ( )( s + b)
s + a decelerating, which is necessary, if the mass begins and ends
at rest. If the velocity of a mass changes sign during a step
Take care not to confuse the meaning of variables. The b in response, then the system is underdamped and oscillatory.
the Laplace transform table is the opposite of an eigenvalue. Use superposition to construct the pulse response func-
The b in FM( s), the Laplace transform of the response func- tion FM(t), Fig. 3.89. Scale and time shift the unit step re-
tion, is the damping coefficient of the system.
178 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

300 bK
Output ( s ) FK ( s ) M
200 (3.77) = =
Input ( s ) V (s) 2 b K
s + s+
100 M M
FM(t), N 0 Substituting in the parameter values yields, as below,
-100
600 · 400
-200 FK ( s ) 200 F (s) 1, 200
= → K = 2
V (s) 600 400 V ( s ) s + 3s + 2
-300 s2 + s+
-6 -3 0 3 6 9 12 15 200 200
t, sec
Fig. 3.89   The response of the force acting to accelerate the mass, Multiply the transfer function by unit step to yield the
FM( t), Eq. 3.76, to the pulse input, Eq. 3.73, shown in Fig. 3.84 Laplace transform of the unit step response:
1
sponse function, Eq. 3.75, using the same scaling and time v(t )=us (t ) → L {v(t )} = L {u (t )} s → V ( s )=
s
shift used to construct the input function, v(t), Eq. 3.73, to
create the pulse response function Eq. 3.76: FK ( s ) 1 1, 200 1, 200
V (s) = → FK ( s ) =
V ( s ) s s + 3s + 2
2
u .s.
2
s s + 3s + 2 ( )
m m m
v (t ) = −0.2 us (t + 5τ max ) + 0.4 us (t ) + 0.6 us (t − 2) Perform the inverse Laplace transformation using the La-
sec sec sec
m m place transform pair.
− 1.2 us (t − 5) + 0.4 us (t − 10) Factor the denominator to match the form of overdamped
sec sec
signals:

( )
FM (t ) = −0.2 1, 200Ne −2(t + 5) − 600Ne − (t + 5) us (t + 5)
FK (s) =
1, 200
→ FK (s) =
1, 200
(
+ 0.4 1, 200Ne −2t − 600Ne − t us (t ) ) u .s. ( 2
s s + 3s + 2 ) u .s. s ( s + 1)( s + 2)

( )
+ 0.6 1, 200Ne −2(t − 2) − 600Ne − (t − 2) us (t − 2) The relevant Laplace transform pair is
− 1.2(1, 200Ne ( −2 t − 5 )
− 600Ne ( ) )u (t − 5)
− t −5
 1  1 − bt 
s
 f (t ) = ab 1 + a − b be − ae 
− at
( )
+ 0.4(1, 200Ne ( − 600Ne ( ) )u (t − 10)
−2 t −10 ) − t −10   
s 
 F (s) = 1
(3.76)  s ( s + a )( s + b )

Pulse response of the force acting through the spring, FK(t)    1, 200 
L  FK ( s ) = L
−1 −1

 
Create the transfer function by performing the Laplace  u .s.   s ( s + 1)( s + 2) 
transformation on the system equation, Eq. 3.66, neglecting
the initial condition terms:  1 
FK (t ) = 1, 200L 
−1


u .s.  s ( s + 1)( s + 2) 
 bK   d 2 FK b dFK K 
L  v (t )  = L  2 + + FK  1, 200  1 
M   dt M dt M  FK (t ) =
u .s.
1+
1⋅ 2  1 − 2
(
2e −1t − 1e −2t 

)
L {v(t )} = L  d F2K  + b L  dFK  + K L { FK }
2
bK
M  dt  M  dt  M u .s.
( (
FK (t ) = 600 1 − 2e − t − e −2t ))
bK b K
V ( s ) = s 2 FK ( s )+ sFK ( s ) + FK ( s ) The unit step response of the spring force, FK( t), Eq. 3.78
M M M
below, is plotted in Fig. 3.90,
bK  b K
V ( s ) =  s 2 + s +  FK ( s )

FK (t ) = 600 1 − 2e − t + e −2t
(3.78) ( )
M M M u .s.

Note that the unit step response of the spring force, Fig. 3.90,
The transfer function is the ratio of the output variable over does not overshoot its steady-state value.
the input variable,
3.11  Solved Problems 179

600
600
400

400 200
FK (t), N FK(t), N
u.s.
0
200
-200
0
0 2 4 6 8 -400
t, sec -6 -3 0 3 6 9 12 15 18
t, sec
Fig. 3.90   The unit step response of the spring force, FK( t), Eq. 3.78 Fig. 3.91   The response of the spring force, FK( t), Eq. 3.79, to the pulse
input, Eq. 3.73, shown in Fig. 3.82

Use superposition to construct the pulse response


function. Use the same scaling for the time shifting, as was −2 ± ( 2)2 − 4 ·10
used to create the input function v( t), Eq. 3.73, Fig. 3.91: s1 , s2 = → s1 , s2 = −1 ± j 3
2
m m m The eigenvalues are complex conjugates, indicating that the
v (t ) = −0.2 us (t + 5τ max ) + 0.4 us (t ) + 0.6 us (t − 2)
sec sec sec system is underdamped and oscillatory. The frequency of os-
cillation is magnitude of the imagery component, 3 rad/sec.
m m
− 1.2 us (t − 5) + 0.4 us (t − 10) The real component is the coefficient of the exponent of the
sec sec decay envelope.

( (
FK (t ) = −0.2 600 1 − 2e − (t + 5) + e −2(t + 5) us (t + 5))) The time constant of the system equals the absolute value
of the real component of the eigenvalues of the system,
( (
+ 0.4 600 1 − 2e − t + e −2t us (t ))) τ=
1
→ τ=
1
= 1sec.
w −1
+ 0.6 ( 600 (1 − 2e ( − t − 2)
))
+ e −2(t − 2) us (t − 2) The unit step response is the product of the Laplace
transform of a unit step and the transfer function. The La-
− 1.2 ( 600 (1 − 2e ( − t −5 )
+ e ( ) )) u ( t − 5 )
−2 t − 5
s place transform of the unit step input is
+ 0.4 ( 600 (1 − 2e ( + e ( ) )) u (t − 10)
 − t −10 ) −2 t −10
(3.79) 1
s
v(t )=us (t ) → L {v(t )} = L {u (t )} s → V ( s )=
s
3.11.3.2 Pulse Response of the Underdamped Output ( s )
Input ( s ) = Output ( s )
System Input ( s )
The parameter values b = 400 N · sec/m , M = 200 kg, and
K = 2, 000 N/m yield an underdamped system. FM ( s ) 1 bs 2
V (s) =
Pulse response of the force accelerating the mass, FM(t) V (s) s s2 + b s + K
The transfer function for the force acting to accelerate M M
the mass, FM( t), is Eq. 3.74, reproduced below,
bs 400 s
FM ( s ) = → FM ( s ) = 2
Output ( s ) FM ( s ) bs 2 b K s + 2 s + 10
(3.74) = = s2 + s+
Input ( s ) V (s) b K M M
s2 + s+
M M The relevant Laplace transform pair is written in terms of the
damping ratio, ζ, and the ideal undamped natural frequency,
Calculate the eigenvalues. The denominator of the transfer ωn, because the response is underdamped and oscillatory,
function is the characteristic function,

b K

 f (t ) = −
1
1
− ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t − φ )
s2 + s+ =0 
M M   1−ζ 2 
  wd 
 where φ = tan −1
 σ  = tan −1
 
400 2, 000     ζ 
s2 + s+ =0 → s 2 + 2 s + 10 = 0 
200 200 s
F (s) =
 s 2 + 2ζw n s + w n2

180 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

300 300

200 200

100
100
FM(t), N FM(t), N 0
0
-100
-100 -200

-200 -300
-6 -3 0 3 6 9 12 15
0 1 2 3 4 5 6 7 t, sec
t, sec
Fig. 3.93   The pulse response of the force acting to accelerate the mass,
Fig. 3.92   The unit step response of the force acting to accelerate the FM( t), Eq. 3.81
mass, FM( t)

   400 s   1 − ( 0.316)2 
L  F M ( s ) = L
−1 −1

 2   1−ζ 2 
 u . s .   s + 2 s + 10  φ = tan −1
 = tan 
−1
 = − 0.69 rad
 ζ   0.316 
 s 
F M (t ) = 400L
−1

 2 
u .s.  s + 2 s + 10  The unit step response of FM( t), Eq. 3.80, is plotted in Fig. 3.92,

The damping ratio, ζ, and the ideal natural frequency, ωn, are F M (t ) = −422e − t sin (3t − 0.69 rad )
(3.80)
calculated by equating the coefficients of like powers of s in u .s.

the denominator,
The pulse response, FM (t), Eq. 3.81, Fig. 3.93, is created by
400 s 400 s scaling, time shifting, and superposing the unit step response
=
s 2 + 2 s + 10 s 2 + 2ζw n s + w n2 with the same scaling factors and time shifts as the input
function v( t), Eq. 3.73, below,
s1 10 = w n2
s 2 + 2 s + 10 = 2ζw n s + w n2 →
s 0 2 = 2ζw n m m m
v (t ) = −0.2 us (t + 5τ max ) + 0.4 us (t ) + 0.6 us (t − 2)
sec sec sec
rad
w n = 10 = 3.16 → 2 = 2ζw n m m
sec − 1.2 us (t − 5) + 0.4 us (t − 10)
sec sec
1 1
ζ= = = 0.316
wn 3.16
(
FM (t ) = −0.2 −422e − (t + 5) sin (3 (t + 5) − 0.69 rad ) us (t + 5) )
The unit step response of the force acting to accelerate the (
+ 0.4 −422e sin (3t − 0.69 rad ) us (t )
−t
)
mass, FM(t), is
(
+ 0.6 −422e − (t − 2 )
)
sin (3 (t − 2) − 0.69 rad ) us (t − 2)
  − 1.2( −422e ( sin (3 (t − 5) − 0.69 rad ))u (t − 5)
( )
− t − 5)
1
FM (t ) = 400  − e −ζw n t sin w n 1 − ζ 2 t − φ  s

1−ζ + 0.4( −422e ( sin (3 (t − 10) − 0.69 rad ))u (t − 10)


2
u .s.   − t −10 )
s

 1−ζ 2  (3.81)
where φ = tan −1  
 ζ  Pulse response of the force through the spring, FK(t)
The transfer function for the force acting to accelerate
400 400 the mass, FK( t), is Eq. 3.77, from above,
− =− = − 422
1−ζ 1 − ( 0.316)
2 2

bK
−ζw n t − (0.316)(3.16)t −t Output ( s ) FK ( s ) M
e =e =e (3.77) = =
Input ( s ) V (s) 2 b K
s + s+
w n 1 − ζ 2 t = 3.16 1 − ( 0.316) t = 3t
2 M M
3.11  Solved Problems 181

Evaluate the transfer function using the parameter values for 600
an underdamped system, b = 400 N · sec/m , M = 200 kg, and
500
K = 2, 000 N/m:
400
400 · 2, 000 FK (t), N
FK ( s ) F (s) 4, 000 u.s. 300
= 200 → K =
V ( s) 2 400 2, 000 V ( s ) s 2 +2 s + 10 200
s + s+
200 200
100

Multiply the transfer function by the Laplace transform of a unit 0


step to yield the Laplace transform of the unit step response: 0 1 2 3 4 5 6 7
t, sec
FK ( s ) 1 4, 000 4, 000
V (s) = → FK ( s ) = Fig. 3.94   The unit step of the spring force, FK( t), Eq. 3.82
V ( s ) s s 2 +2s + 10 s ( s 2 +2s + 10)

The relevant Laplace transform pair from Table 2.3 is the


following: 400


 f (t ) = 1 −
 1
1
− ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t + φ ) 200

 FK(t), N 0
  wd   1−ζ 2 
 where φ = tan −1
 σ  = tan −1
 
    ζ  -200

 F (s) = w n2
-400

 (
s s 2 + 2ζw n s + w n2 ) -6 -3 0 3
t, sec
6 9 12 15

There must be an exact match with the Laplace-domain Fig. 3.95   The response of the spring force, FK( t), to the input function,
function of the transform pair. Multiply the response func- v( t), Eq. 3.66
tion by the unity ratio, 10/10:
10 4,000 4,000 10
FK ( s ) = → FK ( s ) = The unit step response of the spring force FK( t) is Eq. 3.82,
u .s. 10 s ( s 2 +2 s + 10) u .s. 10 s ( s 2 +2 s + 10)
below, and is plotted in Fig. 3.94,
10
FK ( s ) = 400 2 FK (t ) = 400 − 422e − t sin (3t + 1.25 rad )
(3.82)
u .s. s ( s +2 s + 10) u .s.

The damping ratio and the ideal, undamped natural frequen- The unit step response is scaled, time shifted, and super-
cy for this system and parameters, as calculated above, are posed to create the pulse response, FK(t), Fig. 3.95, for the
input function, v( t), Eq. 3.73,
rad
ζ = 0.316 and w n = 3.16
sec m m m
v (t ) = −0.2 us (t + 5τ max ) + 0.4 us (t ) + 0.6 us (t − 2)
sec sec sec
Perform the inverse Laplace transformation using the trans- m m
− 1.2 us (t − 5) + 0.4 us (t − 10)
form pair: sec sec

 
FK (t ) = 400 1 −
u .s. 
1
1 − ( 0.316)
2 (
e − (0.316)(3.16)t sin 3.16 1 − ( 0.316) t + 1.25 rad
2
)


182 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

( ) ( )
FK (t ) = −0.2 400 − 422 e − (t + 5) sin (3 (t + 5) + 1.25 rad ) us (t + 5) + 0.4 400 − 422 e − t sin (3 t + 1.25 rad ) us (t )

( ) ( )
+ 0.6 400 − 422 e − (t − 2) sin (3 (t − 2) + 1.25 rad ) us (t − 2) − 1.2 400 − 422 e − (t − 5) sin (3 (t − 5) + 1.25 rad ) us (t − 5)

( )
+ 0.4 −422 e − (t −10) sin (3 (t − 10) − 0.69 rad ) us (t − 10)

Summary of which elements in the system store energy, and how the
energy stored in an element is calculated.
Linear graphs are circuit-like representations of energetic The set of energetic equations is reduced using elimina-
systems. Draw a linear graph by drawing and identifying tion by substitution to derive the differential system equa-
the velocity nodes. Then add the elements between the tion, which relates the input variable to an output variable
nodes. The crucial step is to identify all nodes of the across of interest.
variable on the schematic of the system. The symbol for a Solution of differential system equations. We have two
source is a circle with the source variable in it. The other complementary ways of solving the system equation for an
branches are lines with an arrowhead, indicating the posi- input function constructed by superposing scaled and time
tive direction of the through variable (in this case, force) shifted Heaviside step functions. The method of superpos-
and the positive direction of the drop, or decrease in the ing scaled and time shifted unit step responses is much more
across variable, velocity. The lower half of the branch M is efficient than the initial condition method. The initial condi-
a dashed line to indicate that no reaction force acts on the tion method requires a new solution for each segment of the
mass from the ground. input pulse, whereas superposition method does not.
Write all equations which can be stated for the energetic Initial condition method. View the previous inputs as es-
model, represented by the linear graph before beginning any tablishing the initial value of the output variable for the next
algebra. Organize the equations for easy reference. The equa- input. The response of the system to the input establishes
tions which describe an energetic system are categorized the value of the output variable at time t = 0 −. Knowing the
as continuity (node), compatibility (loop or path), element values of the input and output variables at time t = 0 − , and
(constitutive), and energy. They will be used to derive the also that the steady-state values of all variables in the system
differential system equation and then to establish the initial must be constants, provide sufficient information to estab-
conditions needed to solve the system equation. lish the value of the energy storage variable at time t = 0 − .
The compatibility and continuity equations depend solely The Heaviside step input transitions instantaneously from
on how elements are connected to form a network. All ener- zero to one when its argument becomes positive. Although
getic systems have a variable which sums at a point or node. we model the input’s power variable as capable of an instan-
In mechanical systems, the through variables, forces and taneous change, an energy storage (state) variable’s value
torques, sum at a node to yield equilibrium equations. We cannot change instantaneously. Power must flow into or out
write one node equation less than there are nodes in the sys- of an energy storage element for the state variable’s value
tem. It is best to omit the ground node from the node equa- to change. Flow takes time. Only infinite flow rates would
tions. Similarly, all energetic systems have a variable which allow finite changes over zero time. Infinite power flows are
changes between nodes. In mechanical systems, the across impossible. Therefore, the energy (state) variable’s value at
variable is velocity. The differences between the values at times t = 0 − and t = 0+ are equal. Important. The output vari-
nodes must be consistent and yield the same sum along two able of the system equation may, or may not, be the energy
different paths between node A and node B, or zero when storage (state) variable. In any case, the only variables whose
summed around a loop back to the starting node. We can values are known at time t = 0+ , at the instant after a step
write no more, and often fewer, independent loop or path input has transitioned, are the input variable and the state
equations than there are “holes” in the linear graph. variables. If the output variable is not the state variable, then
The element or constitutive equations are the model of a it must be determined by using the input and state variables
single energetic ­attribute. The element equations, which re- values at t = 0+ in the energetic equations.
late the power variables of individual elements and energy Superposition method. Sum the responses due to the
equations, which state the amount of energy stored in the previous input and the input applied at time t = 0. First,
system, must be written with the nodes and assumed posi- create the input function as the sum of step functions, except
tive direction of the through variables defined on the linear that the step input applied at an unknown negative time must
graph. We must use linear element equations to yield linear be represented as a constant, since we do not know when its
differential equations. The energy equations are statements transition occurred, other than it was long enough ago, that
Problems 183

the system’s response to it reached steady-state. We solve system’s responses to all inputs received, we must assume
for the response of a unit step of the input variable, as if that the system to be de-energized for each input, or we will dou-
input acts on a de-energized system, even when is clearly not ble-count the effect of an input.
true. The response function for an arbitrary input pulse is
created by the following procedure:
1. Scale the unit step response by the magnitude of the step Problems
inputs, which are superposed to create the input pulse;
2. Time shift the scaled unit step responses, by the same Reminders
time shifts used to create the pulse input; 1. Write energetic equations with proper notation. The
3. Multiply each term of the response function by a Heavi- problem statements do not explicitly state that the prob-
side unit step function with the same time shift, in order to lems require (a) the energetic equations and (b) proper
zero-out that term, until its corresponding input step has notation, because those are part of the linear graph
acted on the system; and method and implied by drawing a linear graph. The
4. Sum (superpose) the response functions which are scaled energetic equations consist of compatibility, continuity,
and shifted in time as the individual input step function, element, and energy equations. Proper notation refers to
which creates that response function. using (a) node subscripts to identify the positive direc-
When the system is de-energized, before a step input is ap- tion of the drop in the across variable of an element, and
plied to the system, we know that the values of the energy (b) the element parameter as the subscript, which identi-
storage (state) variables at time t = 0 − are zero. We also fies the through variable of an element.
know that state variables cannot change instantaneously. 2. Clear fractions and create common denominators as
Therefore, the values of the state variable remain zero, at the you work the reduction. Improper fractions must be
instant immediately after the Heaviside step input is applied, cleared to present the result in standard form. It is gener-
at time t = 0+ . The initial value of the output variable, t = 0+, ally easier to place a sum of ratios over a common denom-
is determined by using the values of the input variable and inator when it is created, rather than carrying improper ra-
the energy storage (state) variable at time t = 0+ in the ener- tios forward and clearing them at the end of the reduction.
getic equations. One advantage of placing sums of ratios over a common
If the system is energized and running in steady-state at denominator is that the resulting ratio can be cleared from
time t = 0 −, under a step input applied at an unknown nega- a product by multiplying its fractional inverse.
tive time, ta, the response of the system to additional step in- 3. Standard form. (a) Time constant form applies only to
puts can be solved by superposition, by including the steady- the first-order system equations. The terms are ordered,
state response (the particular solution) of the previous input with the derivative first, followed by the zero-order
to the response function. In general, knowledge of the values term. (b) Standard form for higher order differential
of the input and output variables is not sufficient to calculate equations requires the coefficient of the highest-order
all remaining power variables in a system. However, in the derivative of the output variable to be cleared. This is
specific case in which a system was subjected to a prior step the same standard form used for polynomials. (c) The
input and has reached steady-state, all power variables in the system equation cannot have improper ratios. Improper
system will have reached constant values, since the input is ratios are not standard form, because they promote error.
constant in steady-state. Consequently, all derivatives with 4. Unit checks. Check units of the system equation before
respect to time equal zero. This additional knowledge allows solving it, by expressing element parameters in terms
us to determine the value of the non-state variable of each of the system’s power variables and time. Do not check
energy storage element, giving us enough information to es- units in terms of fundamental units. Although fundamen-
tablish the values of the state variables at time t = 0 −. tal units are straightforward in a system of single energy
The remaining possibility is that the system is in a tran- type, they become very cumbersome in “hybrid” sys-
sient state at time t = 0 − , due to a prior step input. This case tems of more than one energy type, for example, electric
can only be solved by using superposition, if the time at motors. For example, how would you combine mechani-
which the prior step input was applied is known. If it is un- cal units of torque with electrical units of voltage?
known, then the response must be determined with the initial 5. Convert units to SI for calculations. Convert result to
value method. US customary units, if required. Remember that met-
Again, superposition only works with linear systems. If ric units may not be SI, for example, centimeter. The SI
a system can be reasonably approximated (modeled) using unit of length is meter. Likewise, remove scaling pre-
linear elemental equations, then each input to the system cre- fixes and express values in the base unit, for example,
ates an output of the same form, but its amplitude is scaled kN = 1, 000 N .
by the amplitude of the input step function. When a response 6. Causality. If the system is de-energized for time, t < 0,
function is formulated as the sum (or superposition) of the then the response of the system for time t < 0 is zero.
184 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. P3.1   Input pulses


a b
10
20
5
F(t), N
v(t)
10 m 0 0
___ 2 4
sec t, sec
0 -5
0 1 2
t, sec
-10

c d
20

200 15
10
100 F(t)
F(t) 5
kN
N 0 0
0 5 10 15 0 10 20 30
t, sec -5 t, sec
-100
-10

However, evaluating response functions for negative axes. Select “Traces” in the Format menu to change the
time will yield a non-zero result, if the response func- thickness, color, and type of line.
tion is not multiplied by a Heaviside step function time 10. MATLAB plots. Show the entire transient period of
shifted to when the corresponding input is applied to the response but not too much of steady-state. Edit the
the system. Do not plot negative time unless you use a limits of the ordinate to maximize the proportion of the
Heaviside unit step function to zero-out each response, plot occupied by the trace. Add gridlines and change the
until its corresponding input is applied to the system, trace thickness.
and the response has the correct time shift.
7. Inspect the plots of your results. Look at your results, Problem 3.1 For the pulses shown in Fig. P3.1:
and check that they are reasonable and complete. Is the 3.1.a Sketch the time shifted and scaled Heaviside unit step
response a reasonable shape? Does the plot show a re- functions which superpose to form the pulse.
sponse prior to the application of the input? Is the verti- 3.1.b Express the pulse as a function consisting of time
cal extent of the trace shown, or must the axis limits be shifted and scaled Heaviside unit step functions.
edited? Are the limits of the vertical axis too far from 3.1.c Using Mathcad or MATLAB.
the extent of the response, thereby squashing the trace? i Plot the time shifted and scaled Heaviside step
Does the plot show the beginning of steady-state but not functions sketched in part a on the same plot.
excessively so? ii Plot the pulse function derived in part b.
8. Title plots and label axes. Unidentified plots and unla-
beled axes are unacceptable. Problem 3.2 A translational mechanical system consisting of
9. Mathcad plots. When an x–y plot is inserted in a a force source, a mass, M = 5 kg supported on ideal friction-
Mathcad worksheet, Mathcad automatically sets the less rollers, and damper, b = 3 N · sec/m , is shown schemati-
ranges of both axes. Mathcad always sets the limits of cally in Fig. P3.2a and b.
the abscissa (the independent variable) from − 10 to + 10. 3.2.a Use the information presented in the schematics and
Click on a limit to edit it. Set the lower limit of time to linear graph shown in Fig. P3.2 to derive a complete
zero, and set the upper limit, such that the response has set of energetic equations for the system.
reached steady-state. Show the entire transient period of 3.2.b Derive the system equation that relates the force input
the response but not too much of steady-state. Edit the to these variables:
limits of the ordinate to maximize the proportion of the i The velocity of the mass.
plot occupied by the trace. Right click on the plot to ii The force acting to accelerate the mass.
bring up the “Format” menu. Add gridlines to the two iii The force acting through the damper.
Problems 185

Fig. P3.2  a System with a


translational force source, mass,
a x,v
b x,v
and damper system. b Schematic
annotated with nodes of distinct b b
values of velocity. c Linear graph F(t) F(t) 1
of the system. d Force input M M g
pulse which acts on the system g

c 1 d

20

F(t) F(t), N
M b
10

0
0 1 2 3
g t, sec

Fig. P3.3  a System with a trans-


lational spring, damper system
a b
acted on by a velocity source. x,v x,v
b Schematic annotated with
b b
v(t) K v(t) K
nodes of distinct velocity.
g
c Linear graph of system. g 1 2
d Velocity input pulse which
acts on the system

c d 10
K
1 2
5
v(t)
m 0 0
___ 2 4 t, sec
v(t) b sec
-5

-10
g

Check the units of the system equation in terms of power 3.3.a Use the information presented in the schematics and
variables and time. linear graph shown in Fig. P3.3 to derive a complete
3.2.c The system is at rest at time, t < 0. Determine the unit set of energetic equations for the system.
step responses of the system equations derived in Part 3.3.b Derive the system equation that relates the applied
b. Use superposition to determine the responses to the velocity input to
input force pulse, F( t), plotted in Fig. P3.2d. i The force in the spring.
3.2.d Plot the responses, using Mathcad or MATLAB. ii The velocity drop across the spring.
Check the units of the system equations, in terms of power
Problem 3.3 A translational mechanical system consist- variables and time.
ing of a velocity source, a spring with spring constant, 3.3.c The system is relaxed at time, t < 0. Determine the
K = 4 N / m , and a dashpot (damper) with damping coeffi- unit step responses of the system equations derived in
cient, b = 2 N · sec / m , is shown in the schematics, as given part b. Use superposition to determine the responses to
in Fig. P3.3a and b. the applied velocity pulse, v( t), plotted in Fig. P3.3d.
3.3.d Plot the responses, using Mathcad or MATLAB.
186 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. P3.4  a System with a


translational spring, damper
a b
x,v x,v
system acted on by a velocity
source. b Schematic annotated
b b
v(t) K v(t) K
with nodes of distinct velocity.
g
c Linear graph of system. g 1 2
d Velocity input applied to
the system

c d
K cm
v(t), ___
1 2 sec
10 = v0-

v(t) b 0 t, sec

-10 = v0+
g

Problem 3.4 A translational mechanical system consist- 3.5.b The input velocity, v( t), plotted in Fig. P3.5d is
ing of a velocity source, a spring with spring constant, applied to the damper at time t = 0. Determine the unit
K = 2 N/m, and a dashpot (damper) with damping coeffi- step response of the system equations, derived in part
cient, b = 4 N · sec/m, is shown in the schematics, as given in a. Use superposition to solve the system equations for
Fig. P3.3a and b. The velocity, 10 cm/sec, was applied at an the velocity input, as shown in Fig. P3.5d.
unknown time t < 0. The system reached steady-state by time 3.5.c Plot the responses using Mathcad or MATLAB.
t = 0 − . At time t = 0 , the velocity applied to the system is re-
versed, v( t) = −10 cm/sec, for t > 0, as shown in Fig. P3.3d. Problem 3.6 A translational mechanical system consisting of
3.4.a Use the information, as presented in the schematics a force source, a spring with K = 100 N / m, a damper with
and linear graph shown in Fig. P3.4, to derive a com- b = 300 N · sec / m, and a massless, rigid bar constrained to
plete set of energetic equations for the system. translation is shown in Fig. P3.6a. The system has reached
3.4.b Derive the system equation that relates the applied steady-state under the application of an input of − 40  N,
velocity input to ­applied at an unknown time, t < 0. At time t = 0, the input
i The force in the spring. force is increased to + 60 N, as shown in Fig. P3.6d.
ii The velocity drop across the spring. 3.6.a Derive the system equation which relates the input
Check the units of the system equations, in terms of power force F( t) to
variables and time. i The force acting through the spring, FK.
3.4.c Use the initial condition method to determine the ii The force acting through the damper, Fb.
response of the system’s power variables from Part b to iii The velocity of the point of application of the
the velocity applied at time t = 0, shown in Fig. P3.4d. input force, F( t).
3.4.d Plot the responses using Mathcad or MATLAB. Check the units of the system equations in terms of power
variables and time.
Problem 3.5 A translational mechanical system which con- 3.6.b Use the method of undetermined coefficients to solve
sists of a velocity source, a mass, and a damper is shown in the system equations of part a, for the input shown in
Fig. P3.5. Its mass is M = 5 kg, supported on ideal, frictionless Fig. P3.6d.
rollers. The damping constant is b = 7 N · sec / m. The system 3.6.c Plot the response using Mathcad or MATLAB.
was at rest before the input shown in Fig. P3.5d was applied.
3.5.a Derive the system equation that relates the applied Problem 3.7 A translational mechanical system which con-
velocity input to sists of a force source, a mass, a spring, and a damper is
i The force acting to accelerate the mass. shown in Fig. P3.7. The system was de-energized, before
ii The velocity of the mass. the input shown in Fig. P3.7d was applied. The param-
iii The velocity drop across the damper. eter values are Case I: b = 800 N · sec/m , M = 250 kg, and
Check the units of the system equations in terms of power K = 400 N/m and Case II: b = 400 N · sec/m, M = 250 kg,
variables and time. and K = 800 N/m .
Problems 187

Fig. P3.5  a System with a


velocity source, damper, and a b
2
mass. b Schematic annotated b b
with nodes of distinct values of v(t) g v(t) 1
velocity. c Linear graph of the M M
system. d Velocity input applied
to the system

c b d
1 2 10

v(t)
0
v(t) m
___ 0 2 4 t, sec
M sec
-10

Fig. P3.6  a Translational


mechanical system. b Schematic
a x,v b x,v
annotated with nodes of distinct
values of velocity. c Linear K K
graph of the system. d Force
input applied to the system
F(t) F(t)
g
b g b

c 1 d F(t), N
F(0+) = 60 N

F(t) b K
0 t, sec
F(0 -) = -40 N

3.7.a Derive the system equation that relates the applied response of the system equations derived in part a.
velocity input to Use superposition to solve the system equations for
i The force acting to accelerate the mass. the velocity input, shown in Fig. P3.7d.
ii The velocity of the mass. 3.7.c Plot the responses using Mathcad or MATLAB.
iii The force acting through the spring.
Check the units of the system equations in terms of power Problem 3.8 A translational mechanical system which con-
variables and time. sists of a force source, a mass, a spring, and a damper is
3.7.b The input force, F( t), plotted in Fig. P3.7d is applied shown in Fig. P3.8. The system was in steady-state under
to the damper at time t = 0 . Determine the unit step a previous step input, before the force pulse shown in
188 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Fig. P3.7  a System with a


force source, mass, spring, and
a x,v b x,v
damper. b Schematic annotated b b
with nodes of distinct values of 1 g
velocity. c Linear graph of the
system. d Force input applied to
F(t) F(t)
M M
the system g g
K K

Frictionless rollers Frictionless rollers

c 1 d
4
F(t)
F(t) kN 2
b K M
0
0 5 10 15
t, sec
-2
g

Fig. P3.8  a System with a


force source, mass, spring, and a x,v b x,v
damper. b Schematic annotated b b
with nodes of distinct values of 1 g
velocity. c Linear graph of the
system. d Force input applied to
F(t) F(t)
M M
the system. The system was in g g
steady-state under a previous step
input at time, t = 0, when pulse K K
input, F( t), was applied

Frictionless rollers Frictionless rollers

c 1 d
4
F(t)
F(t) kN 2
b K M
0
0 5 10 15
t, sec
-2
g

Fig.  P3.8d was applied. The parameter values are Case I: Check the units of the system equations in terms of power
b = 400 N · sec/m , M = 250 kg, and K = 5, 000 N/m and variables and time.
Case II: b = 400 N · sec/m, M = 25 kg, and K = 1, 200 N/m. 3.8.b The input force, F( t), plotted in Fig. P3.8d is applied
3.8.a Derive the system equation that relates the applied to the damper at time t = 0. Determine the unit step
velocity input to response of the system equations, derived in part a.
i The force acting to accelerate the mass. Use superposition to solve the system equations for
ii The velocity of the mass. the velocity input, shown in Fig. P3.8d.
iii The force acting through the spring. 3.8.c Plot the responses using Mathcad or MATLAB.
Problems 189

Fig. P3.9  a System with a ve-


locity source, damper, mass, and
a x,v b x,v

spring. b Schematic annotated b K b K


v(t) g v(t) 1 2
with nodes of distinct values of M g
M
velocity. c Linear graph of the
system. d Velocity input applied
to the system. The system was
in steady-state under a previous Frictionless rollers Frictionless rollers
step input at time, t = 0, when the
pulse input, v( t), was applied c b
d 100
1 2
50

v(t)
cm 0 0
___ 5 10 15
v(t) M K sec t, sec
-50

g g -100

Fig. P3.10   a System with


a velocity source, damper,
a x,v b x,v

mass, and spring. b Schematic b K b K


v(t) g v(t) 1 2
annotated with nodes of distinct M g
values of velocity. c Linear M
graph of the system. d Velocity
input applied to the system. The
system was in steady-state under Frictionless rollers Frictionless rollers
a previous step input at time,
t = 0, when pulse input, v( t), c b
d 100
was applied
1 2
50

v(t)
cm 0 0
___ 5 10 15
v(t) M K sec t, sec
-50

g g -100

Problem 3.9 A translational mechanical system which con- 3.9.b The input velocity, v( t), plotted in Fig. P3.9d is applied
sists of a velocity source, a damper, a mass, and a spring is to the damper at time t = 0. Determine the unit step
shown in Fig. P3.9. The system was in steady-state under response of the system equations derived in part a.
a previous step input, before the velocity pulse shown in Use superposition to solve the system equations for
Fig.  P3.9d was applied. The parameter values are Case I: the velocity input, shown in Fig. P3.9d.
b = 600 N · sec/m , M = 150 kg, and K = 500 N/m and 3.9.c Plot the responses using Mathcad or MATLAB.
Case II: b = 600 N · sec/m, M = 150 kg, and K = 5, 000 N/m.
3.9.a Derive the system equation that relates the applied Problem 3.10 A translational mechanical system which con-
velocity input to sists of a velocity source, a damper, a mass, and a spring is
i The force acting to accelerate the mass. shown in Fig. P3.10. The system was in steady-state under
ii The velocity of the mass. a previous step input, before the velocity pulse shown in
iii The force acting through the spring. Fig. P3.10d was applied. The parameter values are Case I:
Check the units of the system equations in terms of power b = 400 N · sec/m , M = 250 kg, and K = 5, 000 N/m and
variables and time. Case II: b = 800 N · sec/m, M = 50 kg, and K = 2, 000 N/m.
190 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

3.10.a Derive the system equation that relates the applied The alternative method to cutting and pasting is to define
velocity input to a function of two variables, time, t, and the time shift, ts,
i The force acting to accelerate the mass.
ii The velocity of the mass. Fb2 ( t,ts ) := e − 2· (t − ts) · Φ ( t − ts )
iii The force acting through the spring.
Check units of the system equations, in terms of power vari- The response function is the sum or superposition of Fb2 ( t,ts)
ables and time. with the same scaling factors used for the input,
3.10.b The input velocity, v( t), plotted in Fig. P3.10d is
Fb ( t ) := Fb2 ( t,0)·10 − Fb2 ( t,1)· 20 + Fb2 ( t,2)·10
applied to the damper at time t = 0 . Determine the
unit step response of the system equations, derived
in part a. Solve the system equations for the velocity MATLAB: Plotting Superposed Functions
input shown in Fig. P3.10d.
3.10.c Plot the responses using Mathcad or MATLAB. We will revise the script, Plotting.m, from the Appendix of
Chap. 2 to plot superposed, scaled, and time shifted step re-
sponse functions. The duration of the plot is now the sum of
Chapter 3 Appendix the duration of the input function plus a minimum of five
times the maximum time constant of the system to allow the
Mathcad: Plotting Superposed Functions response to reach steady-state after the pulse input ends. If
there is an existing step input acting on the system prior to
Superposed, scaled, and time shifted step responses can be time t = 0 , then we must add an additional duration of five
plotted in Mathcad, either by cutting, pasting, and editing the time constants.
response function or by defining a function of two variables: if statement
time, t, and time shift, ts. We will use the input, F( t), and There is no Heaviside step function in MATLAB, because
response function, Fb( t), of Sect. 3.8.1 as examples, there is no need for one. We will implement the equivalent
of a Heaviside step function, by using an “if” statement. If
F (t ) = 10 N us (t ) − 20 N us (t − 1) + 10 N us (t − 2)
(3.61) statements permit the conditional execution of a block of
code. The syntax of an if statement varies with the program-

K
t −
K
(t −1) ming language, but it is a fundamental instruction in all pro-
 Fb (t ) = 1 N e b
10 us (t ) − 1 N e b
20 us (t − 1)
(3.62) gramming languages. The essence of an if statement is as
K
− (t − 2 ) follows:
+1N e b
10 us (t − 2)
if (expression A Relational Operator expression B) is
true then execute the instructions which follow.
Evaluating Eq. 3.62 for the parameter values, K = 5, 000 N/m
and b = 2,500 N · sec/m , yields if (expression A Relational Operator expression B)
Instruction executed if conditional is true
Instruction executed if conditional is true
Fb (t ) = 1 N e − 2t 10 us (t ) − 1 N e − 2(t −1) 20 us (t − 1)

+ 1 N e − 2(t − 2)10 us (t − 2) Instruction executed if conditional is true


end
Editing, cutting, and pasting a function in Mathcad require
attention to the location and extent of the L-shaped insertion The instructions between the conditional statement with the
cursor. The input function assignment statement is the sum relational operator and the end statement are executed if the
of three Heaviside step function, and Mathcad’s notation for conditional statement is true. If the conditional statement is
the Heaviside step is Φ() false then execution (or “control”) skips to the instruction
following the end statement. Important: MATLAB is case-
F ( t ) := 10 · Φ ( t ) − 20 · Φ ( t − 1) + 10 · Φ ( t − 2) sensitive. The “keywords,” “if” and “end”, are all lowercase.
MATLAB highlights the leading if and the closing end in
Fb ( t ) := e − 2·t ·10 · Φ ( t ) − e − 2· (t −1) · 20 · Φ ( t − 1) blue to demarcate the limits of the instructions which will
+ e − 2· (t − 2) ·10 · Φ ( t − 2) execute, if the conditional statement is true. To further de-
marcate the if statement as a “block” of code, it is good prac-
tice to indent the instructions which are executed by the if
The subscript b of Fb ( t ) is a literal subscript, created by typ- statement.
ing a period after F, as appears here, F·b(t).
Superposition Using Nested Loops 191

MATLAB’s relational operators are listed in Table 8.6, MATLAB Time Shift


reproduced below. 1
Table 8.6   MATLAB’s relational operators
0.8
< Less than

Amplitude
> Greater than 0.6

< = Less than or equal to 0.4


> = Greater than or equal to
0.2
= = Equal
~ = Not equal 0

0 1 2 3 4 5 6 7
Note the use of a double equal sign for the relational oper- time, seconds
ator, Equal. MATLAB uses a single equal sign as its assign-
ment operator. Also note the symbol for Not Equal, ~ =. Most Fig. A3.1   Time shifted re­sponse function plot created with TshiftPlot.m
programming languages use an exclamation point as the and then formatted
“Not” or logical inversion operator. However, MATLAB’s
symbol, ~ =, reads as “approximately equal,” which is why % Time shift ts. Positive value for shift into t>0.
MATLAB chose it to represent Not Equal. The relational ts = 3
operator Equal fails unless the two variables or expressions % Duration of calculation equals time shift + seven
compared are not exactly equal. For this reason, the Equal % time constants
operator should not be used with “floating point” (scientific tmax = ts + 7 * tau
notation) variables or expressions. Floating point variables % Time Step
carry so many figures it is unlikely that two variables will be dt = tau/200
exactly equal after any significant amount of computation. % Number of iterations
N = tmax/dt
The Equal operator should only be used with MATLAB’s
% Beginning of for loop
“logical” (Boolean) variables or with integer variables or ex-
for n=1:N
pressions. MATLAB’s logical variables can have one of two
t(n)=(n-1)*dt;
values: true or false and equivalently 1 or 0. MATLAB is
% Beginning of if statement
case-sensitive; “true” and “false” are all lowercase. if (t(n) - ts >= 0);
Us = 1;
end;
Time shift % End of if statement
% Response function
The MATLAB equivalent of a Heaviside step function is the y(n) = Us * (1 - exp(-(t(n)-ts)/tau));
following if statement, where t is the time vector, ts is the end
time shift in seconds, and Us is the variable, which plays the % End of for loop
role of the unit step. Us is initialized with the value of zero plot(t,y)
before its use.

Us = 0

Superposition Using Nested Loops


if (t(n) - ts >= 0);
Us = 1; The superposition can be performed, using nested loops or
end; repeated function “calls.” We will defer the topic of user-
defined functions in MATLAB until Chap. 8. Briefly, a func-
This logic is illustrated in the following code, which plots a tion is a script with an “argument list,” allowing it to be used
stable exponential growth with a time shift of 3 sec. in MATLAB instructions in the same manner as familiar
mathematical functions, such as sin().
% TshiftPlot.m “Nested” loops are two or more loops, where an “inner”
% Time shifted step response plotting script loop completely iterates through the range of its counter
% Initialize unit step variable Us to zero variable for each single iteration of the outer loop’s counter
Us = 0 variable. Fig. A3.2 is a flowchart of a nest loop which initial-
% Time Constant izes a three by four array with ones.
tau = 0.5
192 3  Introduction to the Linear Graph Method, Step Responses, and Superposition

Enter Outer Loop Start


of the Nested Loops

Assign values to scalar variables


tau_max, tchar, and nsteps
r=0
Assign values to vector variables
ts( ) and In( )

r=r+1 for loop to initialize vector variable


Us(m)=0, for m= 1 to nsteps
c=0
Enter Inner Loop Calculate duration tmax, time step dt,
and number of interations N

Initialize outer loop counter


c=c+1 n=0
Array(r,c) = 1 Enter Outer Loop

n = n+ 1
t(n) = (n-1)*dt
No y(n) = 0
c == 4 ? m=0
Enter Inner Loop

Yes
Exit Inner Loop m=m+1

No
No t(n) - ts(m) >= 0 ?
r == 3 ?
Yes
Yes
Us(m) = 1
Exit Outer Loop

Fig. A3.2   Flowchart of nested loops which creates a three by four


y(n) = y(n) + Us(m)*In(m)*StepResponse(t(n) - ts(m))
array named Array filled with ones

We will use a similar structure of two nested loops to per- No


m == nsteps ?
form superposition of scaled and time shifted response func-
Yes
tions, to plot the response function for a pulse input. We will Exit Inner Loop
demonstrate the procedure using the response of the damper
force, Fb( t), of the force source spring-damper system of No
n == N ?
Fig. 3.54 to the input function, Eq. 3.59. The response func-
tion, Fb( t), Eq. 3.62, evaluated using K = 5, 000 N/m and Yes
b = 2,500 N ⋅ sec/m is the following: Exit Outer Loop

plot(t,y)
K K
− t − (t −1)
 Fb (t ) = 1 N e b
10 us (t ) − 1 N e b
20 us (t − 1) Fig. A3.3   Flowchart of MATLAB script, SuperposedResponse.m
K
(3.62)
− (t − 2 )
+1N e b
10 us (t − 2)
the time constant, and it is, therefore, the value of both
The pulse response plotting program is structured as shown variables tau_max and tchar. Next, the vector variables for
in the flowchart, Fig. A3.4. The program begins with as- time shifting and scaling the input function’s Heaviside step
signment statements which define the maximum time con- function are assigned values. The vector of time shifts is
stant, the minimum characteristic times of the system, and named ts. The vector of scaling factors is named In. A vec-
the number of step inputs, nsteps. In the example of a first- tor variable named Us will play the role of the Heaviside
order step response, there is only one characteristic time, step functions. It is initialized to zero using a for loop. The
References and Suggested Reading 193

MATLAB Plot of Fb(t), Eq.3.62 % Input time shift ts(m). Positive value for shift into t>0.
10 ts(1) = 0
5
ts(2) = 1
ts(3) = 2
0 % Input scaling factors
Fb(t), N

In(1) = 10
−5
In(2) = -20
−10
In(3) = 10
% Initialize unit step vector Us(m)to zero
−15 for m=1:nsteps
Us(m)=0
−20
0 0.5 1 1.5 2 2.5 3 3.5 end
t, sec % Duration of calculation equals time shift +
% seven time constants
Fig. A3.4   Response func­tion Eq. 3.62 plotted with MATLAB script, tmax = ts(3) + 7 * tau
SuperposedResponse.m, and then formatted % Time step
dt = tchar/200
duration of the calculation, tmax, depends on both the maxi- % Number of iterations of the for outer loop
mum time constant of the system and the duration of the N = tmax/dt
input function. % Beginning of the outer for loop
The outer loop increments through the time step of the for n=1:N
calculation, as in the case of a single step response. It begins t(n)=(n-1)*dt;
by calculating the current value of the time variable, t(n), y(n)=0;
and initializing the current value of the output variable, y(n), % Beginning of the inner for loop
to zero. Control then passes to the inner loop, which incre- for m=1:nsteps
ments through the number of step inputs in the pulse input % if statement to transition unit step variable Us(m)
function. The inner loop’s if statement checks the time shift, if (t(n) - ts(m) >= 0);
ts(m), against the time variable, t(n). If the unit step should Us(m) = 1;
transition from zero to one, the unit step variable, Us(m), is y(n) = y(n) + Us(m) * In(m) * ( exp(-(t(n)-ts(m))/0.5 ));
assigned the value of one. If the test fails, Us(m) retains its end;
initialization value of zero. The product of the input scaling % End of if statement
factor, In(m), the unit step variable, Us(m), and the unit step end;
% End of inner for loop
response function evaluated for t(n) with the time shift ts(m)
end
is summed to the output variable, y(n).
% End of outer for loop
The MATLAB script, SuperposedResponse.mfollows.
plot(t,y)
% SuperposedResponse.m
% Superposition Plotting Script
% References and Suggested Reading
% Maximim Time Constant
tau_max = 0.5 Hildebrand FB (1976) Advanced calculus for applications, 2nd edn.
% Minimum Characteristic Time Prentice-Hall, Englewood Cliffs
tchar = 0.5 Rowell D, Wormley DN (1997) System dynamics: An introduction.
Prentice-Hall, Upper Saddle River
% Number of steps Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
nsteps = 3 dynamics. Addison-Wesley, Reading
Mechanical Systems
4

Abstract
A mechanical system is a system in which the dominant forms of energy storage, transfer,
and dissipation are described by Newton’s laws. Fluid systems are described by Newton’s
laws, but use different variables due to fluid mass. They will be addressed in Chap. 5.
Mechanical energy is stored as kinetic energy and as strain energy. Gravitational potential
energy is presented as a force source. Mechanical energy is dissipated due to shear of a
fluid, a material in its plastic state, or between two solid surfaces. In order to work with
scalar equations, motion is restricted to single axes. Translational and rotational motions are
separate energy storage modes. The parameter values of the mechanical elements can be
calculated from the geometry and material properties of mechanical components in many
instances. Otherwise, the parameters are experimentally determined by dynamic tests. The
mechanical properties of viscoelastic materials, which include natural and synthetic poly-
mers and biological tissues, are described by the dynamic response.

4.1 Translational Mechanical System tion). Gravitational potential energy will not be represented
Elements as an energy storage mode in translational mechanical sys-
tems. Gravity will be represented as a force source acting
Translational mechanical system elements were introduced on a mass element. Kinetic energy is dissipated as heat
in Chaps. 1 and 2 and then used in Chap. 3 to introduce the through friction, which is lost from the mechanical system.
linear graph method. This chapter investigates how to esti- We will categorize mechanical systems based on the type
mate the linear element equations from analytical models of of motion, either translation (linear) or rotation. Further,
machine components and dynamic tests. Rotational systems we will restrict motion in a subsystem to one dimension, so
are then introduced and the similarities and differences be- that we can work with scalar velocities rather than vector
tween translational and rotational elements are investigated. velocities. It may be that we will need a number of subsys-
We begin with a brief review and summary of transla- tems, each representing motion in one dimension, to repre-
tional mechanical power, energy, elements, and models. sent the motion of a machine.
An energetic model of a mechanical system does not have In translational mechanical systems, power, the flow rate
a physical resemblance to the system. An energetic model of energy, is the dot product of force and velocity,
represents the energetic behavior of the mechanical system
as the power flow in a network of interconnecting elements  dW
= P = F ·v (1.17)
which supply, store, or dissipate energy. Only the dominant dt
aspects of a system’s energetic nature are modeled, in part,
not only for economy of effort but also due to the limitation The elemental and energy equations for translational me-
of our knowledge of a system’s actual behavior. chanical systems written in terms of the power variables
Mechanical systems store energy as kinetic energy (en- force and translational velocity are summarized in Table 4.1.
ergy of motion) or strain energy (energy of elastic deforma-

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_4, © Springer Science+Business Media New York 2014 195
196 4  Mechanical Systems

Table 4.1   Translational mechanical elements only a limiting case. For example, aluminum does not have a
Elemental equation Energy equation linear elastic response. The modulus is defined by a line off-
dv1g 1 2 set from and parallel to the approximate slope of the initial
Mass Fm = m Em = mv1g
dt 2 region of the stress–strain curve. The greatest uncertainties
dFK FK2 in material properties tend to be in those governed by flaws
Spring = Kv12 EK =
dt 2K within the material, such as strengths, and the condition of
Damper Fb = bv12 None. the surface of the material, such as friction.
Energy dissipater Although many material properties are presented as if
they were certain, the true uncertainty may not be apparent.
An example of the uncertainty inherent in material proper-
ties is the yield strength of metal alloys. Yield strengths
4.2  Modeling Translational Elements for different grades of steel and aluminum are published
in various handbooks, but what do the values represent?
All engineering calculations are estimates to some degree. The published yield strength is surely not the average yield
Reasonably accurate estimates of the element parameters strength. If it were, then half of all specimens would yield
for the elemental and energy storage equations are needed below that value. The mean yield strength is surely greater
in order to use dynamic system theory. The difficulty level than the published yield strength, but that value and the
in obtaining a reasonably accurate estimate varies with the standard deviation are not available since they would re-
type of physical component and attribute to be modeled. Dif- veal the steel mill’s control over the manufacturing pro-
ficulty also varies with the point in the engineering cycle, cess, which is proprietary information. Further, strength
i.e., does the physical component exist or is it still in design? is affected by the processing history of the part. Although
There is also a chicken-and-egg aspect to the computation. the history of a part is well documented in some instances,
Our “lumped” models include elements to represent signifi- such as in aerospace applications, in others, it may be un-
cant energetic attributes of the system, where significance is known.
defined as the amount of energy stored or dissipated in an at- We will begin with the simplest approach, calculating
tribute relative to the other attributes of the system. We must elemental parameters using published material properties,
first compute the elemental parameter before we can judge and then consider how to estimate parameters from experi-
the significance of the element in the system. mental data. Most of the materials and devices we work
When possible, the best estimates of elemental param- with are non-linear. We investigate the process of linear-
eters are calculated from the dynamic response of systems. ization in Sect. 4.2.4.1. We shall see that “linearizing about
Characterization of the transient period of first-order step re- an operating point” is only accurate over a limited range
sponses yields the time constant. The steady-state value of a of the system’s power variables. If a single straight-line
step response yields a coefficient. Likewise, as we will see approximation of a constitutive relationship is not accurate
when we work with second and higher-order systems, the enough to yield useful results from linear calculations (i.e.,
transient period of an oscillatory underdamped second-or- superposition), then, we must resort to “numerical meth-
der step responses yields the damped natural frequency and ods” so that we can use non-linear models of springs and
damping ratio. The steady-state value of both underdamped friction.
and overdamped second-order step responses yield useful We will defer using non-linear element equations until
coefficients. we have introduced the state-space representation of dy-
If the component, machine, or system does not yet exist, namic systems in Chap. 7 and finite difference methods
parameter values can be estimated using the geometric con- in Chap. 8. The Runge–Kutta algorithm, which we shall
figuration of the machine or system component, values for program in MATLAB, will solve state-space formulations
the various material properties, and estimates of the operat- which include non-linearities. An alternative approach for
ing range of the power variables. If possible, estimates pro- dealing with non-linear energetic elements is the use of dy-
duced by interpolating or extrapolating parameters obtained namic simulation software, such as MathWorks’ Simulink.
from similar, existing, well-characterized components will Simulink has a graphical user interface in which the engineer
have the least uncertainty. If no data are available from simi- constructs a block diagram of the dynamic system. Block
lar components, then estimates must be based on published diagrams are discussed in Chap. 9. Simulink, however, is not
material properties and engineering models. The published addressed in this text for two reasons. First, instruction in
values of “bulk” properties, such as the density of most en- the use of graphically based engineering software which in-
gineering materials or the elastic properties of steel, can be cludes tutorials and examples is unnecessary for an educated
used with great confidence. In other cases, material proper- and enterprising engineer. Second, those readers for whom
ties can be very uncertain with published values representing this text is a reference or part of their continuing self-edu-
4.2  Modeling Translational Elements 197

cation will not have access to MATLAB or Simulink. The at speeds which are substantial fractions of the speed of light.
MATLAB programming exercises can be performed using Macroscopic matter does not. The maximum velocity an en-
free, open-source programming language R, which was cre- gineer may deal with is the escape velocity from earth orbit
ated to resemble MATLAB. Alternatively, those with some of approximately 11 km/sec. The relativistic increase in mass
programming experience may choose to work in a general at that escape velocity is vanishingly small
purpose open-source programming such as C. Simulink is
not free, in fact, it is very expensive for those who do not m0 m0
m= = → m = m0
have student status. 2
1 − 1.34 × 10 −9
 km 
11 
sec
1− 2
4.2.1  Mass, Kinetic Energy Storage Element  km 
 300, 000 
sec 
An element in a model represents a single energetic property.
All physical components in machines possess mass, but the Although the true and non-linear elemental equation for
property of mass is not assigned to all physical components. mass is relativistic, Newton’s second law is extremely accu-
Mass elements are added to the model of a dynamic sys- rate for all practical purposes above the atomic scale.
tem, if the mass of a component (or assembly) accounts for Although the element equation of a mass is linear, it may
a significant amount of power flow or energy storage within still be uncertain due to uncertainty in the parameter value,
the system, relative to power flows and energy storage in i.e., the amount of mass. In practice, we are often unable to
other modes. A significant power flow can be created by a weigh a machine part to determine its mass because we are
large acceleration of a modest mass. Likewise, a significant not allowed to disassemble the machine or the part is too
amount of kinetic energy can be stored in a small mass at large to weigh. We must calculate the object’s mass from its
high velocity. Consequently, the nature of the system and the geometry and density. Variation of the part geometry from
magnitude and rate of change of its power variables, in other that specified in the drawings or, if drawings are not avail-
words, the system’s operating conditions, must be consid- able, the accuracy of our measurements, is a source of error.
ered when modeling a system. Often, the material or the specific alloy is unknown. The un-
F = ma is the familiar form of Newton’s second law. certainty of the estimated mass depends on how variable the
More formally, Newton’s second law equates force and the densities of the alloys of the base material are. The density
rate of change of momentum, of steel can be assumed with certainty. However, the densi-
ties of some metal alloys and most polymers vary with the
dp dmv specific formulation. Finally, when the system is open, in the
F= → F=
dt dt sense of allowing mass flow across the system boundary, the
accuracy of the measurement of the mass flow rate limits the
If mass m is constant, then precision with which we know the mass of the system.
The precision of our result is no greater than the least
dv precise datum used in the calculation. However, precision
F=m
dt is usually the lesser of our concerns. After we take all our
Restricting motion of a single dimension (or direction) al- measurements and calculations, a final uncertainty in the pa-
lows this vector equation to be written in scalar form rameter value remains in how representative was the part of
the population from which it was drawn. The uncertainty or
dv variability in the values used in calculating the parameters
F=m
dt has a greater effect on limiting the precision we report for
our result. We must decide how many significant figures of
As you know from your study of physics, as part of his the- the result we believe and report no more than those.
ory of relativity, Einstein revised Newton’s second law to Mass is the easiest energetic attribute to calculate when
correct for the velocity dependence of mass a machine component is (1) in translational motion, (2) can
be reasonably modeled as rigid (i.e., all points on the com-
m0 ponent have the same velocity), and (3) the density/ies of the
m=
v2 material(s) which comprise the part are known.
1−
c2 Sadly, a common error is made when computing the mass of
a machine component. Mass is not weight. Mass density does
where mass m0 is the “rest mass” and c is the speed of light not equal unit weight. Weight is a force. Specifically, weight is
in a vacuum, 300,000 km/sec. Subatomic particles can move the force of gravity acting on mass at sea level on earth
198 4  Mechanical Systems

F Table 4.2   Density variation of aluminum alloys


F = Ma → M = Aluminium alloys Density (kg/m3)
a
Lowest Highest Percent variation (%)
2000 Series 2,580 2,850 10.5
and
6000 Series 2,680 2,740  2.2
weight = Mg Casting alloys 2,650 2,950 11.3

where g, the gravitational constant, is the acceleration of What should an engineer do when the aluminum alloy
gravity at sea level is unknown? Play the odds. Assume the density of an alloy
commonly used for the application. For example, if one
ft m were calculating the mass of an aluminum component which
g = 32.2 2
= 9.81
sec sec 2 was machined, as opposed to cast, then Aluminum 2024,
ρ2024 = 2780 kg/m3, or Aluminum 6061, ρ6061 = 2700 kg/m3 ,
Pound mass is a unit used to perform fluid mechanics and would be reasonable guesses, yielding a 3 % uncertainty. It is
thermodynamics calculations in the US customary units. A a mistake to report a result with more significant figures than
pound mass is the mass of a material which weighs a pound can be justified by the certainty of the information available.
force if the mass is not accelerating. Students who have a For the 2000 Series alloys, one certain figure and one es-
taken the short-sighted formulaic approach to their engineer- timated figure is reasonable. If you believe the alloy to be
ing studies become confused by “pound mass” because they 2024, then use its density to three significant figures in your
do not understand why the gravitational constant is included calculation, but only report your result to two.
in the calculations. The confusion is compounded by the vari- Actual machine components are not rigid. When a part is
ety of peculiar units used in the US customary units to express loaded in compression or tension within its elastic response,
weight and gravitational acceleration. There is no pound mass the deformation due to elastic strain is small unless the part
in dynamics and system dynamics. Pound is a unit of force. is very large. Although there is kinetic energy stored dur-
The unit of mass in the US customary units is a “slug” ing elastic axial deformation, it is usually negligible. When
a part is loaded in elastic bending, the maximum displace-
weight ment and the kinetic energy stored during bending of a small
weight = Mg → slug ≡
g part can be significant. A tuning fork is an example. Another
example is a dynamic experiment performed by all students
where weight is in pounds and the gravitation constant early in their education by bending and then releasing a ruler
g = 32.2 ft /sec 2 . The units of a slug are: cantilevered from a table. The resulting vibration or oscil-
lation requires both strain and kinetic energy storage in the
 lb  single physical object, the ruler. The displacement and veloc-
 weight 
[slug ] =  g  → [ ]  ft
slug = 

ity of the cantilever increase with distance from its support.
   sec 2  The lack of rigidity leads to the distributed mass of the ruler
not storing kinetic energy uniformly but as a function of po-
 lb ⋅ sec 2  sition along the cantilever. The “effective” mass is calculated
[slug ] =   by “weighting” (no pun intended) the mass at a point by its
 ft  displacement or velocity using a “shape” function. We will
formulate the displacement of a cantilevered beam loaded at
It is far easier to avoid wrestling with the US customary units its end in Sect. 4.2.3.4, when we investigate elastic energy
and convert to SI units at the beginning of the calculation in bending and calculate the effective mass of a cantilevered
and then convert back to the US customary units at the end, beam in Sect. 4.2.3.2.
if required to report the results in the US customary units.
Although mass is easy to calculate if the geometry of a
machine component is known, and a published value for den- 4.2.2  Spring, Strain Energy Storage Element
sity is available, it is often the case that only the type of mate-
rial—but not the particular alloy metal or type of plastic—is Springs of many different shapes and sizes are used in ma-
known. For example, say a metallic material is identified as chine design. Compression and extension coil springs are the
aluminum. Aluminum, a low density metal, ρ = 2700 kg/m3, most familiar. Other common springs are torsion coil springs,
is commonly alloyed with copper, a very dense metal, leaf springs, and belleville springs. Torsion coil springs are
ρ = 8960 kg/m3. As a result, aluminum alloys vary signifi- used in mouse traps and clothes pins. Leaf springs consist
cant in density, Table 4.2. of a single or a nested stack of rolled or bent steel strips.
4.2  Modeling Translational Elements 199

Large leaf springs are used in vehicle suspensions. Belleville Elastic Deformation
washers, also known as disk springs, are washers or disks ∆L
deformed into a profile which approximates a fraction of a
L
F(t) F(t)
sphere. An endless variety of custom-designed coil springs
can be made on spring machines.
Any matter that deforms elastically stores strain energy. If Cross-sectional area A
the amount of strain energy stored in a machine component Young’s modulus E
is significant relative to the amount of energy stored else-
Fig. 4.1   Axially loaded element with elastic deformation
where in the system, then the machine component’s attribute
of strain energy storage is included in the energetic model
of the system. A component of a given size loaded in bend- Table 4.3   Moduli of carbon and alloy steels (excluding stainless steels)
ing undergoes substantially more deformation than the same Young’s Modulus 28.5 to 30.0 × 106 psi
component loaded axially. Shear Modulus 11.0 to 11.9 × 106 psi

4.2.2.1  Translational Spring Constant


The most familiar expression for a translational spring is Calculation of a translational spring constant is an exer-
cise in deriving a force–displacement or force–deflection
FK = K ∆ x
(1.6) relationship for a machine part of assembly. The force–dis-
placement relationship is then expressed as the ratio
∆ x is the conventional notation for the deformation or
change in length of a spring and K is the stiffness, spring  Force Force
or = Spring Constant ≡ K
rate, or spring constant. It is important to keep the distinc- Displacement Deflection
tion between position, displacement, and deformation clear. (4.2)
A displacement is a change from an initial position. Know-
ing the displacements of both ends of the springs allows us 4.2.2.2  Axial Strain
to calculate the deformation of the spring. Calculation of the spring constant of an element loaded axi-
The reciprocal of stiffness is called “compliance” ally in tension is an application of the definition of linear
strain, Fig. 4.1. Compressive loading must always be con-
1  deformation  sidered as a critical calculation due to the possibility of a
K
≡ compliance where [compliance] =  force  catastrophic buckling failure, which is outside the scope of

this text.
(4.1) For tensile axial loads below yield and compressive axial
Compliance is often given the symbol C. This is unfortunate loads below buckling, the extension or compression of an
because C is also used in place of b as the symbol for “pro- axial member is the product of the length of the member and
portional” or linear damping in vibration analysis. the linear strain.
Translational spring constants can be calculated if the ge-
ometry of the machine elements and the moduli of the materi- σ Faxial L
Elastic Deformation ∆ L = ε L = L=
als are known or if a reasonable estimate is available. Most E AE
materials do not have a true Young’s modulus, since their
Force
stress–strain relationships are non-linear. Steel is an important The ratio yields the spring constant K
exception, since there is a linear region of the stress–strain Displacement
curve below the yield stress. In contrast, the slope of the initial 
Force F AE
portion of an aluminum alloy’s stress–strain curve is used in = axial = ≡K (4.3)
the 0.2 % offset construction to establish a “yield” stress. Displacement ΔL L
Although the moduli of metals may be only linear
approximations, the stress–strain behavior of metals below We often know very little about the materials in a system
yield is considered a bulk property and insensitive to mi- we need to analyze or modify and must use published mate-
croscopic material flaws and common manufacturing opera- rial properties in our calculations of the energetic properties.
tions. Hence, one can place greater confidence in published Fortunately, the Young’s modulus and shear modulus values
values of moduli of metals than in, say, values of ultimate of carbon steel vary relatively little as a function of carbon
strength. content, processing (i.e., cast vs. cold rolled), and heat treat-
ment, Table 4.3.
200 4  Mechanical Systems

Fig. 4.2  a Linear bending stress


distribution b Infinitesimal force
a b y
and moment -σz y

x
Neutral y
Axis z dy

σz dFz = σz dA
+σz
dM x = ydFz = yσz dA

4.2.2.3  Deflection in Bending axial stress distribution. The area moment of inertia com-
Forces acting transverse to a member induce a bending mo- bines dependence on distance from the neutral axis of both
ment. The deflection of the point of application of the force the linear axial stress distribution and the resisting moment
is calculated by beam theory. In contrast to deflection under created by that stress acting on an elemental area of the
axial forces, the deflection due to bending does not depend cross-section.
of the cross-sectional area of the beam. It depends on the The deflection equations can be derived from the relation-
area moment of inertia I, a purely geometric quantity. ships for linear bending stress
The connections, or support conditions, at the ends of an
element in bending greatly influence the magnitude of the   −y 
σ z = σ z max  (4.4)
deflection of the point of application of the force. In order  ymax 
to calculate bending deflection manually, the support condi-
tions must be statically determinant. This is generally not the infinitesimal internal force
case in machine design, where statically indeterminate con-
nections are stiffer and can carry greater loads because they dFz = σz dA
(4.5)
have lower local stresses.
Modeling a statically indeterminate machine element in and infinitesimal internal moment
bending as if it were a statically determinant beam will typi-
cally overestimate the bending deflection, yielding a spring  dM x = ydFz (4.6)
constant which is smaller than the actual. Conversely, the
energy stored in bending strain will be overestimated be- Expressing the infinitesimal internal moment in terms of the
cause of the inverse relationship between the spring con- linear bending stress distribution yields, Fig. 4.2
stant and strain energy storage. Manual calculations are still
useful as initial estimates and may suffice, if other energetic   −y   σz max 
parameters of the system are as uncertain. However, when dM x = y σz max  dA = − 
2
 y dA (4.7)
 ymax   ymax 
more refined estimates of the spring constant of a com-
ponent with indeterminate end conditions or non-uniform
cross-section are warranted, then finite element calculations where the area moment of inertia is
are necessary.
A brief review of elastic beam theory follows. The as-  I = ∫ y 2 dA (4.8)
sumption that planar cross-sections of a member remain
planar when the member is bent allows the radius of curva- The assumption of a linear stress distribution allows the mo-
ture of the member to be related to the rotation of a plane ment to be related to the radius of curvature in terms of the
section and, consequently, to the axial strain distribution Young’s Modulus, E, and the area moment of inertia of the
across the beam in the y-direction. If the material model cross-section, I
is linearly elastic, i.e., σ = E ε , the assumption that plane
sections remain plane leads to a linear variation in axial  1 M ( x)
= (4.9)
stress with distance from the neutral axis. Moment equi- rcurvature EI
librium of a free body cut from the beam results from the
4.2  Modeling Translational Elements 201

y Substituting the expression for the internal moment into


x
Eq. 4.10 and evaluating the integral for a cantilevered beam
L with a constant cross-section and Young’s modulus
FR F
y
d2y F
FR
x ∫ dx 2
dx = −
EI ∫ ( L − x ) dx
∆y
MR
dy F  x2 
=−  Lx − + C1 
Area Moment of Inertia I dx EI 2 
Young’s Modulus E

Fig. 4.3   Model of a cantilevered beam dy F x2


∫ dx dx = − E I ∫ Lx − 2
+ C1 dx

y
F  Lx 2 x 3 
x y=−  − + C1 x + C2 
L-x F EI  2 3 
FR
y
Mz Mz Fy
FR
x The end conditions of a cantilevered beam are that the dis-
MR
Fy placement and slope of the beam are zero at the support

dy
Fig. 4.4   Free body diagram of beam cut at distance x from the support y ( 0) = 0 and =0
dx

Deflection of a beam is calculated by approximating the cur- F  L02 03 


y ( 0) = −  − + C1 0 0 + C2  = 0 → C2 = 0
vature of the beam for small deflections as EI  2 0 3 0 
 1 d2y
≈ (4.10)
rcurvature dx 2 dy ( 0) F  02 
=−  L 0 0 − + C1  = 0 → C1 = 0
dx EI  2 0 
Equating Eqs. 4.9 and 4.10, we get
The deflection equation for a cantilevered beam loaded at its
 d 2 y M ( x) end a distance L from the support is
= (4.11)
dx 2 EI
 F  Lx 2 x 3  (4.12)
y ( x) = − − 
Integrating Eq. 4.11 twice with respect to x yields the dis- E I  2 6
placement y( x) of the beam. Evaluation of the integral re-
quires knowledge of the variation of the bending moment The spring constant calculation requires the deflection at the
along the beam M ( x), which is found from statics; hence, point of application of the load. Evaluate Eq. 4.12 for x = L
the requirement of statically determinant beams. The two
constants of integration require knowledge of the slope and F  L3 L3  F  6 L3 − 2 L3  F L3
y ( L) = −  − =−   =−
displacement somewhere on the beam. EI  2 6  EI  12  3E I
Two of the most commonly used models of translation-
al springs are a cantilevered beam and a simply supported  Force −F 3E I
beam. = 3
= 3 ≡ K Cantilevered Beam
Displacement FL L (4.13)
− Loaded at End

4.2.2.4  Cantilevered Beam 3E I


The moment on an internal cross-section of the cantilevered
beam shown in Fig. 4.3 is found by creating a free body of 4.2.2.5  Simply Supported Beam
the right-hand section of the beam with an imaginary cut a Derivation of the deflection equation of a simply supported
distance (L–x) from the end, Fig. 4.4, and then using mo- beam, Fig. 4.5, with a single load is more involved than that
ment equilibrium at the cut of the cantilevered beam, because the end conditions lead to
the elimination of one of the two constants of integration.
M cantilevered ( x ) = − F ·( L − x )
The reaction forces are:
202 4  Mechanical Systems

Fig. 4.5   Model of a simply


supported beam with a single y
concentrated load F L
x χ
F
a b

+M +M

∆y
FR
x
Area Moment of Inertia I
FR Young’s Modulus E FR
y yB
A

F ·b F ·a We will evaluate the special case of loading the beam at


FRA = − and FRB = −
L L mid-span first. Symmetry leads to the two reaction forces
being equal
and the moments on cross-sections through the beam are:
L L
M simply ( x ) = FR · x for 0 ≤ x ≤ a −F ⋅ −F ⋅
supported
A
F RA = 2 =−F and F RB = 2 =−F
L 2 L 2
and
L
The slope and deflection equations for 0 ≤ x ≤ are derived
M simply ( x ) = FR ·( L − x )
B
for a ≤ x ≤ L 2
supported
by integrating the relationship between moment and bending
twice with respect to x
Evaluating the integral for a simply supported beam is
more involved than for a cantilevered beam. The end con-
d2y FR L
ditions of a simply supported beam are that the displace- ∫ dx 2 dx = − E IA ∫ xdx for 0 ≤ x ≤
2
ment of the beam is zero at the supports. However, the
slope of the beam at the supports is unknown, because the FR  x2
dy  L
beam is free to rotate. Hence, only a constant of integration = − A  + C1  for 0 ≤ x ≤
can be determined by using the displacement of the end dx EI  2  2
of the beam. It is tempting to assume that the maximum F RA
dy x2 L
deflection of the beam occurs at the location of the load, ∫ dx dx = − E I ∫ 2
+ C1 dx for 0 ≤ x ≤
2
but that may not be true and cannot be established until
the deflection equation is known. The one exception is the F RA  x 3  L
special case loading the beam at its mid-span, where sym- yA ( x) = − + C1 x + C2  for 0 ≤ x ≤
E I  6  2
metry requires the slope of the deflected shape to be zero
at the location of the load. In the general case when the
L
load is not at the mid-span, what we know must be true is Evaluating the slope equation at x =
2
that the deflection of the point of application of the load
and the slope of the beam at that location is the same, re-  L
gardless of whether we formulate the deflection and slope dy     L 2 
 2 FR
equation from support A or support B. Consequently, the =0=− A   2  
dx EI  + C1 
second constant of integration is found by equating the de-  2 
flection and the slope of the point of application of the
load calculated from both supports. The general case is no  L
2

more conceptually challenging than the symmetrical spe-   L2


2
cial case but the algebra does become involved. + C1 = 0 → C1 = −
2 8
4.2  Modeling Translational Elements 203

Evaluating the deflection equation at x = 0 F R A  03 


y A ( 0) = 0 = −  + C1 ⋅ 0 0 + C2  → C2 = 0
EI  6 0 
F R A  03  L2  
y A ( 0) = 0 = −  −   ⋅ 0 + C2  → C2 = 0
E I  6 0  8   Hence
0

F RA  x 3 
FR  x 3 L2  L yA ( x) = −  + C1 x for 0≤ x≤a
y A ( x) = − A  − x for 0≤ x≤ EI  6 
EI  6 8  2
The deflection equation for a ≤ x ≤ L can be formulated in
Substituting for the reaction force at support A terms of x or, more conveniently, in terms of position mea-
sured from support B, χ
F  x 3 L2  L
yMid − span ( x) = − x for 0≤ x≤
2 E I  6 8  2 d2y F RB
Loading
∫ dχ 2
dx =
EI ∫ χ dχ for 0≤ χ ≤b
Our objective is the spring constant (or spring rate) of a
beam bent by a transverse load. Consequently, we evaluate dy F RB  χ 2 
the deflection equation at the point of application of the load, =  + C3  for 0≤ χ ≤b
dχ E I  2 
x = L /2
  L 3  dy FR χ2
 L F   2  L2   
L
∫ d χ d χ = E IB ∫ 2
+ C3 d χ for 0≤ χ ≤b
yMid − span   =  −  
Loading
 2  2E I  6 8 2 
  FR  χ 3 
yB ( χ ) = B
+ C3 χ + C4  for 0≤ χ ≤b
E I  6 
 L F  L3 L3  F  L3 3 L3  FL3
yMid − span   =  − =  −  =−
Loading
 2  2 E I  48 16  E I  84 84  48 E I Evaluating this deflection equation for χ = 0, the location of
support B, eliminates one constant of integration
Signs are always a problem in mechanics. Negative sign of
the displacement equation is correct relative to the positive F RB  03 
y-direction. The applied force is also in the negative direc- yB ( 0) =  + C3 0 0 + C4  → C4 = 0
EI  6 0 
tion, yielding a positive spring constant. Energetic param-
eters must be positive! F RB  χ 3 
 yB ( χ ) =  + C3 χ  for 0≤ χ ≤b
Force −F 48 E I EI  6 
= = ≡ K Simply supported
Displacement F  L3  L3
− Beam Loaded at
E I  48  Mid − Span
The slope of the beam at this location is independent of how
(4.14) we chose to formulate the equations. Our formulation written
The general case of simply supported beam loaded at a in terms of χ approaches the load position from the negative
location, x = a, where a ≠ L /2 is more involved to derive, x-direction. Hence, the slope in terms of χ is the opposite of
because we cannot assume that the slope of the beam is zero the slope in terms of x. Inverting the sign and equating the
at the location the load is applied. We begin as we did previ- two deflection equations evaluated for the location of the ap-
ously and derive the same slope and deflection equations for plied load, x = a and χ = b, yields an equation with constants
0 ≤ x ≤ a, where x is measured from support A of integration, C1 and C3

FR  x2 dy A ( a ) dy (b ) F RA  a 2  F RB  b 2 
dy  =− B → + C = − + C3 
= − A  + C1  for 0≤ x≤a dx dx 
EI  2
1
 
EI  2 
dx EI  2 

F R  x3  F b  a2  F a  b2 
y A ( x ) = − A  + C1 x + C2   + C1  = − + C3 
EI  6 
for 0≤ x≤a EI L 2  E I L  2 

We can evaluate the deflection equation since we know  a2   b2 


y A ( 0) = 0 b  + C1  = − a  + C3 
 2   2 
204 4  Mechanical Systems

Evaluating the two deflection equations at the location of the


− F  a b (b + a ) 
2 2
− F  a 2 b3 + a 3b 2 
applied load yields a second equation with C1 and C3 yA (a) = =  
E I L  3 ( a + b )  E I L  3 ( a + b ) 
F RA  a 3  F RB  b3 
y A ( a ) = yB (b ) →  + C1 a  =  + C3b
EI  6  EI  6  − Fa 2 b 2
yA (a) =
3E I L
F b  a3  F a  b3 
 + C a
1  =  + C3b
EI L 6  EI L 6  We can check this equation against our previous result by
evaluating it for a = b = L /2
 a3   b3 
b  + C1a = a  + C3b  L  L
2 2
 6   6  −F    
− Fa 2 b 2  2  2 − F L3
yA (a) = = =
Form a set of simultaneous equations in terms of C1 and C3 3E I L 3E I L 48E I

a 2 b ab 2 This checks. The result can be expressed in many different


+ = −bC1 − aC3
2 2 forms because L = a + b
ba 3 ab3
− = − abC1 + abC3
− F  a 2b2  − F  a ( L − a ) 
2 2
6 6
yA (a) = =  
E I  3L  E I  3L 
We can now use Gaussian elimination. Multiplying the first
equation by b and adding it to the second equation eliminates
− F  ( L − b) b 
2 2
C3, yielding C1
yA (a) =  
EI  3L 
a 2 b ab 2
+ = −bC1 − aC3
2 2
ba 3 ab3  a 2 b ab 2  Similarly, the spring constant K can have a number of forms.
− + b + = − abC1 + abC3 + b ( −bC1 − aC3 )
6 6  2 2  Of the above, the latter two are the most convenient for com-
putation
a 2 b ab 2 Force −F 3L E I
+ = −bC1 − aC3  = = 2 2 ≡ K Simply Supported
2 2 Displacement  2 2
 ab
F ab Beam Loaded at x = a
ba 3 ab3 a 2 b 2 ab3 −  
− + + = − abC1 − b 2 C1 E I  3L 
6 6 2 2 (4.15)
and
a 2 b ab 2
+ = −bC1 − aC3
2 2  Force −F
=
Displacement F a ( L − a) 
 2
 ba 3 ab3 a 2 b 2   −1  2

 6 + 3 + 2   ab + b 2  = C1 −  
EI  3L 

Force 3L E I (4.16)
The deflection equation y A ( x) can now be evaluated for = 2 ≡ K Simply Supported
x=a Displacement a ( L − a )2 Beam Loaded at x = a

− F RA  x 3  ba 3 ab3 a 2 b 2   −1  
yA ( x) = + + +   x
E I  6  6 3 2   ab + b 2   4.2.2.6  Non-Linear Springs
The relationship between force and elastic deformation may
for 0 ≤ x ≤ a be non-linear, either because the stress–strain properties
of the material are non-linear (e.g., aluminum), or because
the object may deform in a way that alters the stress state
− ( − Fb )  a 3  ba 3 ab3 a 2 b 2   −1   in the object. Consequently, the greatest source of error in
yA (a) = + + +  a
E I L  6  6 3 2   ab + b 2   the linear model of a spring is often the need to approximate
non-linear behavior with a single straight line described by a
4.2  Modeling Translational Elements 205

6
a x,v
K
5 a b c
F(t) 1
Uniaxial g
4
Nominal
a b c
Stress 3 b F(t) 1
g
MPa Biaxial
2 ∆x1g
∆a
1 ∆b
0 ∆c
1 2 3 4 5 6 7 8
Stretch Fig. 4.7   Linear variation in deformation with position along a transla-
tional spring with constant stiffness per unit length. a Relaxed spring.
Fig. 4.6   Stress vs. stretch data for rubber under uniaxial and biaxial b Compressed spring
load. Note the unit of stretch rather than strain due to the large defor-
mations. (The data are from Trealor and reproduced from Boyce and x,v
Arruda 2000)
K
a b c
F(t) 1
g
spring constant or spring rate K and not the variation in the vcg
performance of nominally identical springs. vag vbg
v1g
Non-linear springs are often desirable in machine design.
We may wish to have a spring which is soft with a small L
spring constant K for small force and becomes stiffer with Fig. 4.8   Linear variation in velocity with position along spring K dur-
a larger K for larger forces. Elastomeric (rubbery) materials ing deformation
have this property, Fig. 4.6. Elastomers are used as “vibra-
tion isolators” to dissipate vibration and as “shock mounts” John William Strutt (1842–1919), made significant contribu-
to protect components from impulse loads. An example of tions, and common methods bear his name.
the use of elastomers added to design for their vibration An assumption of vibration theory is that the maximum
damping characteristics are motor mounts in automobiles. strain energy stored in an element equals the maximum ki-
Although non-linear springs are useful in mechanical de- netic energy. In other words, there is no energy dissipation in
sign, they are problematical for the analysis. Superposition the system. The notation used in physics and vibrations is U
requires linear differential system equations. A linear differ- for potential energy and T for kinetic energy. The deformed
ential system equation, in turn, requires a linear constitutive shape of the object must be established. Three approaches
or elemental equation for every element in the system. are used. The deformed shape of the object can be calculated
as given in Sects. 4.2.2.3 and 4.2.2.4 using beam theory. Al-
ternatively, for objects with constant cross-section and prop-
4.2.3  Effective Mass erties, such as beams, a sinusoidal shape which satisfies the
end conditions is assumed.
If a machine component stores a significant amount of both
kinetic and elastic energy, then its energetic model must in- 4.2.3.1  Effective Mass of a Spring
clude both a spring and a mass. The spring constant K is When an elastic object undergoes deformation, the displace-
calculated as the ratio of force over displacement, Eq. 4.2, ment of points on the object must vary with position. If all
as developed in Sect. 4.2.2. The corresponding lumped points on an object displace equally, then the object trans-
parameter mass is the “effective” mass of the element. A lates without deformation. In the special case of a machine
system with two independent energy storage modes will component with constant stiffness along its length, the elas-
have internal energy transfers during its natural or unforced tic deformation varies linearly from one end to the other, as
response. The transfer of energy between independent en- shown with spring K in Fig. 4.7. The elastic energy stored is
ergy storage modes is manifested as an oscillation in the uniform along the length of spring K since it is stored in elas-
system’s power variables, which, in a mechanical system, tic strain. If the machine component, say spring K, also stores
is termed a vibration. Accordingly, calculation of effective a significant amount of kinetic energy during the elastic de-
mass uses theory developed in the mechanical engineering formation, the velocity of the mass elements which comprise
discipline of vibrations. The study of vibrations and waves the component will also vary linearly, Eq. 4.17, Fig. 4.8.
are topics of Newtonian mechanics and, as such, predate the
study of system dynamics by centuries. Lord Rayleigh, a.k.a.   x
v( x) = 1 −  v1g (4.17)
 L
206 4  Mechanical Systems

Cross-sectional area A 4.2.3.2  Effective Mass of a Cantilevered Beam


Area-moment of inertia I
Young’s modulus E
If a beam acting as a spring also stores a significant amount
y,v Density ρ of kinetic energy during its deflection, then the lumped pa-
F
x dm
rameter model of the beam must include the effective mass
of the beam. We will use a cantilevered beam with a constant
g 1 cross-section and properties for this example, Fig. 4.9.
L We derived the deflected shape of a cantilevered beam
with uniform cross-section loaded at its end, Eq. 4.12, and
Fig. 4.9   End-loaded cantilever beam with infinitesimal mass element the spring constant K, Eq. 4.13, in Sect. 4.2.2.4
dm at x
 F0  Lx 2 x 3  (4.12)
y ( x) = − − 
where L is the instantaneous length of the spring. E I  2 6
Consequently, kinetic energy is not stored uniformly in
spring K. The effective mass is calculated in terms of the  3E I
K Cantilevered Beam = (4.13)
kinetic energy stored in the spring calculated using the maxi- Loaded at End
L3
mum velocity, v1g, Eq. 4.18:
If the external force F is suddenly removed, the cantilevered
 1 beam will vibrate or oscillate, as anyone who has bent and
E kinetic = M effective v12g (4.18)
2 released a ruler cantilevered off the edge of a table can at-
test. The vibration occurs due to the transfers in strain energy
The effective mass of spring K is the inertia “felt” at node from bending into kinetic energy and back. When a canti-
1. Since v1g is the maximum velocity of any mass element levered beam is loaded and then released, its oscillation is
in the spring, the effective mass must be less than the total a “free,” as opposed to a forced, vibration. Free vibrations
mass. If the mass of the spring is Mspring, then a mass element are mechanical oscillations described by the homogeneous,
dm of the spring at any instant during deformation is: unforced, or natural response of the mechanical system.
The deflected shape of a freely vibrating beam is the su-
 M spring perposition of “modes” of a number (infinite, in the abstract)
dm = dx (4.19)
L of possible deflected shapes, all of which are sinusoidal. The
primary or first mode has the lowest frequency of vibration,
The effective mass is calculated by scaling the contribution known as the fundamental frequency, and the simplest shape,
of a mass element dm by the square of the ratio of the veloc- with the fewest sinusoidal cycles or fractions of cycles.
ity at that location to the maximum velocity, Eq. 4.19 The calculation of the effective mass requires assuming
a sinusoidal mode shape. For an assumed mode shape to be
2
1 M spring  
L
x  reasonable, it must satisfy the boundary conditions of a can-
2 ∫0 L   L  
E kinetic = 1 −  v1g dx tilevered beam. Where the beam is built-in to a support, the
deflection, y( x), and the slope of the beam, dy( x)/dx equal
1 M spring   L − x  
L 2 zero.
2 L ∫0   L  
E kinetic =   v1g dx
dy ( 0)
y ( 0) = 0 =0
1 M spring 2
L dx
v1g ∫ ( L − x ) dx
2
E kinetic = 3
2 L 0
The boundary conditions for the free end are the absence of
1 M spring 2 2
L

E kinetic = v1g ∫ L − 2 Lx + x 2 dx shear and bending ­moment. Recall that moment is related to
2 L3 0 curvature, which for small deflections can be written as
L
1 M spring 2  2 x3  d 2 y ( x) M ( x) d 2 y ( L)
E kinetic = 3
v1g  L x − Lx 2 +  = → =0
2 L  3 0
dx 2
EI dx 2
1 M spring 2 L 3
E kinetic = v1g Also recall that shear, V( x), integrates to yield moment, M( x),
2 L3 3 and, conversely, moment is differentiated to yield shear
1 M spring 2
E kinetic = v1g
2 3 d 3 y ( x) V ( x) d 3 y ( x)
= → =0
dx 3 EI dx 3
M spring
 M spring =
effective 3 (4.20)
4.2  Modeling Translational Elements 207

0.0 The contribution of the mass element dm to the kinetic en-


Mode 1 shape,
Eq. 4.21 ergy of the beam is:
-0.2
  dy ( x ) 
2
1 1
d E M ( x ) = dm v ( x ) d E M ( x ) = dm 
2
y(x)
___
-0.4 →
2 2  dt 
y(L)
-0.6 (4.23)
End load bending
displacement, Eq. 4.12
-0.8
The uniform cross-section and properties of the beam allow
the mass element dm to be expressed as the mass per unit
-1.0
0.0 0.2 0.4 0.6 0.8 1.0 length of the beam
x
_
L
 M beam
dm = Aρ = (4.24)
Fig. 4.10   Normalized plots of the deflected shape of a cantilevered L
beam due to an end load, Eq. 4.12, and the first mode shape of a vibrat-
ing cantilever, Eq. 4.25
The kinetic energy of the mass element at distance x from
the support is:
The first mode shape of a cantilevered beam is quarter cycle
1 M beam  dy ( x ) 
2
of a cosine, (4.25)
d EM ( x) =
2 L  dt 
  πx 
y1 ( x ) = ymax 1 − cos   
(4.21)
  2L  
We assume that the velocity dy( x)/dt varies along the beam
velocity with the same function as displacement, as we did to
where ymax = y( L). calculate the effective mass of a spring, with the maximum
The deflected shape calculated from beam theory, used to velocity at the end, vmax = v( L) ≡ v1g .
determine the elastic strain energy of the beam, and the de-
flected shape assumed by vibration theory to calculate the ki-    π x   π x
v ( x ) = vmax 1 − cos    → v ( x ) = v1g 1 − cos   
netic energy of the beam differ slightly as per Fig. 4.10. The   2L     2L  
deflection in the middle of the cantilever is slightly greater (4.26)
under the end load than in first mode shape. If we were to as-
sume that the beam oscillated freely in its end-loaded shape, Substituting into the expression for the kinetic energy of a
we would calculate a larger effective mass than that calcu- mass element
lated assuming that beam vibrates in the first mode shape.
2
The first mode stores most of the mechanical energy of a vi- 1 M beam    π x
bration. The “harmonics,” or frequencies of the higher modes, d EM ( x) =  v1g 1 − cos  2 L   
2 L
are integer multiples of the fundamental frequency. The higher
mode shapes contain additional sinusoidal cycles which pass
through “nodes” with zero deflection. This type of oscillation Integrating yields the kinetic energy of the beam
is called a “standing” wave, because the sinusoidal oscillation
2
does not progress down the beam. The amplitude of the oscil- L L
1 M beam    π x
lation varies cyclically, but the locations of the maxima and ∫0 d E M ( x ) = ∫0 2 L  v1g 1 − cos  2 L    dx
the nodes of zero deflection do not change location.
Since the end of the cantilevered beam at x = L is not load- L 2
ed in a free vibration, the moment and shear at x = L are zero: 1 M beam 2   π x
EM = v1g ∫ 1 − cos    dx
2 L 0
  2L  
d 2 y ( L) M ( L) d 3 y ( L) V ( L)
2
= =0 3
= =0
dx EI dx EI 1 M beam 2
L
 π x  π x
2

EM = v1g ∫ 1 − 2 cos   + cos   dx


2 L  2 L   2L 
The velocity, v( x), of a mass element of the beam, dm, at dis- 0

tance x along the beam is the time derivative of the deflection


L
y( x) at that location 1 M beam 2  4L  π x x L  π x 
EM = v1g  x − sin   + + sin   
2 L  π  2 L  2 2π  L 
 dy ( x ) 0
v ( x) = (4.22)
dt
208 4  Mechanical Systems

4
1 M beam 2  3L 4 L  π L L  π L  Fb, kN
EM = v1g  − sin   + sin    3
2 L  2 π  2L  2π  L 
1 0
2 b
1 M beam 2  3 L π 2 4 L  1
EM = v1g  −
2 L  2 π 2 π  -5 -4 -3 -2 -1 1 2 3 4 5
-1 m
v12 , ______
1  3π − 8  sec
EM = M beam v12g  -2
2  2π 
-3

-4
1
EM = M beam v12g 0.227
2 Fig. 4.11   Linear translational damping. The slope of the line is the
damping coefficient b
We now equate the kinetic energy of the beam to the equa-
tion for kinetic energy written in terms of the effective mass
of the beam and the velocity of the end at x = L  Fb = bv1g (3.11)

1 1 Different symbols are used in different subdisciplines to rep-


M beam v12g = E M = M beam v12g 0.227
2 effective 2 resent the same physical quantity. In the area of mechanical
vibrations, the symbol “c” represents a linear viscous damp-
 M beam = 0.227 M beam (4.27) ing coefficient b. Likewise, linear viscous damping is known
effective
as “proportional damping” in vibrations. In viscoelasticity,
Sect.  4.4.6, η is the symbol of a linear viscous damping
coefficient.
4.2.4 Damper: Viscous Friction Energy
Dissipation 4.2.4.1 Linearization
We now investigate how to create a linearized model of the
Energy is dissipated in a mechanical system by friction as typically non-linear relationship between damper force, Fb,
heat. Friction is the result of shear, either the shear between and the velocity difference across the damper. The behav-
two solid surfaces, Coulomb friction, the shear of a fluid, ior of all real systems is non-linear to some degree. Recall
or the plastic deformation of a solid. The frictional proper- the simple test for linearity from Sect. 2.3, “Does doubling
ties of mechanical systems generally introduce the great- the input, double the output?” Eq. 3.11 is a linear relation-
est uncertainty and error in energetic models. The type of ship. Doubling the difference in velocity between the piston
friction that can be reasonably accurately represented in a and cylinder, v12, doubles the force, Fb, acting through the
dynamic model is “viscous” friction, which was modeled in damper.
Sect. 3.2.1. Viscous friction describes the relationship be- The elemental equation
tween shear force and translational shear velocity created by 
a thin lubricant film between solid surfaces. Under the proper Fb = b1v12 (4.28)
combination of geometry, lubricant, and relative velocity be-
tween the two surfaces, the shear force required to move one is linear. However,
surface parallel to the other is approximately proportional to 
the shear rate, where the shear rate is equal to the difference Fb = b1v12 + b2 v12 v12 + b3 v123 (4.29)
in velocity of the two surfaces divided by the distance be-
tween them. “Ideal” viscous friction has a linear relationship Fb = b1v12 + b2 v12 v12
(4.30)
between shear force and shear rate. We apply the model of
ideal viscous friction to dashpots and other damping devices, Fb = b2 v12 v12
(4.31)
in which energy is dissipated in the flow of a viscous fluid
through orifices, tubes, and filters, by formulating a linear re- are non-linear relationships; doubling v12 does not double Fb.
lationship between the power variables associated the shear Note that the squared term is expressed as the product of v12
or pumping a fluid. In a translational mechanical system, the and its absolute value, rather than as v122 to preserve the effect
power dissipated in ideal viscous friction is the product of of the sign of the velocity on the resulting force.
the force acting through the damper and the velocity differ- If the operating range of v12 is centered on zero, then we
ence or drop across the damper, Eq. 3.11 and Fig. 4.11. “linearize,” or approximate with a linear equation, a relation-
4.2  Modeling Translational Elements 209

Fig. 4.12  a Non-linear viscous


friction. b Linearized models of a Damper
b Damper
viscous friction Force Fb Force Fb b II

bI

Velocity, v12 Velocity, v12

Linear Viscous
Friction Models

1,000 which may or may not be centered in the middle of the oper-
3
Fb(v12 ) = 10 v12+ |v12 |v12+ v12
ating range. Rather than using a formal Taylor series where
500 the coefficient of the second term is the derivative of the func-
Fb, N tion at the operating point, and the straight line is tangent to
Damper 0 the curve at the operating point, we may choose to fit a line
Force Fb(v12) ≈ 37 v12 to the curve at the operating point, by minimizing either the
-500 absolute value or the square of the error. In any case, the lin-
earization is the sum of the value of the function at the operat-
-1,000
ing point, plus a term which is the product of a coefficient and
-10 -5 0 5 10
the difference between the value v12 and the operating point
m
v12 Velocity Difference, ___
sec  dF
Fb = Fb vop . pt . (
+ v12 − vop. pt . ) dvb
(4.32)
Fig. 4.13   Linearization of non-linear translational damping 12 vop . pt .

For example, if the non-linear damping relationship


ship of the form, Eq. 4.17, by truncating the squared and
cubed terms and then adjusting magnitude of the coefficient N ·sec
Fb = 50 v12 v12
b1 to best fit the curve over the operating range, Fig. 4.12. m
The criterion for “best” is generally left to the judgment of
the engineer. Fitting a straight line to the curve by minimiz- was to be linearized for the operating point v12 = 2 m/sec and
ing the sum of either the square or the absolute value of the the operating range from v12 = 1m/sec to v12 = 3 m/sec then
error between the non-linear and linear relationships is a we create the two-term Taylor series expansion.
common method. For example, approximating
N ⋅ sec 2 m m
Fb = 50 2 2
N ⋅ sec N ⋅ sec 2 N ⋅ sec3 3 m2 sec sec
Fb = 10 v12 + 10 v12 v12 + 1 v12
m m m  m  N ⋅ sec 2  m
+  v12 − 2
  ( 2 )  50 m 2  2 sec
sec 
as
N ·sec  m N ⋅ sec 
Fb ≈ 37 v12  (4.33)
m Fb = 200 N +  v12 − 2   200 
 sec m 

by a linear relationship to fit the operating range


−6 m/sec < v12 < 6 m/sec with the minimum sum of the 4.2.4.2  Shock Absorbers and Snubbers
error squared requires increasing the coefficient of the first- “Shock absorbers” and “snubbers” (large shock absorbers)
order term from 10 to 37; whereas, the fit, which minimized are dampers used to dissipate the energy of disturbance forc-
the sum of the absolute value of the error, yields the coef- es and motions. They have a piston-cylinder design and are
ficient 33, Fig. 4.13. the basis for the symbol of a dashpot. There are a variety
If the operating range of the system is not centered on of designs which differ in how the working fluid, oil, flows
zero, then we linearize non-linear relationships with a two- from one side of the piston to the other under load. The most
term Taylor series expansion at the expected operating point, common example are the shock absorbers used as dampers
210 4  Mechanical Systems

900 “pressure tube” into the outer “reserve tube.” The two valves
Fb(v12 ) = 50 |v12|v12 have multiple valve “stages,” which are disks with orifices
500 that become exposed and pass flow at different pressures.
The disk stack presses against a compression spring which
F b, N
controls the valve opening.
Damper 0 The force–velocity difference traces of Fig. 4.15 are
Force
clearly non-linear. A less obviously non-linear relationship is
-500 dFb(2)
( m
___
Fb(v12 ) ≈ Fb(2)+ ________ v12 - 2 sec
dv12
)  Fb = bv12 + Fb0 where Fb0 is constant (4.34)

-900 Although it is the equation for a straight line with familiar


-4 -2 0 2 4
m form, y = mx + b, the line does not pass through the origin.
v12 Velocity Difference, ___ It is “offset” vertically by the additive constant. Doubling v12
sec
does not double Fb. A straight line that does not pass through
Fig. 4.14   Linearization of non-linear translational damping about a the origin does not describe a linear model. It is an “incre-
non-zero operating point mentally linear” model. Compare the plot of Fig. 4.16a and b.
The linear model is described by Fb = bv12, in which the
500 force is zero at zero velocity. The incrementally linear model
is described by Fb = bv12 + Fb . The offset Fb is the non-zero
0 0

force at zero velocity. The test “does double the input yield
250 double the output” fails. Superposition cannot be applied.
Fb Any functional relationship which does not pass through the
origin is non-linear.
lbf
0 Linear ordinary differential equations with constant co-
efficients result from using linear elemental equations with
constant elemental parameters to model an energetic system.
Linear elemental equations are often only poor approxima-
-250 tions of the behavior of components in real systems. Common
-10 -5 0 5 10
in machine elements with significantly non-linear behavior are
v12 , ___
sec energy dissipaters, where Coulomb friction is particularly
troublesome, and transformers and transducers are subject
Fig. 4.15   Shock absorber force–velocity difference curves in US cus- to saturation and power sources with power limits. Unfor-
tomary units tunately, the only type of differential equation of practical
importance we can solve analytically is a linear ordinary dif-
in vehicle suspensions. A “shock mount” has no working ferential equation with constant coefficients. Consequently,
fluid. It is an elastomer, a rubbery polymer or synthetic rub- we must use linear elemental equations, even when we know
ber. The elastomeric motor mounts between the engine of they are only poor approximation, because our design tech-
a vehicle and its frame or subframe to reduce vibration are niques require linear differential equations.
shock mounts. When we need to improve the accuracy of our math-
A shock absorber or snubber with a fixed orifice diam- ematical model due to the significance of the system’s non-
eter has a force–velocity relationship similar to the trace in linear behavior, we use numerical solutions, such as the
Fig.  4.14, where the damper becomes stiffer with increas- Runge–Kutta method, which allow us to include non-linear
ing velocity difference between the piston and cylinder. This elements. We will introduce numerical method in Chap. 7.
response is unacceptable for a vehicle suspension. Vehicle However, even when we use a numerical solution, we will
shock absorbers are designed to produce either an approxi- still start our analysis with a linear approximation. Ironically,
mately linear or a “digressive” force–velocity difference even when a linearized model is too inaccurate to yield use-
curve, where digressive refers to a curve which becomes soft- ful results, it still has one useful property: a linear model has
er (flatter) with increasing velocity difference, per Fig. 4.15. only one solution. Non-linear models can have a number of
The force–velocity difference relationship is created by the solutions. If we use the inaccurate results of our linear model
valve design. Conventional automotive shock absorbers as a starting point to begin the iteration required to fit a non-
have two valves, one in the piston, which controls flow of oil linear model to our data, we save time and are more likely to
from one side of the piston to the other, and a second valve at converge to a physically meaningful solution.
the base of the cylinder, which controls flow from the inner
4.2  Modeling Translational Elements 211

Fig. 4.16  a Linear model.


b Incrementally linear model.
a Damper b Damper
Force Fb Force Fb
The equation of a straight line
which does not intersect the
origin is “incrementally linear”

Velocity, v12 Velocity, v12

Line passes Line does not pass


through the origin through the origin

Fig. 4.17  a Static and kinetic


friction coefficients model of
a b Average Actual
Coulomb friction. b Typical Kinetic Friction Friction Friction Force Friction
average actual friction force Force Force
plotted with the friction coef-
ficients model. The friction force
has the opposite sign of the force
acting through the damper due to Static Friction
inverse sign conventions

Velocity Velocity

Consider the static and kinetic model of Coulomb fric- points), rather than one straight line. We must simplify the
tion, described in terms of coefficients of friction. The model model to use it analytically. The first thing we will do is ne-
consists of two horizontal lines for v ≠ 0 and two points for glect the case of static friction, since we are developing a
v = 0, Fig.  4.17a. It yields four different values of friction, dynamic model. If a mechanical system isn’t moving, then
two for the static case and two for the kinetic case, which it isn’t dynamic. The response either hasn’t started yet, or it
depend on the direction of motion, since the frictional force has reached steady-state. Further, it is difficult to implement
opposes motion. Typical measured Coulomb friction is more mathematically. How is our model to know which way to
difficult to describe, Fig. 4.17b. apply a static force to oppose motion which hasn’t happened
In the case of Coulomb friction, the conditions of the sur- yet? We can model kinetic friction, if we know the direction
faces in contact are subject to change. Surface contaminants, of motion, by using a force source to apply a constant force
i.e., dirt and wear particles, and chemical and mechanical al- to represent the kinetic friction resisting motion. This meth-
teration of a surface change its frictional properties. To start od works well for first-order dynamic systems, since they
with, there are many different types of friction. The most cannot oscillate, and we will know the direction of the mo-
familiar is dry friction between two surfaces. A very use- tion. The only trouble is that we cannot turn off the kinetic
ful model consists of static and kinetic friction coefficients, friction when the velocity reaches zero. We must keep this in
which are presented as constants. Although this model is mind when interpreting the results of our model.
very useful, it isn’t very realistic. Shouldn’t the relationship
between the normal and tangential force acting on surfaces
depend on the displacement of the surfaces, not to mention 4.2.5  Translational Mechanical Sources
the wear, temperature, lubrication, cleanliness, and chemi-
cal alteration of the surfaces? Friction depends on all these Although some energetic mechanical systems operate from
factors. Further, there is a stochastic, or probabilistic, aspect energy stored in springs or flywheel over the period we wish
evident in the transition from kinetic to static friction; or, in to model them, it is more common that the system draws en-
other words, when a sliding object stops. ergy from a power supply. Mechanical power is the product
Although the static and kinetic Coulomb friction model of force and velocity, P = Fv. Each element in a linear graph
is conceptually simple, it is not mathematically simple, represents a power flow in the energetic system, where the
because it is non-linear and consists of two lines (and two power sourced, stored, or dissipated by the element equals
212 4  Mechanical Systems

a b 1 or velocity, as the system draws power in response to the


change of the input. Force sources must be powerful enough
x,v to maintain the specified force over the range of the velocity
b of the point of application of the force. Likewise, velocity
F(t) F(t) b M sources must be powerful enough to maintain the specified
M velocity, by applying the force needed to move the point of
application of the force at the specified velocity. The more
g powerful the device, the more nearly perfectly it can behave
as a source. On earth, gravity is a nearly perfect source of
Fig. 4.18   Schematic and linear graph of a force source-mass-damper force acting on a mass, since the gravitational force is virtu-
system. The power flow in each element is the product of the through
variable, force, and the across variable, velocity
ally independent of the velocity of the mass, and the change
is the distance of the mass from the center of the earth.
Some devices can be either a force source or a velocity
the product of the force acting through the element and the source, depending on how they are controlled. For example,
velocity difference across the element. laboratory tensile testing machines typically can be placed
A power source is identified by which of its two power vari- under either force or velocity control. A powerful hydraulic
ables of the system is either known or controlled. Mechanical piston under feedback control can function as either a force
systems have force sources and velocity sources. If a source is source or a velocity source. Again, if we control the force
powerful enough, it is possible to control (or specify) the time applied by the source, we cannot control the velocity of the
history of one, but not both, of its two power variables. The point of application, unless we dictate the laws of physics.
magnitude of the second power variable is determined by the Similarly, a velocity source for a device provides power to
response of the system. For example, if power is supplied to a mechanical system with a known or prescribed velocity. If
a system at a specified force, then the response of the system we control the velocity of the point of application of a veloci-
determines the velocity of application of that force. ty source, we cannot control the force that the source applies.
An abstract force source, Fig. 4.18, is the mechanical Force sources, illustrated as force vectors, are the typical
power source most familiar to engineering students, since mechanical source used in introductory courses. However,
they have seen many vectors labeled F or F( t) on free body in machine design, energy is often provided by a “velocity”
diagrams. The convention of representing power sources as source. A velocity source is a mechanical element that follows
vectors indicates that whatever is providing power to the a specified motion at a specified velocity. Whatever is mov-
system is powerful enough that it can maintain the specified ing the machine element must be powerful enough to provide
value of the input variable. In order to apply the specified the force necessary to move the point of application of the
force to the system, the point of application of the force must source at the specified velocity. Velocity sources are more
move with the velocity of the system. If the force source familiar than they seem at first. We more commonly set or
were to lose contact, then it could not apply the force. Con- control the speed of a machine than we do the force it applies.
sequently, a force source must also be capable of displace- Power limitation determines what “type” of source can
ment at the velocity of the system it is driving. A mechanical act on what type of element. We will classify both sources
power source cannot float in space and apply force to an and energy storage elements by one of the power variables,
object, as the input force vector is portrayed in Fig. 4.18. A force or velocity in mechanical systems. We can control only
mechanical source must react against something, since for one of the two power variables of a power source. As stated
action there must be reaction, per Fig. 4.19a. If a force vector above, a force source is a device which provides power to a
is shown floating in space, then the necessary reaction force mechanical system with a known or prescribed force.
is assumed to be provided by ground, per Fig. 4.19b.
In order to model a device as a power source, it must be 4.2.5.1  Limitations of Mechanical Sources
able to supply enough power such that it can approximately The limitations of mechanical sources are imposed by the mag-
maintain the specified magnitude of the input variable, force nitude of the power drawn by the system. Ideal power sources

Fig. 4.19  a Force source mass-


damper system shown with
a x,v b x,v
support to provide the reaction
needed by the mechanical power b b
source. b Conventional represen- F(t) 1 F(t) 1
tation of the power source as a g M g g M g
vector with the reaction against
ground implied
4.2  Modeling Translational Elements 213

a b 1 The corresponding limitation applies to strain energy. We


cannot apply a step change in force to a spring with a force
x,v source, since it too would require infinite power. From the
b perspective of power and energy,
v(t) v(t) b M
M ΔE K ΔE K ΔE K
lim = ≈ → P =∞
Δt → 0 Δt dt 0
g
From the Newtonian perspective, it is impossible to apply
Fig. 4.20   Schematic and linear graph of a velocity source-mass- a Heaviside step input of force directly to a spring, since
damper system. It is impossible to command a step change in velocity
the point of application of the force must displace to create
because that would require infinite power from the source
the spring force, Fig. 4.21. Instantaneous deformation of a
spring requires infinite velocity, which is impossible.
maintain the specified value of the input variable regardless The restriction that a force source cannot impose a force
of the power drawn by the system. However, for an energetic on a spring runs contrary to many models you have used in
model to be useful for all possible inputs which might be ap- statics and machine design. A force represented as a vector
plied to the system, a power source cannot directly drive an acting directly on a spring is a common model in physics
energy storage element with the state (energy storage) variable and engineering texts. What is unstated in those models is
identical to the source. that either (1) the system has reached steady-state, and the
The two mechanical system energy storage elements, a force was not constant during the transient period, or (2)
mass and a spring, store kinetic energy and strain energy, in the input function increased the applied force gradually, so
terms of velocity and force, respectively that the power drawn from the source by the system did
not exceed the capacity of the source. We will see when
2
1 2 1 1 F  F2 we investigate frequency response in Chap. 10 that a force
Em = mv1g EK = Kx12 2 = K  K  = K
2 2 2  K 2K source can act directly on a spring, because a sinusoidal
input varies gradually enough to limit the power demanded
If we control a source, then we must be able to command it from the source. Force inputs which start at zero then ramp
to apply any arbitrary input to the system. A Heaviside step up or down can also be applied directly to springs. Text-
input is an extreme input, since it instantaneously chang- book figures showing force sources acting on springs imply
es values. We cannot apply a step change in velocity to a one of these restrictions or special cases, because in the
mass with a velocity source, Fig. 4.20. A finite change in general case, when there are no restrictions on the type of
the amount of energy stored in the mass over the infinitesi- input function, a force source cannot arbitrarily control the
mal time period from t = 0 − to t = 0+ would require infinite amount of energy stored in a spring.
power from the source Violation of the restriction that a force source cannot im-
pose a force directly on a spring is more difficult to detect,
ΔE M ΔE M ΔE M because we need to look beyond the element which is the
lim = ≈ → P =∞
Δt → 0 Δt dt 0 point of application. For example, a force source acting on
a damper seems reasonable, until it is recognized that the
From the Newtonian perspective, a finite change in veloc- damper is in series spring, and the two elements carry the
ity over infinitesimal time is an infinite acceleration, which same force.
would require an infinite force to impose on a mass.

Fig. 4.21  a Schematic and


a b
b linear graph of a force source K
spring-damper system. It is 1 2
impossible to command a step x,v
change in velocity because that b
would require infinite power
F(t) K
from the source
F(t) b

g
214 4  Mechanical Systems

Fig. 4.22  a Schematic and


b linear graph of a force source a b b
damper-spring system. Although 1 2
the force source F( t) acts on x,v
damper b, it also imposes its b
force on spring K. It is impos- F(t) K
sible to apply a Heaviside step K
input to this system
F(t)

Fig. 4.23  a Schematic and b


linear graph of a force source a b b
damper-spring system with a 1 2
mass element added at the point
of application of the force source. x,v
The force applied by the source
divides at node 1 between accel-
b
erating mass M and compressing F(t) K F(t) M K
damper b in series with spring K.
M
At the instant after a step change
in force, all of the step change
acts to accelerate mass M g

The concern of demanding infinite power from a source in order to use linear models. If it is the case that the source
commanded to undergo the instantaneous transition of a can only push and not pull, or only move forward but not in
Heaviside step may seem artificial, since no physically real reverse, then we are limited to modeling only part of the re-
power source can instantaneously change the magnitude sponse of the system without resorting to numerical solutions.
of its output variable. The flow of energy into or out of the
energy storage element of a first-order system is governed
by the homogeneous response. The difficulty we encounter 4.2.6 Summary
using a model in which we attempt to dictate the value of an
energy storage (state) variable is that the system cannot have It is important to keep clearly in mind that a schematic is a
a transient response. This is why models of that sort are only model of an energetic system. The energetic properties repre-
in steady-state. sented by the elements in the schematic are not all of the prop-
If we were modeling a physical system with the dominate erties of the system. They are the dominant energetic prop-
energetic properties of strain energy storage and energy dis- erties that must be included, in order to model the behavior
sipation; and the input to the system was better modeled as of the system with the precision needed for the particular en-
a force source than a velocity source, we would modify the gineering analysis being performed. You will discover as we
model shown in Fig. 4.22 by adding a mass element at the work with dynamic models that what is significant and must
point of application of the source, per Fig. 4.23. The amount be included in a model for it to be reasonably representative of
of energy stored in the mass element is small enough, so that the system depends on many factors, including the magnitude
we would usually neglect it. However, in this case, it serves and type of input and the time scale of the response.
to limit the power drawn from the source. A step change in
force applied to mass from the source results in a step change
in the acceleration of the force, not a step change in the de- 4.3  The Sign Problem of Mechanical Systems
formation of the spring, thereby limiting the power flow
from the source. We need to consider signs. There are two incompatible defi-
A significant problem for modeling energetic systems is nitions of a “positive” force in mechanical engineering. The
that linear systems must have a source which both provides convention we choose to use depends on what is important
(source) and accepts (sink) power, while maintaining the spec- to us in the particular analysis. The convention which arose
ified value of one of its two power variables. There are rela- from the mechanics of rigid bodies defines a force as posi-
tively few sources which can push and pull with the same force tive based on the direction of force in space. A positive force
or velocity. We must ignore this limitation of the hardware produces a positive acceleration on a mass, that is, an accel-
4.3  The Sign Problem of Mechanical Systems 215

Fig. 4.24  a Force in the positive


direction puts the spring into
a b
x,v x,v
tension. b Force in the positive
direction puts the spring into
compression K F(t) F(t) K
M M

Frictionless rollers Frictionless rollers

x 1, v 1 x 2 , v2 to the crux of the sign problem. It looks wrong and, indeed,


FK K FK if equal forces acted in the same direction on both ends of a
translational spring, the spring would not deform.
Node 1 Node 2 In a conventional Newtonian analysis, the conflict be-
tween the sign conventions based on the direction of the
Fig. 4.25   Translational spring showing both displacement and veloc-
force, used for mass, and the sign conventions based on the
ity variables at the nodes. The force FK is shown in the positive direc-
tion at either end of the spring effect of the force, used for stress, strain, and springs, is re-
solved by drawing free body diagrams of every mechanical
element in the system. This is a lot of work and, unfortunate-
a b ly, error-prone due to confusing action and reaction forces.
K K The linear graph sign conventions limit the effort and
1 2 1 2 possibility of error, by arbitrarily defining a positive di-
rection for the force in acting through each element in the
Fig. 4.26  a Linear graph spring element oriented by the arrowhead network, by drawing an arrowhead near the middle of the
with the positive direction from node 1 and node 2. b Spring element line connecting the two nodes. This process is known as
oriented from node 2 to node 1 “orienting” the element. The direction of the arrowhead
is arbitrary, but once chosen, it defines the positive direc-
eration in the positive direction as defined by the axes. The tion for the force acting through the element and, therefore,
convention which arose from solid or continuum mechanics the order of the nodes for calculating the positive velocity
defines a positive force based upon its effect on materials difference across the element. The direction a force acts
or machine elements. A force which produces tension on an on a node will be interpreted using a flow analogy and the
infinitesimal element is defined as positive. “bank account” sign convention. Force directed into a node
As it happens, the positive definition of force from con- is positive. Force directed out of a node is negative. The
tinuum mechanics corresponds with the effect of the forces arbitrary positive directions of elements work out the same
on the design of machine and structural components. The way that arbitrary signs assumed in a truss analysis are cor-
common meaning of the word, “positive,” is good or favor- rected by the algebra.
able. For example, we welcome a positive review of our For example, the positive direction for the spring in
performance. Tensile force is positive, in the sense of being Fig.  4.26a is from node 1 to node 2, yielding the element
favorable, because engineers generally prefer to have ele- equation
ments loaded in tension rather than compression. Compres-
sion can lead to buckling, a sudden and catastrophic failure dFK
= Kv12
mode. Tension cannot produce buckling. Elements designed dt
to carry a compressive are more complicated and more ex-
pensive than elements that carry a tensile load of the same Conversely, the positive direction for the spring in Fig. 4.26b
magnitude. Hence, compression is bad, or at least unwanted, is defined by the arrowhead as from node 2 to node 1. Con-
and negative. sequently, the element equation is
Consider the two similar spring-mass systems in Fig. 4.24. dFK
They are both acted on by a force F( t) in the positive direc- = Kv21
dt
tion. In Fig. 4.24a, positive displacement of mass M places
spring K in tension whereas, in Fig. 4.24b, positive displace- Recall that reversing the order of the node subscripts on ve-
ment of mass M places spring K in compression. locity differences is equivalent to inverting the sign,
Isolating a translational spring and showing the equal
−v12 = v21
forces acting on it in the positive direction, Fig. 4.25, gets
216 4  Mechanical Systems

Fig. 4.27  a Linear graph of a


spring-mass-damper system with
a b b b
the positive direction damper 1 2 1 2
b and spring K defined from
nodes 1 to 2 and from nodes 2
to g. b Linear of the system of
4.27a, but with the positive di- K
rections of damper b and spring
F(t) M F(t) M K
K reversed

g g

We will demonstrate that reversing the arbitrarily defined


b dF (t )
2
dv1g bM d v1g
positive directions for elements does not affect the resulting + F (t ) = M + bv1g +
differential system equation, by reducing the linear graphs in K dt dt K dt 2
Fig. 4.27a and b for the velocity of the mass, v1g. The linear
graphs have opposite directions defined as positive for damp- 1 dF (t ) K d 2 v1g K dv1g K
+ F (t ) = + + v1g
er b and spring K. The sign reversals of the through variables, M dt bM dt 2 b dt M
forces Fb and FK, are included in the continuity and element
equations. The sign reversals of the velocity differences, v12 The equations for the linear graph in Fig. 4.27b, in which the
and v2g, appear in the compatibility and element equations. positive direction for damper b and spring K are the reverse
The continuity, compatibility, and element equations of of the linear graph in Fig. 4.27a are
the linear graph in Fig. 4.27a are
Continuity: F (t ) + Fb = FM Fb = FK
Continuity: F (t ) = FM + Fb Fb = FK
Compatibility: v1g = −v21 − vg 2 v1g = v1g
Compatibility: v1g = v12 + v2 g v1g = v1g
dv1g dFK
dv1g Elements: Fb = bv21 FM = M = Kv21
dFK dt dt
Elements: Fb = bv12 FM = M = Kv12
dt dt
Reducing the second set of energetic equations to the system
Reducing these equations to the system equation for the equation for the velocity of the mass, v1g, will yield the previ-
velocity of the mass, v1g ous result

dv1g F (t ) + Fb = FM → F (t ) = − Fb + FM
F (t ) = FM + Fb → F (t ) = M + bv12
dt
dv1g
dv1g F (t ) = −bv21 + M
F (t ) = M
dt
(
+ b v1g − v2 g ) dt
dv1g
dv1g (
F (t ) = −b − v1g − vg 2 + M) dt
F (t ) = M + bv1g − bv2 g
dt
dv1g
dv1g F (t ) = bv1g + bvg 2 + M
b dFK dt
F (t ) = M + bv1g −
dt K dt
b dFb dv1g
F (t ) = bv1g + +M
dv1g b dFb K dt dt
F (t ) = M + bv1g −
dt K dt dv1g
b d
F (t ) = bv1g +
K dt
( − F (t ) + FM ) + M
dt
dv1g b d
F (t ) = M + bv1g − ( F (t ) − FM )
dt K dt b dF (t ) b dFM dv1g
+ F (t ) = bv1g + +M
K dt K dt dt
b dF (t ) dv1g b dFM
+ F (t ) = M + bv1g +
K dt dt K dt b dF (t )
2
bM d v1g dv1g
+ F (t ) = bv1g + 2
+M
K dt K dt dt
4.4  Drawing Linear Graphs from Mechanical Schematics 217

Fig. 4.28   Linear graph symbols 1 2 3 4 6


for translational mechanical
systems

F(t) v(t) M K b

g g g 5 7
Force Source Velocity Source Mass Spring Damper

1 dF (t ) K d 2 v1g K dv1g K labeling the nodes and then adding the elements between
+ F (t ) = + + v1g
M dt bM dt 2 b dt M their respective nodes.

As long as we are consistent in using the arbitrarily defined


positive direction for each element in the linear graph, the 4.4.1  Linear Graph Symbols
signs of the equations will be consistent. The convention of
declaring an arbitrary positive direction for each element The across variable in mechanical systems is velocity.
simplifies the reduction of the linear graph networks to a Springs and dampers have a node at both ends, since they
differential system equation. It is at the end, after we have extend and compress. However, an ideal mass has a single
a system equation, that we will use a free body diagram to velocity because it is rigid. The other velocity node associ-
establish the relationships between the positive directions for ated with the mass is the velocity of the ground node. The
the input and output variables in the linear graph with the linear graph element is drawn between the node represent-
forces and velocities in the actual physical system. ing the velocity of the mass and the ground node. Ground
We will always need to exercise care in interpreting and does not supply a reaction force for a mass, only an inertial
communicating the results of our analyses. Graphical defini- velocity reference. All of the force flowing into a mass ele-
tion of the positive directions of the input and output vari- ment acts to accelerate the mass. The lower half of a branch
ables on the mechanical schematic (physical model) of the representing a mass is dashed to indicate that the force is not
system is the most effective method. transmitted to ground.
We shall see that the different sign conventions of sub- Each branch of a linear graph represents a power flow.
systems of dissimilar types or “energy domains” may be Power can flow into or out of the energy storage and source
contradictory. Unfortunately, the signs of systems with elements. However, power can only flow into damper
energy transformation or conversion are further compli- elements where it is dissipated as heat and lost from the me-
cated by the way subsystems are coupled in a linear graph. chanical system. Power sources differ from “passive” ele-
Further, when a differential system equation is used as the ments. Sources “pump” energy uphill; the across variable
model of a “plant” placed under feedback control, its sign velocity increases in the positive direction of the through
must be “fixed,” so that a positive input to the system yields variable force in a source. In all other elements, the across
a positive value of the output variable or the closed-loop variable, decreases in the positive direction of the through
feedback control system will be unstable. For the mean- variable. Source symbols also differ from the symbols of
time, we will accept the sign yielded by the linear graph the other elements. The convention from electrical circuits
method and not trouble ourselves to determine how that is used, and sources are identified by a circle containing the
sign corresponds to one of the many sign conventions used power variable under our control, Fig. 4.28.
in mechanical engineering.

4.4.2 Force Source Acting on a Parallel


4.4 Drawing Linear Graphs from Mechanical Mass-Damper System
Schematics
The ideal, massless and rigid bar shown in Figs. 4.29 and
The first step in drawing a linear graph is to find the nodes of 4.30 is a force “spreader” which distributes the input force to
distinct values of the across variable and identify them on the mass M and damper b. Since the ideal bar is rigid and mass-
schematic of the system. Once you have identified the nodes less, it has no energetic properties. It functions as an ideal
on the schematic, draw the linear graph by first drawing and force “conductor.” The bar cannot rotate. This constraint is
218 4  Mechanical Systems

Fig. 4.29  a An energetic


schematic of a force source
a Ideal massless b
mass-damper system. b Nodes of Ideal massless and rigid rods
and rigid bar x,v x,v
distinct values of the across vari- b b
able velocity located and labeled 1
on the schematic. The rigid bar g
has a single translational veloc-
ity. The bar and the ideal rigid
F(t) F(t)
1
rods connected to have the same g 1
velocity
M M

Fig. 4.30  a Redundant nodes 1 a x,v


b
b
removed but redundant ground
nodes retained. b Equivalent
force source-mass-damper g
x,v
system F(t)
1 b
g F(t) 1
M M g
g

g g

Fig. 4.31  a The linear graph a 1 b 1 c 1


represents the power flows in
the general case and during the
transient period of a step response.
b The instant after a step input is F(t)
applied to the system, all of the
F(t) b M F(t) b M b M
power from the source flows into
the mass. c When the system has
reached steady-state, all of the
power from the source flows into g g g
the damper and is dissipated as Power flow during Power flow at Power flow in
heat; no power flows into the mass transient period time t = 0 + steady-state

implied, since only the x-axis identified. The horizontal lines the general case. Assume the system is initially de-ener-
connected to the piston and cylinder of the dashpot and to gized, when a step input of force F( t) = F0us( t) is applied
the mass are ideal, massless and rigid rods. The velocity dif- to the system. From the Newtonian perspective, the system
ference which can exist across the dashpot is between the tends towards force equilibrium. Initially, all of the force
piston and the cylinder. In an actual shock absorber, the de- acts to accelerate the mass. When the mass begins to move,
formation of the steel components is minute, relative to the the damper compresses and begins to exert force. Eventu-
displacement of the piston within the cylinder, validating this ally, as the velocity difference across the damper increases,
model of a damper. the force acting to compress the damper equals the applied
Draw the linear graph by drawing and identifying the force, leaving none to accelerate the mass.
across variable nodes and then adding the elements be- From the complementary energetic perspective, the sys-
tween the appropriate nodes, per Fig. 4.31. There are just tem tends towards energy equilibrium. When the energy of
two unique velocities and, hence, two nodes, node 1 and the power supply is made available to the de-energized sys-
ground. We may use multiple ground nodes in large linear tem, the power flow travels into the energy storage element.
graphs for clarity but never multiples of any other nodes. As the power flow into the system continues, and the system
The linear graph shows how both the through variables approaches an energetic equilibrium with the power supply,
and power flows divide. The force applied to node 1 by the less mechanical energy flows into the kinetic energy storage
force source divides between the mass and the damper, in element, and more is dissipated as heat by the damper. When
4.4  Drawing Linear Graphs from Mechanical Schematics 219

Fig. 4.32  a Energetic sche-


matic of a force source mass-two
a Ideal massless and rigid bar b
damper system. b Nodes of x,v x,v
b1 b1
distinct values of the across
variable translational velocity g
located and labeled
F(t) F(t)
b2 1 b2
g 2
M M

4.4.4 Force Source Acting on System


b2 of a Mass and Two Dampers
1 2
A second configuration of a system with a mass and two
dampers is for the force source to act directly on damper
F(t) b1 M b1, Figs. 4.34 and 4.35. The series connection removes the
dynamic response of the damper, since its through and across
variables, Fb and v12, are imposed by the force source. How-
ever, the velocities v1g and v2g are unknown.
g
Fig. 4.33   Linear graph of the force source mass-two damper system
shown schematically in Fig. 4.32 4.4.5 Force Source Acting on a Mass-Spring-
Damper System

the system reaches energetic equilibrium with the power A force source acting on a mass-spring-damper system is
source, no more energy flows into the kinetic energy stor- shown in Fig. 4.36 and its linear graph in Fig. 4.37. Note that
age element. All of the mechanical energy flowing into the there are two ground nodes in this linear graph. The horizon-
system is dissipated as heat. tal line at the bottom is the “ground plane.” Every node on
the ground plane is ground. We could use a single ground
node, but two are used to spread out the graph and make it
4.4.3 Force Source Acting on a System easier to read.
of a Mass and Two Dampers

Rigid elements have a single translational velocity. Masses 4.4.6  Viscoelastic Models
are rigid and are always assigned a velocity node. The ideal
massless and rigid bar of Fig. 4.32, which distributes the ap- Polymers, i.e., plastics, and elastomers, i.e., synthetic rub-
plied force to dampers b1 and b2, also has a distinct veloc- bers, have both elastic strain energy storage and damping
ity and is assigned a node. The force source needs a reac- properties. Materials which combine these two properties at
tion. Since none is shown, we assume that the source reacts room temperature and under low stress are called viscoelas-
against ground, Fig. 4.33. tic, where the prefix, visco-, denotes the viscous energy dissi-
pation property and the suffix, -elastic, denotes strain energy
storage. Although a plastic, elastomeric, or naturally occur-
ring polymer part or specimen is a single object, it requires

Fig. 4.34  a Energetic sche-


matic of a force source mass-two
a x,v
b x,v
damper system. b Nodes of b1 b2 b1 b2
distinct values of the across
F(t) F(t) 1 2
variable translational velocity
M M g
located and labeled g
220 4  Mechanical Systems

b1 by a numerical subscript. Hence, the last element, the spring


1 2 at infinity, is identified as E∞, or in system dynamics nota-
tion, K ∞. This spring constant should not be misinterpreted
as an infinite stiffness, i.e., a rigid object. It is simply the last
spring in a finite range of stiffness theoretically divided into
F(t) b2 M an infinity of values to create a continuum. Needless to say,
a model which requires two times infinity parameter values
is not practical and, if it were, is unnecessary to achieve cre-
g ate a continuum. Needless to say, a model which requires
two times infinity parameter values is not practical and, if it
Fig. 4.35   Linear graph of the force source-mass-two damper system were, is unnecessary to achieve precision needed to fit the
shown schematically in Fig. 4.34
data. Implementation occurs with a finite number of springs
and dashpots, per Fig. 4.40. The Kelvin–Voight model has
two elements in a lumped parameter model to represent its “instantaneous” displacement due to the spring K ∞, which is
energetic behavior: a damper to represent the energy dissipa- not paired with a dashpot in parallel. As we have established,
tion, and a spring to represent the strain energy storage. it is impossible to apply a Heaviside step input to a spring.
Viscoelasticity uses different symbols for the material pa- The “instantaneous” displacement is the displacement that
rameters than are used in system dynamics. The symbol used occurs during the rapid, but not instantaneous, application of
in viscoelasticity for a spring constant is E. The symbol for a fixed “dead” load.
a damping constant is η. The terminology of viscoelasticity
also differs slightly from system dynamics. A time constant 4.4.6.2  The Maxwell Viscoelastic Model
is known as a “relaxation time” and given the symbol T. The A viscoelastic model which has the spring and damper in
term relaxation time has its origin in tests in which a visco- series was introduced by Maxwell, Figs. 4.42 and 4.43. We
elastic material is put under tension and held at a certain ex- have considered this model previously, Fig. 4.18, and con-
tended length, while the force on the specimen is measured. cluded that it is impossible to drive a spring and dashpot in
The relaxation time is calculated from the decay of the force. series with a force source in the general case. We found that
The use of relaxation time has extended beyond viscoelastic- the force source dictated the value of the state variable, the
ity into other areas of engineering, where models are based force acting through the spring, and, thus, eliminated the dy-
on real exponential decays. namic response of the energetic system. Consequently, in the
When mechanical elements are connected such that they general case, a single spring and dashpot pair in series must
act to share a load, they are described as acting in “paral- be driven by a velocity source instead of a force source, to
lel.” The viscoelastic model with the spring and damper in eliminate the impossible instantaneous deformation of the
parallel was introduced by Kelvin and is known as the Kel- spring. We shall see when we investigate frequency response
vin–Voight viscoelastic model. When mechanical elements in Chap. 10, that a sinusoidal force is a useful test input for
are connected such that they carry the same load, they are characterizing viscoelastic materials.
described as acting in “series.” This is the Maxwell visco- Multiple pairs of spring and dashpots of the Maxwell
elastic model. model are ganged in parallel with a spring, Kn, to approxi-
mate a non-linear viscoelastic response, Figs. 4.44 and 4.45.
4.4.6.1  The Kelvin–Voight Viscoelastic Model
A single spring and dashpot pair cannot replicate non-lin- 4.4.6.3  The Zener Viscoelastic Model
ear viscoelasticity. The models shown Figs. 4.38 and 4.39 The Zener viscoelastic model consists of the first and last
are ganged-up with multiple replicas (an infinite number in elements of the infinitely ganged Maxwell model, a single
theory) in a chain with different damping and stiffness ele- Maxwell model in parallel with a spring, Fig. 4.46. This
ments, Figs. 4.40 and 4.41. The spring constant and damp- model is popular enough to be termed the “standard linear
ing coefficients of each spring and dashpot pair is identified model” in viscoelasticity. The linear graph is Fig. 4.47. The

Fig. 4.36  a Schematic of a a b


force source acting on a mass- x,v x,v
spring-damper system. b Nodes b b
of distinct values of the across F(t) K F(t) 1 K 2
variable velocity M M g
g

g
4.6  Modeling Rotational Elements 221

K Beware that all angular measures are dimensionless. Al-


1 2 though we use a degree symbol of superscripted circle, de-
grees, radians, cycles, and revolutions have no units. Work
in a rotational mechanical system is the product of torque
and angular displacement θ. Because angular measures are
F(t) M b dimensionless, the units of torque, [T ] = [ F L ], are the same
as the units of mechanical work, [W ] = [ E ] = [ F L ].

g W = E = T θ12
(4.36)

Fig. 4.37   Linear graph of the force source-mass-spring-damper sys-


tem shown schematically in Fig. 4.36
4.6  Modeling Rotational Elements

Zener model is shown being driven by a velocity source for Machine design relies on rotational mechanical systems, be-
the general case. The steady-state response of the model of cause there is generally no limit on angular displacement.
a sinusoidal force input of varying frequency, termed its fre- Translational displacement of machine elements is almost
quency response, is also used in viscoelastic material testing. always limited. Rotation is an independent motion with ad-
ditional degrees of freedom of movement and must be ac-
counted for separately from translation. Most engineers
4.5  Rotational Mechanical System Elements find translational motion and translational mechanics much
easier to conceptualize than the corresponding quantities of
Rotational mechanical systems are directly analogous to rotational motion. A case in point is momentum. We have
translational mechanical systems. The power variables in a better intuitive sense of the effects of translational mo-
a rotational mechanical system are torque and rotational mentum p than we have for rotational momentum H, which
velocity. We must express rotational velocity as “angular” is why gyroscopes are so fascinating. The parameters of
velocity Ω in radians/sec. Power in rotational mechanical rotational mechanics are also more difficult to estimate than
system is the product of torque and angular velocity. their translational analogs. We have an intuitive sense of the
acceleration of a mass under the influence of a force. If the
 P = T Ω 12 (4.35) object is of one density, the mass scales directly with size.
The rotational acceleration of inertia due to a torque is not
The easiest way to introduce the equations for rotational sys- as intuitive, because both the quantity and the distribution of
tems is by direct analogy to translational systems, Table 4.4. mass relative to the axis of rotation are factors in the mass
We can use a one-to-one set of analogies to relate rotational moment of inertia. The need to use SI units increases the dif-
mechanical systems to translational mechanical systems, ficulty of estimating rotational parameter.
because the mechanisms of energy storage and dissipation
are physically similar. Note that the systems are analogous
to the extent that the same symbols are used for the spring 4.6.1  Mass Moment of Inertia
constants and damping coefficients in the two types of sys-
tems. A subscript “R,” denoting “rotation,” can be used to Recall from your study of dynamics that the relationship for
distinguish K and KR and b and bR, but this is not necessary. rotational analogous to F = ma is
It is clear from the variables used in an equation whether the
motion is rotation or translation. T = Iα

Fig. 4.38  a Energetic schematic


of the Kelvin–Voight visco-
a x,v b x,v
Ideal massless
elastic model. b Kelvin–Voight and rigid bar
viscoelastic model with nodes
of distinct values of the across K K
variable velocity g
F(t) F(t)
b 1 b
g
g
222 4  Mechanical Systems

1 sions of the mass are much smaller than the distance r from
the axis of rotation, then the mass can be treated as a “point
mass.” The mass is accelerated by a torque applied at the axis
of rotation, Fig. 4.48.
F(t) b K Although there is centrifugal acceleration because of the
circular motion, the centrifugal force does no work on the
mass, since there is no displacement in the radial direction,
because the mass moment of inertia is rigid. The force acting
g to accelerate the infinitesimal mass is tangential to the circu-
lar motion. Its magnitude is
Fig. 4.39   Linear graph of the Kelvin–Voight viscoelastic model shown
in Fig. 4.38 dT
= dFm
r
where T is a torque acting to accelerate an inertia in rotation, The instantaneous tangential velocity is the product of the
I is the mass moment of inertia, and  is the angular accelera- angular velocity W and the radius r from the axis of rotation
tion. We will use this expression, but will restrict rotation to
one axis, so that we can use scalar mathematics, and we will vtangental = Ω r
the change symbol for mass moment of inertia. We will use
J for the mass moment of inertia rather than I, so that we can Writing F = Ma in terms of the applied torque and the an-
use I for the area moment of inertia in bending or for electric gular velocity
current. For brevity, we will also refer to the mass moment
of inertia J as “rotational inertia.” dvtangental dT dΩ r dΩ
dFm = dm → = dm → dT = dmr 2
The rotational equivalent of mass M is mass moment of in- dt r dt dt
ertia J. The inertial effect of mass in resisting acceleration in
rotation depends on both the amount of mass and the square All infinitesimal elements of mass dm at radius r from the
of its distance from the axis of rotation. An ideal mass mo- axis of rotation must undergo the same acceleration under
ment of inertia J is rigid, because its only energetic property the same infinitesimal torque dT, since they must rotate to-
is inertia. If it were able to deform, then it could dissipate en- gether in a rigid body. Integrating over all infinitesimal mass
ergy due to internal or external friction or store strain energy. elements dm at radius r yields the torque T and the mass of
We will write the elemental equation for a mass moment the mass moment of inertia,
of inertia J, T = Jα, with angular acceleration α expressed as
the first derivative angular velocity Ω, so that the equation is dΩ dΩ
∫ dT = r ∫
2
dm → T = Mr 2
in terms of the power variables of a rotational system, dt dt

 d Ω1g The inertial effect of the mass at radius r felt at the axis of
TJ = J (4.37)
dt rotation is the mass m
­ oment of inertia

The angular velocity of rotational inertia must be referenced Mass − Moment of Inertia ≡ J = Mr 2
to ground!
The mechanics of rotation can be understood in terms of It is generally the case in machine design that a rotating
translational motion. If we examine a mass moment of in- mass cannot be reasonably modeled as a point mass or a thin
ertia taking an infinitesimal mass dm, such that the dimen- walled cylinder or ring to keep the mass or the rotational

Fig. 4.40   Energetic schematic x,v


of the Kelvin–Voight viscoelastic
model with n-1 dashpots and n
springs K1 K2 K n-1

F(t) Kn
b1 b2 b n-1
4.6  Modeling Rotational Elements 223

Fig. 4.41   Linear graph of K1 K2 Kn-1


the Kelvin–Voight model of
Fig. 4.40
1

b1 b2 bn-1
F(t) Kn

g g

Fig. 4.42  a Energetic schematic


of the Maxwell viscoelastic
a x,v
b x,v
model. b Maxwell viscoelastic b b
model with nodes of the across v(t) K v(t) 1 K 2
variable translational velocity g
g

extension of our previous analysis modified, so that the


K mass of an infinitesimal ring is a function of its distance
1 2
from the axis of rotation. In order to evaluate the integral
over the cross-section of a part, the infinitesimal mass is
expressed as the product of mass density and an infinitesi-
v(t) b mal volume formulated in terms of the distance r from the
axis of rotation
Mass − Moment of Inertia ≡ J = ∫ r dM = ρ ∫ r d vol
2 2

g mass vol

For example, the mass moment of inertia of a cylinder rotat-


Fig. 4.43   Linear graph of the Maxwell viscoelastic model shown in
Fig. 4.42 ing about its axis is formulated by representing the infinitesi-
mal mass as a ring with circumference 2π r , thickness dr, and
length of the cylinder, L, Fig. 4.49,
inertia a radius r from the axis. Although the mass of the part
R R
at the greatest radius from the axis of rotation has the great- r4 R4
J = ρ L 2π ∫ r 2 ⋅ r dr → J = ρ L 2π → J = ρ Lπ
est contribution to the mass moment of inertia, the rotational 0
4 2
0
inertia of mass closer to the axis is typically not negligible.
In these cases, the mass moment of inertia must either be When M is substituted for the product of the materials den-
calculated using integral calculus or with tables. sity and the volume of the cylinder, we find that the mass
Simplest integral formulation of mass moment of in- moment of inertia of a uniform cylinder rotating about its
ertia is that for a revolved body. The formulation is an

Fig. 4.44   Schematic of the


Maxwell viscoelastic model F(t)
Ideal massless and rigid bar

x,v

K1 K2 Kn-1

Kn

b1 b2 bn-1
224 4  Mechanical Systems

1 preciation for the effect of geometry on rotational inertia, to


compute the rotational inertia in terms of density. This is the
approach presented in Figs. 4.52, 4.53 and 4.54. The effort
b1 b2 bn-1 required to calculate the mass moment of inertia of an object
depends on the geometry of the object and the location of
the axis of rotation relative to the axis of symmetry of the
F(t) 2 3 n Kn object, if it has one. The solids are assumed to be of uniform
density ρ .
K1 K2 Kn-1

4.6.3 Mass Moment of Inertia Calculated from


g Area Moment of Inertia

Fig. 4.45   Linear graph of the Maxwell viscoelastic model shown in The mathematical similarity between the area moment of
Fig. 4.44 inertia I and the mass moment of inertia J allows one to
be derived from the other. The similar mathematical form
exists, because the effectiveness of an elemental area in the
axis of symmetry is half that of the same mass at a distance cross-section of a beam in resisting bending increases with
R from its axis of revolution, Fig. 4.50. the square of its distance from the “neutral” axis, where the
axial stress is zero. This is mathematically analogous to the
R4 R2 R2 kinetic energy in rotation increasing with the square of the
J = ρ Lπ → J = ρ
Lπ R 2 → J=M
2 Cylinder 2 2 distance of a mass element from the axis of rotation.
Volume
If a machine component has uniform density and constant
The parallel-axis theorem is used to calculate the mass mo- cross-section, the mass moment of inertia J x − x for rotation
ment of inertia of a solid which rotates about an axis parallel about the x-axis normal to the cross-section through the cen-
to but at a distance from an axis for which the mass moment troid, and the area moment of inertia I x − x for bending about
of inertia of the solid is known. The parallel-axis theorem adds the z-axis through centroid along the length of the compo-
the square of the distance between the mass center and axis of nent are related as
rotation, Fig. 4.51. For example, the mass moment of inertia of
a cylinder rotating about an axis parallel to its own axis yields J x−x = ρ L I x−x
(4.38)

2 R2 where L is the length normal to the cross-section as defined


J = J Cylinder + Mr
 =M + Mr 2
about its Parallel − Axis 2 in Fig. 4.55.
axis Theorem

4.6.4 Mass Moment of Inertia Calculated by


4.6.2 Mass Moment of Inertia of Primitive Superposition
Shapes
The mass moment of inertia of solid volumes not shown in
Although the parallel-axis theorem is easy to remember, it the table can be calculated by superposition by adding and
is not convenient to express the rotational inertia in terms of subtracting mass moment of inertia of solid primitives or
mass, since any change in geometry also changes the mass. It with the parallel-axis theorem, after finding the mass and
is more efficient, and generally provides better physical ap- centroid of the solid.

Fig. 4.46  a Energetic schematic a b


of the Zener viscoelastic model. x,v x,v
b Zener viscoelastic model with Ideal massless
and rigid bar
nodes of the across variable
K1 K1
translational velocity
g
v(t) v(t)
b 1 b
K2 g K2 2
g
4.6  Modeling Rotational Elements 225

K2 by removing two rings and the cylindrical shaft bore from


1 2 a cylinder with the diameter and height of the pulley. When
there is rotation about a single axis, then the rotational inertia
of an object depends only on the distribution of mass in the
radial direction, normal to the axis of rotation, and is indepen-
F(t) K1 b dent of the distribution of mass in the longitudinal direction,
parallel to the axis. Consequently, the geometry can be altered
to simplify calculation without affecting the result, Figs. 4.58
and 4.59. In the case of rotation about two different axes, we
g
would represent the motion using two rotational subsystems
Fig. 4.47   Linear graph of the Zener viscoelastic model shown in
with two different rotational inertias to store the kinetic ener-
Fig. 4.46 gies due to the angular velocities about the two axes.
Pushing the web to the bottom surface of the hub and
Table 4.4   Analogies between translational and rotational mechanical flange allows the mass moment of inertia to be calculated as
elements either the sum of a positive cylinder plus a negative cylinder
Elemental equation Energy equation and a negative ring.
Kinetic energy storage
dv1g 1 4.6.4.1 Example Mass Moment of Inertia
Mass FM = M EM = Mv12g
dt 2 Calculation
d Ω1g 1 Working with the cross-section of the pulley shown in
Rotational inertia TJ = J E J = J Ω 12g
dt 2 Fig. 4.56 and the construction of the positive and negative
Strain energy storage primitives shown in Fig. 4.59a, the mass moment of inertia
dFK FK2 equals the rotational inertia of the Positive Cylinder A minus
Translational spring = Kv12 EK =
dt 2K the rotational inertia Negative Cylinder B and twice the rota-
dTK TK2 tional inertia of Negative Ring A
Rotational spring = K Ω 122 EK =
dt 2K
Energy dissipation (
J = J Cylinder A − J Cylinder B + 2 J Ring A )
Translational damper Fb = bv12 None. Energy
 ρπ LA rA4    ρπ LB rB4   ρπ Lring 4 
dissipater
J =
 2    − 
2   + 2
 2
(
ro ring − ri 4ring  

)
Rotational damper Tb = bΩ 12 None. Energy
dissipater
J=
ρπ  4
2 
( (
LA rA − LB rB4 + 2 Lring ro4ring − ri 4ring  ))
Example Calculate the rotational inertia of a pulley
shown in Fig. 4.56 by the superimposition of the positive Although it increases the time required for a computation, it
and negative mass moment of inertia of solid primitives. The is best to convert units as a discrete step, because it permits
pulley is 2024 aluminum with a unit weight, 0.101 lb/in 3 . a “reasonableness” check before the results are obscured by
The dimensions are in inches. Report the result in SI units. other operations, such as raising the dimension to the fourth
The mass moment of inertia can be calculated by summing power, in this case. If the reader is thinking, “I wouldn’t
the rotational inertia of three rings, Fig. 4.57. know whether the result of the unit conversion is reason-
This is not the only construction which yields the mass able or not,” performing calculations of this sort is how one
moment of inertia. The same shape can be constructed develops a sense of scale

Fig. 4.48   Infinitesimal mass z


element dm in a mass moment of
inertia at radius r from the axis
θ, Ω
of rotation is accelerated by in-
vi = Ω ×r
finitesimal torque dT of applied
H dm
torque T
r
y
T
x
226 4  Mechanical Systems

We must reexamine the routine conversion from pounds to


R kilograms. The conversion factor 1 kg/2.2 lb is a ratio of
dissimilar units since [ kg ] = [ mass ] and [ lb ] = [ force ] . It is
a result of the different fundamental units of US customary
r dr
units, force, length, and time, and SI, mass, length, and time.
Weight is the force exerted by gravity on a non-accelerating
+T,θ,Ω mass

W
F = Ma → W = Mg → M =
Fig. 4.49   Cross-section of a cylinder of length L and radius R normal g
to axis of symmetry
Converting only like to like units yields an apparently dif-
ferent result
6.5 in  2.54 cm 
rA = = 82.6 × 10 −3 m
2  1 in  1lb  0.225 N   3.28 ft 
= 0.453
 32.2 ft   1lb   1 m 
0.5 in  2.54 cm   
rB = = 6.4 × 10 −3 m sec 2 
2  1 in 
N N ⋅ sec 2
= 0.453
m m
5.5 in  2.54 cm 
r0 ring = = 69.9 × 10 −3 m sec 2
2  1 in 
1.5 in  2.54 cm  Recall that 1 N is defined as the force needed to accelerate a
ri ring = = 19.1 × 10 −3 m 1 kg mass at 1 m/sec2,
2  1 in 
kg ·m 1 N ·sec 2 1N
1N = 1 → 1kg = =
 25.4 × m  −3
sec 2
m m
LA = LB = 1 in   = 25.4 × m −3
 1 in  sec 2
Hence, our result is one gram shy of the conversion factor
 25.4 × m −3  −3 0.454 kg/1 lb.
Lring = 0.25 in   = 6.6 × m
 1 in Unfortunately, US customary units are still used by the
3 aerospace industry. Unit conversion errors has led to costly
lb  0.454 kg   39.36 in  kg and well publicized errors, including the crash of a Martian
ρ = 0.101 3     = 2800 3
in  1 lb   1 m  m probe due the incorrect assumption that force was reported in
newtons when, in fact, it was reported in pounds. The most
 kg  unusual US customary units are possibly those used for mass
 2800 3  π
m
(
 25.4 × 10 −3 m 82.6 × 10−3 m)( ) moment of inertia. The rotational inertia of DC servomo-
4
J=  2
2 tors is sometimes given in units of [oz ·in ·sec ] leading to

(( )(
− 25.4 × 10 −3 m 6.4 × 10 −3 m )
4 question “Is that ounce force or ounce mass?” Do not waste
time searching for the appropriate conversion table. Perform
( ) ((69.9 × 10 m) − (19.1 × 10 m) ) a unit analysis. We know that the units of mass moment of
4 4
+ 2 6.6 × 10 −3 m −3 −3
2 2
inertia are [mass ·length ]. Attempt to convert [oz ·in ·sec ]
to SI units presuming that oz. represents ounce force,
J = 3.8 × 10 −3 kg ·m 2 .

Fig. 4.50   Mass moment of


inertia of a right cylinder about
its axis of symmetry r
r4 r2
L J = ρ Lπ =M
2 2
4.6  Modeling Rotational Elements 227

Fig. 4.51   Mass moment of


inertia of a right cylinder about r
an axis parallel to its axis of
symmetry R
R2
J = J Cylinder + M r2 = M + M r2
L about it
Parallel − Axis
2
axis
Theorem

Fig. 4.52   Mass moment of iner-


tia (rotational inertia) J of solids
of uniform density ρ Area A
Radius r
ρ A L3 ρπ r 2 L3
J= =
3 3
Length L

Right Cylinder Rotating About One End

Area A
Radius r
2 ρ A L3 2 ρπ r 2 L3
J= =
3 3
Length L

Right Cylinder Rotating About Its Center

Offset δ Area A
ρA
( L +δ ) −δ 3 
3
J=
3  
Length L

Right Cylinder Offset from Axis of Rotation by Distance δ

ri ro
ρπ L
L J=
2
(r o
4
− ri 4 )

Right Cylindrical Tube or Ring Rotating About Its Axis of Symmetry


228 4  Mechanical Systems

Fig. 4.53   Mass moment of iner- y


tia (rotational inertia) J of solids Radius r
of uniform density ρ

ρπ r 5
J=
x 5

z
Hemisphere Rotating About Its Axis of Symmetry (y)
or an Axis Through the Center of its Base (x and z)
y
Radius r

 r 3δ 2 r 4δ r 5 
Offset δ J = ρπ  + + 
 3 2 5

x
z

Hemisphere Rotating About an Axis Parallel to and Offset


From an Axis Through the Center of its Base an Axis
y Radius r

ρπ r 5
J=
5
x

z
Hemisphere Rotating About an Axis Tangent to its Pole
y Radius r

 r 5 2r 4δ 2r 3δ 2 r 2 ( r + δ )3 
J = ρπ  + + − 
Offset δ
 5 4 3 3 

z
Hemisphere Rotating About an Axis Offset to a Tangent to its Pole

2  1 1b   0.225 N   1m  In this case, oz is ounce force. An engineer would use such


oz ⋅ in ⋅ sec  16 oz   1 lb   39.37 in  a strange unit, in order to avoid working with the inconve-
   
niently scaled magnitudes inherent to most SI units. Scien-
= 3.57 × 10 −4  N ⋅ m ⋅ sec 2  tific notation requires remembering the multiplier and the
exponent. An error in remembering, or manipulating, the
We must introduce the definition of a newton in order to in- exponent creates an order of magnitude error. US custom-
corporate mass into the units, ary units are “human” scaled, meaning that the magnitudes
were greater than one, but not huge, making them easier to
 kg ⋅ m  compute and remember. The older metric system, cgs, for
1
3.57 × 10 −4  N ⋅ m ⋅ sec 2   sec 2  = 3.57 × 10 −4 kg ⋅ m 2 centimeters-grams-seconds, also had this advantage. The re-
 
 1N  sults of conversions from inches to meters performed above
were expressed in the “engineering” notation of m × 10 −3
4.6  Modeling Rotational Elements 229

Fig. 4.54   Mass moment of iner-


tia (rotational inertia) J of solids
of uniform density ρ
ρL
L J=
3
( a b + ab )
3 3

b
a

Rectangular Prism Rotating About an Edge

 a 3b ab3 
L
J = ρL + 
 24 6 
b
_a _a
2 2

Rectangular Prism Rotating About the Center-Line of a Face

L J=
ρL
3
( a (b + δ ) + a (b + δ ) )
3 3

b
ρL
a −
3
( a b + ab )
3 3

Offset δ

Rectangular Prism Rotating About an Axis Offset from an Edge

 a 3 ( b + δ ) a ( b + δ )3 
L J = ρL + 
b  24 6 
 
_a _a
2 2  a 3δ aδ 3 
− ρL + 
Offset δ  24 6 

Rectangular Prism Rotating About an Axis


Offset from the Center-Line of a Face

(millimeters) for that reason. Engineering notation has a z


distinct advantage over scientific notation. By varying the
exponents only by a factor of three, values that are within L
one or two orders of magnitude are likely to be expressed
y
using the same exponent. If a consistent exponent is used,
then to remember the number, one only need remember the
x
mantissa. Engineering notation also reduces errors made in
comparing the magnitudes of values, since one only needs to
compare the mantissas. x z
y
Fig. 4.55   The relationship between the mass moment of inertia J and
area moment of inertia I for a machine element with constant cross-
section is J x − x = ρ L I x − x
230 4  Mechanical Systems

Fig. 4.56   Plan and section of


pulley

A A 0.25

1.0

0.5
1.5
5.5
6.5

Plan Section A-A

Positive lindrical tubes. The rotational (or torsional) spring constant


Positive Positive
Ring A Ring B Ring C is calculated using Eq. 4.28 rearranged as

 TK
K= (4.41)
θ 12

The angular deformation θ12 of an elastic material is cal-


culated using the torsional analog of Young’s Modulus, the
Fig. 4.57   Mass moment of inertia of pulley constructed of three rings Shear Modulus G defined as

 τ=Gγ (4.42)
4.6.5  Torsion Springs
where γ is the shear strain and τ is the shear stress. Shear
Rotational springs store elastic strain energy in torsion. The strain is an angular distortion of a material without volume
common form of the equation for a torsion spring is analo- change. Strain is defined on the infinitesimal scale, Fig. 4.61.
gous to the equation for a translational spring. The torque TK As the distance Δ y approaches zero
acting through the spring is related to the angular displace-
ment θ12 across the spring.  ∆ x dx
lim = ≡γ (4.43)
∆ y→0 ∆ y dy
 TK = Kθ 12 (4.39)
If a machine part’s geometry is constant over a finite distance
Angular displacement θ is not a power variable in a rota- y, and the shear modulus is constant in that volume, then we
tional system, so, as with translational mechanical systems, can approximate shear strain on the macroscopic scale as
the common expression is differentiated with respect to time
to yield an expression in terms of the power variables torque  x
≈γ (4.44)
T and angular velocity Ω y

 dTK dθ Shafts twist when torqued. A shaft with uniform shear mod-
= K 12 = K Ω12 (4.40)
dt dt ulus and constant cross-section will have a constant twist
angle, resulting in a straight line parallel to the axis twisting
Angular deformation of an elastic material stores strain ener- into a helix when torque is applied. The twist of a shaft is not
gy. Calculation of torsional deformation is quite complicated the same as the rotation of a shaft. Twist is the deformation
except for the case of elements with circular cross-sections. of the shaft due to the torque transmitted through the shaft.
Fortunately in machine design, shafts and torsion bars are When a shaft carrying a constant torque rotates or spins to
usually circular in cross-section, either solid cylinders or cy- transmit power, both ends rotate at the same velocity. It is

Fig. 4.58   Two geometries with


the same mass moment of inertia
about their axes of rotation
4.6  Modeling Rotational Elements 231

Fig. 4.59   Two equivalent mass


moment of inertia constructions a b
of positive and negative primi- Positive Negative Positive
tives Cylinder A Ring B Cylinder A

Negative
Rings A

Negative Negative
Cylinder B Cylinder B

Helical line created by


θ1 , Ω 1 θ2 , Ω2 constant twist angle α Angular displacement θ

TK K TK T T
Node 1 Node 2
Straight line before twist
Fig. 4.60   Energetic schematic symbol for a torsion spring
L

a b ∆x Fig. 4.62   A shaft with uniform circular cross-section and shear modu-
τ yx lus will have a constant twist angle when carrying torque. Straight lines
parallel to the axis will be twisted into a helixes

y y

τxy x τxy y x

shear stress, τ(r)


τyx
radius, r
Fig. 4.61   Strain deformation of an infinitesimal element under positive
shear stress τ. The subscript notation for shear stress τ is face, direction

only when the torque carried by the shaft changes, that the
ends of the shaft rotate at different velocities, changing the
twist angle. Fig. 4.63   Distribution of shear stress τ on a plane normal to axis of
rotation and twist of a shaft
The shear strain varies with distance from the axis of rota-
tion on each cross-section. Shear strain equals the limit as the
distance between the cross-sections, ∆L, approaches zero of shear stress increases with the distance r from the axis of
the angular displacement between two cross-sections of the rotation, the torque on an infinitesimal ring increases as r 2.
shaft, θ12, divided by ∆L and multiplied by the radius r from The infinitesimal torque at a distance r from the axis of
the axis of rotation, Eq. 4.34. rotation is

 r θ12 dT = rdF = r τ ( r ) dA
γ ( r ) = lim (4.45)
∆L→0 ∆L
θ
where τ ( r ) = Gγ ( r ) = G r , dF = τ dA, and dT = rdF ,
Shear stress is proportional to shear strain, Figs. 4.62 and Fig. 4.60. L
4.63. Hence, shear stress also varies with the radius, The relationship between torque and angular displace-
ment is derived by integrating the infinitesimal torque over
 r θ12 the cross-section, Fig. 4.64. The surface area of an infinitesi-
τ ( r ) = Gγ ( r ) = G lim (4.46)
∆L→0 ∆L mal ring a distance r from the axis and of width dr is

The torque applied to the shaft is balanced by shear stress
acting on cross-sections normal to the axis. Because the
dA = ∫ r dr d φ = 2π r dr
0
232 4  Mechanical Systems

θ1 , Ω 1 θ2 , Ω2
dr b
Tb Tb
dT = rdF = rτdA
r dφ Node 1 Node 2
r
Fig. 4.65   Schematic symbol for a torsional damper or “drag cup”

Rolling-contact bearing

b b
Fig. 4.64   Distribution of shear stress τ on a plane normal to axis of J
rotation and twist of a shaft T(t)

 θ
T = ∫ dT = ∫ r dF = ∫ r τ dA = ∫ r  r G  dA
A A A A
 L

R
 θ θ 2π r 4 Mass-moment-of-inertia J
T = ∫ r  r G  2π r dr = G keyed or splined to the shaft
0
 L L 4

θ 2π r 4 G 2π r 4 Fig. 4.66   Schematic of a mass moment of inertia J keyed to a shaft,


T =G → T = Kθ ∴ K= which is supported on rolling-contact bearings with damping b and
L 4 L 4 driven by a torque source T( t)

This result contains yet another moment of inertia, in this case, air will be pumped or moved by its motion. Shafts must be
the polar moment of inertia, which describes how the distribu- supported on rolling-contact or hydrodynamic bearings or
tion of area on a cross-section relative to the axis of rotation on simple bushings. These devices have both viscous and
contributes to creating an internal torque. Unfortunately, polar Coulomb friction.
moment of inertia has the symbol J. We will distinguish be-
tween polar moment of inertia and mass moment by identify- 4.6.6.1  Rolling-Contact Bearings
ing a polar moment of inertia with the subscript “polar,” Another common schematic representation of a rotating
element supported on rolling-contact bearings is shown in
 2π r 4 π r 4 Fig. 4.66. The drawing shows ball bearings.
J polar ≡ = (4.47)
4 2 Other types of rolling-contact bearings include roller, ta-
pered roller, and needle bearings.
G J polar
 T= θ The shaft is supported on two rolling-contact bearings.
L
 (4.48) The rolling-contact bearings are shown in cross-section with
Rotational
Spring the plane of the section through the axis of rotation. Often,
Constant
the cross-sections sections of rolling-contact bearings are not
drawn and simply indicated by “X”s. The rotating element is
4.6.6  Rotational Damping indicated to have rotational inertia J. The shaft and rotational
inertia are typically two separate elements keyed, splined, or
Energy is dissipated in rotation by the same mechanisms where coupled together. Each bearing is indicated to have damping
energy is dissipated in translation: slip between surfaces, shear coefficient b. The rolling friction of steel on steel due to the
of fluids, and deformation of viscoelastic materials. The same, deformation of the shaft and rolling-contact bearing is neg-
or a very similar, schematic symbol is used to represent damp- ligible, but the seals contribute Coulomb or dry friction of
ing in rotational systems. The rotational equivalent of a dash- some amount. Bearing are lubricated with either oil or grease
pot is a “drag cup.” Instead of fluid being forced to flow in depending on their use. Each bearing has damping b due to
or out of an orifice or the annular space between a piston and the shear of the lubricant and damping due to the motion of
cylinder, as in a dashpot, fluid is sheared in the annular space the rotating element it supports in air or a more viscous fluid.
between a cylinder and an inner rotating “drag.” The fits and preloads placed on bearings vary, but, in gen-
Although energy dissipation is often deliberately de- eral, the inner race of a rolling-contact bearing is press fit or
signed into rotational systems, it is always present to some a shrink fit onto the shaft and rotates with the shaft. The outer
degree, due to either fluid shear or deformation of viscoelas- race of a rolling-contact bearing is press fit into the bear-
tic materials. Unless a machine rotates in a vacuum, some
4.6  Modeling Rotational Elements 233

a b
b b T(t) T(t) 1
J J J
T(t) g
1 1 b b
g Hydrodynamic
Bearing

g g
The inner bearing race
rotates at the shaft speed
Fig. 4.69   Abstract schematic of a torque source-rotational inertia-
The outer bearing race damper system, where the damping is represented as a hydrodynamic
is fixed in the bearing mount bearing supporting the rotational inertia

Fig. 4.67   Nodes of the across variable angular velocity shown on the
energetic schematic. The inner race of the rolling-contact bearing ro-
by soft tin alloy known as babbit, after its nineteenth century
tates at the shaft’s velocity. The outer race is fixed in the bearing mount inventor, Iaasa Babbitt. The soft alloy protects the surface
and does not rotate of the shaft, in the event of a bearing failure and contact
between the shaft and the bearing surface. Hydrodynamic
bearings were once so common, that they are known as
“plain” bearings. Many former applications are now filled
Rotating by rolling-contact bearings. Although rolling-contact bear-
Shaft ings have a Coulomb friction torque of some magnitude due
Shaft
Ra v1 = Ω1r to seals and packing, we will generally model bearings as
diu
sr
ideal hydrodynamic bearings with a linear relationship be-
Ω1 tween the angular velocity of the shaft relative to the bearing
vg and the torque needed to maintain that velocity.
Bearing In most cases, the damping in a rotational system is de-
Fluid Velocity termined experimentally. In those cases, all of the damping
Profile in the system is reported using a single damping coefficient
Stationary Bearing Ω g
and, if necessary, a single static Coulomb friction torque.
Fig. 4.68   Cross-section through a hydrodynamic bearing in a “pillow An abstract energetic schematic representation of a rota-
block.” The ideal fluid velocity profile is linear between the two solid tional system with a bearing is shown in Fig. 4.69. This rep-
surfaces. A typical gap between the shaft and bearing is 0.001–0.003 in. resentation makes no attempt to represent the configuration
of either the rotational inertia or the type and number of bear-
ing mount or pillow block and does not move relative to the ings supporting the shaft. It simply indicates the presence of
mount, which is often in the machine’s frame and at ground those energetic elements in the system by their parameters,
velocity. In the annotated figure the outer races of the bear- b and J. Note that the torque source acting against the rota-
ings are assumed to be fixed to the machine frame, which is tional inertia must also react against ground, Fig. 4.69b.
an inertial ground. Schematics of rotational systems in which two or more
Again, note that the inner race of bearings rotates with the energetic attributes have the same angular velocity drop
shaft the bearing supports, and the outer race is stationary across them and act in parallel can be difficult to draw and to
relative to the machine frame. Consequently, the angular ve- interpret. If the attributes are possessed by a single machine
locity drops across the bearing from the velocity of the shaft element, or when the attributes are possessed by machine
to ground velocity. elements which are coaxial, then, in a two dimensional draw-
ing, the components lie atop and obscure one another, unless
4.6.6.2  Hydrodynamic Bearings they are drawn in a cross-section normal to the axis of rota-
Hydrodynamic bearings are used to support large forces. The tion. When schematic symbols of two attributes, say a tor-
rotation of shaft shears and drags the lubricating fluid. The sional spring and a drag cup or rotational damper, are drawn
drag pressurizes the lubricant, allowing the fluid to support parallel to one another, so that each is visible, then it appears
the transverse force of the shaft and prevent metal to metal that they could not rotate at the same angular velocity.
contact. Their most familiar application is as the crank shaft
and piston rod bearings in automobile engines. A hydrody- 4.6.6.3  Fluid Coupling
namic bearing cross-section is shown in Fig. 4.68. The bear- A fluid coupling is a device which transmits torque by shear-
ings are made of a steel, copper or, aluminum ring overlain ing fluid captured between two solid components rotating at
234 4  Mechanical Systems

θ1 , Ω 1 θ2 , Ω2 400
b Design A
Tb Tb Design D

Percent of Rated Torque


300
Node 1 Node 2

Fig. 4.70   A fluid coupling symbol. The drag cup symbol is more
200
common

different angular velocities. The most familiar fluid coupling 100


Design C
is the “torque converter” of automatic automobile transmis- Design B
sions, in which vans on one half of the housing pump auto-
matic transmission fluid (oil) at vans on the other half of the 0
0 20 40 60 80 100
housing. The prime advantage of a fluid coupling is that it
Percent of Rated Speed
allows one of the shafts to have zero velocity, while the other
shaft continues to rotate. In its automotive application, this Fig. 4.71   NEMA three phase induction motor standard designs A, B,
allows the engine to run while the drive shaft and the car are C, and D torque vs. speed curves
stopped. The disadvantage of a fluid coupling is the energy
lost in shear of the working fluid. The automatic transmis- power variables. The magnitude of the second power vari-
sion fluid must be cooled to limit its temperature rise. Most able is determined by the amount of power the system draws
automobiles have a cooling coil inside the radiator, which from the source at any instant.
connects to tubes from the transmission. Older automatic Torque sources are generally transducers of some sort, in
transmissions dissipated 10% of the power input as heat. which we control the power variable, which transduces into
Modern automatic transmissions can be “locked-up” at high- torque. Using a positive displacement hydraulic motor as an
way speed with the engine power transmitted through gears example, fluid flow rate is proportional to angular velocity
to increase fuel efficiency. of the motor. Consequently, fluid pressure is proportional to
An actual torque converter is roughly toroidal in shape. torque. If we can (approximately) control the fluid pressure
The schematic symbols used for a fluid coupling is either a entering the hydraulic motor, then we can control the torque
drag cup, Fig. 4.65, or two L-shaped lines, Fig. 4.70, where output. Likewise, in an electric motor, torque is proportion-
the gap between them is the fluid-filled volume, which is al to motor current, and angular velocity is proportional to
sheared. The drag cup symbol is more commonly used. voltage. Although we can control or specify voltage, say, by
closing the switch between a battery and an electric motor,
we cannot control current without a sophisticated feedback
4.6.7  Rotational System Sources device which “modulates” (or changes) the voltage applied
to the system to maintain a specified current. Current sourc-
Shafts are the most common rotational system power sourc- es are available and used in industry specifically so that the
es. During the industrial revolution, entire factories were torque an electric motor produces can be controlled.
powered by a single steam engine and the power distributed AC induction motors are widely used in industry. They
throughout the building by overhead shafts. The legacy of have no “brushes”, which are actually graphite contacts, and
this is seen today on the campuses of older colleges and uni- require less periodic replacement than other types of motors.
versities, where part of the mechanical engineering depart- There are four standard designs of three phase induction mo-
ment remains near the steam plant. tors are available, NEMA (National Electrical Manufacturers
Rotational sources are machines or machine elements in Association) designs, A, B, C, and D. The designs provide
which we can control or specify one of the two rotational different torque-speed curves, as shown in Fig. 4.71. The
power variables, either torque or angular velocity. An exam- torque curve of design C is relatively flat over most of its
ple of a rotational source is a rotating shaft separated from speed range and approximates a torque source.
the rest of the rotational system by a friction clutch. If the A design advantage of induction motors is their torque
machine driving the shaft is powerful enough to approxi- at zero angular velocity. Many rotational motors must rotate
mately maintain the shaft’s speed when the clutch plates are to function. Internal combustion engines are an example. If
closed, then we could model the machine-shaft-clutch as an the crank shaft stops rotating, the engine “stalls” and must
angular velocity source. Note that this criterion depends on be restarted using the auxiliary, electrically powered starter
the amount of power the source can supply and relative to motor. Likewise, a synchronous AC motor cannot produce
that the system draws and can only be satisfied approximate- torque, except when the rotor is in synchronous rotation with
ly. Remember, we cannot control the power drawn from the the rotating magnetic field of the stator. If a synchronous AC
source. Consequently, we can only control one of the two motor falls out of synchrony, an auxiliary motor must bring
4.7  Dynamic Tests 235

Ω max mechanical systems. These phenomena are very difficult to


predict from fundamental principles due to the large effect
Torque-Velocity Data of small geometric features and other factors, as discussed
T0
in Sect. 4.2.4.
Constant
Torque Torque
Viscoelastic materials are not characterized by con-
Model vention tensile tests, because their properties are time
Constant Tmax dependent. Three tests are commonly used. A “creep” test
Velocity
measures the displacement of the material under a fixed
Model
Ω0 load. A “relaxation” test measures the decay of stress under
Angular Velocity a fixed displacement. “Dynamic mechanical analysis” sub-
jects specimens to sinusoidal loads to measure their frequen-
Fig. 4.72   Constant torque and constant velocity models and their op- cy response, the subject of Chap. 10.
erating ranges Ultimately, tests must be performed to validate a model.
A model’s value is its ability to predict the behavior of the
back up to speed. Stepper or stepping motors can produce physically real system. Until a model’s predictions are tested,
a “holding” torque at a constant angular position. Some ro- they remain open to question. Some models cannot be tested
tational hydraulic motors can also produce torque without by subjecting an exemplar of the system to a full scale test.
rotation. We will consider motors in Chaps. 6 and 11. The effect of earthquake ground motions on a large struc-
The operating ranges of the source’s torque and angular ture is an example. In those cases, subsystems or small scale
velocity must both be considered, when modeling a rotation- physical models of the entire system are tested.
al power source. If relatively little power is drawn from the Dynamic test data are interpreted by equating an observed
source by the system, then it may be reasonable to model a value with a corresponding term from the system equation,
source as either a constant torque or a constant angular veloc- or by fitting a response function to the observed response
ity source, Fig. 4.72. If the amount of power drawn from the to test inputs representative of the system’s operating range.
source affects the magnitude of the variable, which we would Parameters for linear models of energetic systems are esti-
like to model as under our control, then we must add a damp- mated by measuring characteristic times, i.e., time constants
er in parallel with the source to represent its behavior. The or periods of oscillation, and initial and final values of ob-
damping coefficient b is the ratio of the maxima of the torque served responses. Dynamic testing of viscoelastic materials
and the angular velocity, where they are chosen to provide is more involved, due to the number of superposed New-
the best fit through the sources performance data, Fig. 4.73. ton or Kelvin–Voight elements of different relaxation times
(or time constants) needed to approximate the response.

4.7  Dynamic Tests


4.7.1 Components with a Single Unknown
Dynamic tests are performed to establish the time-de- Energetic Parameter
pendent parameters of viscous fluids, viscoelastic solids;
to characterize components of energetic systems; and to If an individual machine component has a single significant
confirm or validate an energetic model. The parameters energetic property, then that parameter can be estimated
of the individual elements of energetic models are often from the component’s steady-state response to a load (force
established through testing of a component or the entire or torque) or velocity input, if it is feasible to remove the
actual physical system, when it is feasible. Viscous damp- component from the system, and the means exist to test it.
ing and Coulomb friction are present to some extent in all

Fig. 4.73  a Torque–angular


velocity data for a torque source a b
with two linear approximations System
or models. b Torque source with Tmax Torque-Velocity 1
a damper in parallel to yield a Data
torque–angular velocity model.
This is a Norton source model. A Torque Tmax
Thevenin model is a damper in
T(t) b = ___
series with a source Ω max
Torque-Velocity
Models
Angular Velocity Ω max g
236 4  Mechanical Systems

Fig. 4.74   Load and displace-


ment histories of a linearly Force Extension
elastic component ∆xa = x1- xg
Steady-State Steady-State

ta Time ta Time

The most familiar example is loading an elastic machine be non-negligible. In this case, the mass or mass moment of
component to estimate its translational spring constant K, inertia and its associated friction are estimated by using both
Fig. 4.74. If a material is reasonably modeled as linearly the transient and steady-state portions of the test response, as
elastic, then the deformation of the material is proportional discussed in the following section.
to the force acting through it, where the proportionality con-
stant is the spring constant K,
4.7.2 Components with Multiple Energetic
x1g
( )
FElastic ≡ FK = K x1 − xg = Kx1g ∴ K =
FElastic
Parameters

The discussion above presumed that a machine component


Although less familiar, the torsional spring constant K of had a single energetic property, and the property could be
a shaft or rod is measured similarly, by applying a known calculated from material test data, or by a test of the machine
torque and measuring the resulting angular displacement. component itself. The most common circumstances are that
Although we commonly speak of applying a load, either (1) it is either not feasible or downright impossible to re-
a force or a torque, to an elastic element or structure, we can- move a machine component from a system for testing, or (2)
not instantaneously apply a fixed load. It would require infi- a single machine component has two or more significant en-
nite power. Specifically, the testing machine would need to ergetic properties. In these cases, the unknown parameter or
displace a finite distance at infinite speed. Consequently, we parameters must be determined from the dynamic response
must “build up” the load since an elastic element deforms as of the system to test inputs. Three different test inputs are
it is loaded. Typically, a load is ramped up at a uniform rate, used: an impulse, a step, and a sinusoid. We will defer dis-
Fig. 4.70. The test should span the expected operating range cussion of the response of systems to sinusoidal inputs, until
of the component. In theory, we could use any pair of dis- we discuss frequency response, but, briefly, the steady-state
placement and load data to determine the spring constant K. In magnitude and phase shift of the systems sinusoidal re-
practice, the spring constant is calculated using the expected sponse provide two equations and, consequently, generally
force which the component will most often experience, that is, allows us to solve for two unknown parameters, unless we
its operating point. The initial portion of the load history, near have more unknowns or the unknowns appear in ratios.
the base of the ramp, is not used, because there is slack in the The interpretation of impulse and step response data de-
load path, which is pulled or twisted out at the start of a test. rives directly from the response equations expressed in terms
The damping coefficient b of a translational or rotational of the input and the system parameters. In theory, an impulse
damper is also determined by load tests which span the ex- input produces the homogeneous response of a system. The
pected operating range. Depending on the magnitudes of the steady-state, or particular, response of a system to an impulse
loads and velocities involved and the capability of the testing is zero since the impulse is zero, except at the instant it turns
equipment, the tests may be either continuously varying ve- on. There are two practical difficulties. First, the ideal unit
locities or repeated step inputs of different velocities. In either impulse, δ (t ), is a singularity function defined to have zero
case, the velocity and the corresponding force or torque should duration (or width on the time axis), but its time integral (or
span the expected operating range of the damper. Amateur area) equals unity, which requires its magnitude to be infin-
auto racing enthusiasts characterize the frequency response ity at the point where its argument equals zero. There are
of shock absorbers using an inexpensive “shock dyno,” which no macroscopic physical phenomena which possess these
is a crank and slider mechanism driven by a variable speed properties that we can use as inputs. The most useful tool for
electric motor. Frequency response is the subject of Chap. 10. approximating an impulse input is a large hammer which is
If the Coulomb and viscous friction acting on a mass or a used to excite vibrations in structures. A swinging pendulum
mass moment of inertia are negligibly small, then the mass fills the role of a large hammer for impact testing.
M or mass moment of inertia J are determined by measuring In practice, a step input is often the most convenient and
their acceleration under a known force or torque. It is most realistic test input. Step inputs are convenient, because many
common, however, for the Coulomb and viscous friction to of the systems we design are “on” or “off ”, that is, either
4.7  Dynamic Tests 237

Fig. 4.75  a Force source acting


on a spring K and damper b. a Rigid and
b
b Linear graph of the system massless bar
shown schematically in a F(t) 1

x,v

F(t) b K
b K

g g

connected or disconnected to a source with a constant power The material properties of a tensile specimen can never be
variable. We can perform a test by operating the system in “exactly” those of the corresponding machine component.
its normal mode with its own power supply. For “on-off” The question is always, “How close are the properties of the
systems, this type of test is clearly also the most realistic, test specimen to that of the part?”
and, consequently, most likely to lead to a characterization of To a degree, all material testing are “index” tests, where
the system’s parameters which are useful in dynamic model- the results of the tests correlate with the performance of
ing. A step input has the additional advantage of providing parts, rather than provide material properties. The underly-
non-zero steady-state data, in many cases. ing non-linear nature of materials, their temperature, and
strain rate dependence, and, for flaw-dependent properties,
such as ultimate or rupture strength, the stochastic nature of
4.7.3  First-Order System Step Responses the data, all contribute to making material characterization
challenging. In order to use any of the test results within a
The transient portion of a first-order systems response pro- linear model of an energetic system, we must select a value
vides the system’s time constant which is a ratio of the damp- of the parameter to represent a population of data, limiting
ing coefficient and the energy storage parameter. The initial the precision of the model.
or steady-state value may be a product or ratio of the input
and the damping coefficient. 4.7.3.1 Estimating Spring Constant K and
Polymers (including most biomaterials), amorphous Damping Coefficient b from Step
(or glassy) materials, and crystalline metals at sufficiently Response Data
high temperatures exhibit time-dependent deformation at We will use the first-order force source-spring-damper sys-
constant stress. This phenomenon is known as “creep.” The tem shown in Fig. 4.75 as the first example. The energetic
combination of stress-dependent and time-dependent defor- equations, Sect. 3.1.4.1, and the system equation for the force
mation is called “viscoelasticity.” The size of polymer mol- FK acting through the spring, Eq. 3.29, are reproduced below.
ecules leads to significantly different mechanical behavior
than that of metals and ceramics. Viscoelasticity in polymers Continuity or Node Eq: F (t ) − Fb − FK = 0
is due to the slippage of the “macromolecular” chains past
each other. The slippage is impeded by side groups attached Compatibility or Path Eq: v1g = v1g
to the main molecular backbone. Creep of viscoelastic materi-
als increases with temperature, because the thermal motion of dFK
Element Eqs: Fb = bv1g = Kv1g
the molecules increases the likelihood that side groups which dt
impede slippage will vibrate out the way of one another.
FK2
Manufacturing a dog-bone specimen from the “same Energy Eqs: E sys = E K EK =
2K
material” introduces the possibility that the effects of the
manufacturing processes used to make the actual parts and
b dFK
those used to make the specimens have different effects on F (t ) = + FK (3.29)
the material’s behavior. Solidification processes such as cast- K dt
ing or molding; deformation processes including stamping 
and forcing; and material removal processes such as turn- The step response data are the displacement, x1g, Fig. 4.76.
ing or milling, all have stress, deformation, and temperature We easily derive the system equation for the displacement,
histories which depended on the size and shape of the part.
238 4  Mechanical Systems

12 12
10 10

8 8
x1g(t)
x1g(t) 6
cm 6 x1g (ss)=11.4 cm
cm
4 4 x1g(τ)=7.2 cm
2 2

0 0
0 0.5 1.0 1.5 2.0 0 τ 0.5 1.0 1.5 2.0
t, sec τ ≈ 0.26 sec t, sec
Fig. 4.76   Displacement in centimeters under a step input of 4,000 N Fig. 4.77   Steady-state displacement equals 11.4 cm. The displacement
after one time constant equals 63.2 % of steady-state, 7.2 cm. The cor-
responding time is the time constant τ ≈ 0.26 sec

x1g, by eliminating the spring force FK from Eq. 3.29 with the


substitution FK = Kx1g , Eq. 1.6, shown in Fig. 4.77. The corresponding time t equals one time
constant τ. For these data, 63.2 % of dynamic range is
b dFK b dKx1g
F (t ) = + FK → F (t ) = + Kx1g 0.632 ·11. 4 cm = 7.2 cm
K dt K dt

 1 b dx1g which occurred at approximately 0.26 sec. The time constant


F (t ) = + x1g (4.49)
K K dt is found by unit analysis of the system equation

1 b dx1g b
The step response is a stable first-order growth F (t ) = + x1g → τ= ≈ 0.26 sec
K K dt
 K
 −t
  − t
K
τ
x1g (t ) = x1g ( ss ) 1 − e τ  → x1g (t ) = 0.114 m 1 − e b 
   
N
b ≈ K ⋅ 0.26 sec → b ≈ 35, 000 ⋅ 0.26 sec
The input F (t ) = 4, 000 N us (t ) is constant in steady-state. m
Consequently, the derivative of the output variable is zero in
steady-state, allowing us to determine the spring constant K N ⋅ sec
b ≈ 9,100
m

1 b dx1g ( ss ) 1 Alternatively, we can work from the step response function


F ( ss ) = + x1g ( ss ) → F ( ss ) = x1g ( ss )
K K dt 0 K

F ( ss )    − t
−t K
4, 000 N N x1g (t ) = 0.114 m 1 − e τ  → x1g (t ) = 0.114 m 1 − e b 
=K → = 35, 000 = K
x1g ( ss ) 0.114 m m    

To determine the damping coefficient b, we must first Only positive numbers have logarithms, so, dropping the
estimate the system’s time constant. We can calculate the units of meters, we rearrange the response function as
time constant using any data pair, the initial value and the
dynamic range. However, the calculation is more accurate, if −
K
t 0.114 − x1g (t )
e b
=
we use data from early in the response, where rate of change 0.114
is greatest. A first-order step response progresses 63.2 % of
the remaining change toward its steady-state value in one where the value of 0.114 − x1g (t ) is greater than zero since
time constant τ. 0.114 = x1g ( ss ) . We can now take the natural logarithm

 t
  0.114 − x1g (t ) 
( )
−  
Ω (t ) = Ω ( ss ) 1 − e  − Kt
 → Ω (1τ ) = Ω ( ss ) 1 − e
τ − (1)
ln  e b  = ln  
     0.114 
Ω ( τ ) = Ω ( ss )(1 − 0.368) → Ω ( τ ) = Ω ( ss )( 0.632)
K  0.114 − x1g (t ) 
− t = ln  
A convenient method is to calculate 63.2 % of dynamic range b  0.114 
and then locate that value approximately in the data set, as
4.7  Dynamic Tests 239

Fig. 4.78  a Cross-section of a


flywheel with mass moment of a Mass moment of b
inertia J supported on rolling- inertia J
contact bearings driven by a 1
Rolling-contact
torque source T( t). b Linear bearing
graph of the rotational system
modeling the energy loss from
the bearing and air drag as ideal
T(t) 1
viscous friction with damping b
T(t) J b
g
g

Since the spring constant K is known, the expression above 2,000


can be evaluated using any data pair, other than one in
steady-state, to calculate the damping coefficient b. 1,500

4.7.3.2 Estimating Torque and Damping from Ω(t)


1,000
Step Response Data
RPM
The objective is to determine the energetic parameters of 500
a rotational system consisting of an induction motor driv-
ing a flywheel on a shaft supported by rolling-contact bear-
0
ings, Fig. 4.78. The mass moment of inertia of the flywheel 0 20 40 60 80 100 120
and the damping of the system are unknown. The induction t, sec
motor driving the system has one of the torque vs. speed
Fig. 4.79   Step response of the system to an unknown torque input
curves shown in Fig. 4.71, but the design type of the motor
is unknown. A test is performed in which the motor is ener-
gized and the angular velocity of the flywheel is measured, The dimensions of the flywheel are shown in Fig. 4.80.
Fig.  4.79. Note the response curve resembles a linear first- The flywheel is either iron or carbon steel. Fortunately, the
order step response, but it does not look quite right. Either the weight of the flywheel is known to be 360 lbs. The shaft is a
damping of the system is not linear or the input is not a step. carbon steel, 1.5 in. in diameter and 17 in. long.
In this example, the damping is linear. The torque applied by The dimensions and weight of the flywheel are sufficient to
the motor is not constant. We wish to make an initial esti- calculate the flywheel’s mass moment of inertia. The metal’s
mate of the damping coefficient and the input torque. We will density is needed for the mass moment of inertia calcula-
model the input torque as a step input of unknown magnitude, tion. The flywheel’s volume can be computed by using either
i.e., a constant torque, even though we know that it is not. rings or solid cylinders. The flywheel’s geometry, Fig. 4.80a,

Fig. 4.80  a Cross-section of


the flywheel. b Cross-section a b
of equivalent geometry used for
calculation of the flywheel’s vol- 2.5 in 2.5 in
ume and mass moment of inertia

1.5 in 3.5 in 16 in 3.5 in 1.5 in 16 in

2.5 in 2.5 in

2 in 2 in
4 in 4 in
10 in 10 in
240 4  Mechanical Systems

can be constructed by summing three rings; the flange, web, Express the parameters J and b in terms of the power vari-
and hub. Figure 4.80b was drawn from Fig. 4.80a by shifting ables and time
material parallel to the axis of revolution, leaving the volume
and mass moment of inertia unchanged. Figure 4.80b can dΩ   T  t  T 
be constructed by summing two positive and two negative
[T ] =  J  → [ J ] =   = [T ] = [bΩ ] → [b ] =  

 dt    Ω 
cylindrical volumes; the outer cylinder minus the inner cup
plus the hub projection minus the shaft bore. Check the units of the system equation
The volume is 1, 260 in3 or 0.0207 m3. The density of the
metal is  T (t )   J d Ω1g  T   J Ω 
 b  =  b dt  + Ω1g  →  b  =  b t  + [ Ω ]
 
360 lb lb
= 0.285 3
1, 260 in 3 in  Ω   T  t Ω Ω 
 T T  =  Ω T t  + [ Ω ] → [Ω ] = [Ω ] + [Ω ]
3
360 lb 0.454 kg  1 in  kg
ρ= 3   = 7,890 3 The units check.
1, 260 in 1 lb  0.0254 m  m
We need the steady-state value and the time constant,
both of which we determine from the system equation in
The material is likely a carbon steel. time constant form
The mass moment of inertia will be calculated by su-
perposing the mass moment of inertia of cylinders rotating T0 J d Ω 1g T0
= + Ω 1g ( ss ) → = Ω 1g ( ss )
about their axis, Fig. 4.50, b b dt 0 b
4
r
J cylinder = ρL π
2

J=
ρπ
2
( 4 4
)
4  0.0254 m 
10 in (8 in ) − 8 in (5.5 in ) + 2 in (1.75 in ) − 4 in ( 0.75 in ) 
4

 in


J = 4 . 41kg · m 2 The steady-state velocity is known, Ω1g ( ss ) = 1, 670 RPM.


Rotational speed in RPM must be converted to angular ve-
To proceed, we need the system equation relating the input locity in radians per second calculations in SI units
torque to the angular velocity of the flywheel.
rev 2π rad 1 min rad
Ω1g ( ss ) = 1, 670 = 175
Energetic Equations min rev 60 sec sec

T0 rad
Continuity ( Node ) Eq: T (t ) = TJ + Tb = 175
b sec
Compatibility ( Path ) Eq: Ω1g = Ω 1g In reality, the motor’s torque is a function of its angular
velocity. The motor will reach a torque equilibrium with the
d Ω 1g system in steady-state,
Element Eqs: TJ = J Tb = bΩ 1g
dt
T ( ss ) = bΩ 1g ( ss )
1
Energy Eqs: E sys = E J E J = J Ω12g
2 However, in order to make any estimate, we must model the
torque as constant.
Reduction: Input T( t), Output Ω1g( t) The time constant of the system is

d Ω1g T0 J d Ω1g J
T (t ) = TJ + Tb → T (t ) = J + bΩ1g = + Ω1g → τ =
dt b  b dt b
τ
T (t ) J d Ω 1g We estimate the time constant from the data as if the re-
= + Ω 1g
b b dt sponse were a linear step response, by locating the time in
4.7  Dynamic Tests 241

250
1,600
Motor torque
T(t) 200
1,200
Ω(t) N •m
Ω1g (ss)=1,670 150
RPM 800 Ω(t)
Ω1g (τ)=1,060 rad 100
___
400 sec Flywheel velocity
50
0
0 5 10 15 0
τ ≈ 3.18 sec t, sec 0 5 10 15
t, sec
Fig. 4.81   Time constant estimate
Fig. 4.83   Response of the linear model to the variable input torque of
an AC induction motor
1,600

1,200 The observed response was created using a MATLAB pro-


Ω(t) gram of the Runge–Kutta finite difference solver, using a
Linear model
RPM 800 look-up table for the inductor motor’s torque, discussed in
Observed response Chap. 8. The value used for the damping coefficient b was
400 b = 0.75 N·m·sec/rad . The calculation overestimated the
damping coefficient by 85 %. Correspondingly, the estimat-
0 ed torque applied by the motor, T0 = 243 N · m, is also high.
0 5 10 15
t, sec The maximum torque applied by the motor was 239 N · m,
Fig. 4.83.
Fig. 4.82   Linear model plotted against the observed response The errors in the estimated damping coefficient and torque
are due to the induction motor’s torque varying with its an-
the response at which the system has reached 63.2 % of its gular velocity. The initial portion of the response is driven
steady-state velocity, Fig. 4.81, harder than later portion, decreasing the estimated time con-
stant and leading to the overestimated damping ratio.
0.632 Ω 1g ( ss ) = Ω 1g (τ ) → 0.632 ⋅1, 670 RPM = 1, 060 RPM The initial response of many students is to view the
magnitude of the errors in the estimated values as outra-
We can now estimate the damping coefficient b, geous and to see no value in the calculation. However, it
is important to note that prior to the calculation, we had no
J J inkling of the magnitude of either the torque or the damp-
τ= → b=
b τ ing coefficient. We estimated a value of the motor torque,
which is just outside the range of the actual torque. The
4.41 kg ⋅ m 2 N ⋅ m ⋅ sec input variables and parameter values presented in textbook
b= → b = 1.39
3.18 sec rad problems must be determined by the engineer in practice.
Torque measurements are a particular problem. A torque
the torque input T0, cell of the correct capacity must be inserted into the power
train. This is straightforward in a laboratory situation but
T0 rad
= 175 not when there is an existing system which cannot be taken
b sec out of service and torn apart. Conversely, angular veloc-
ity is simple and cheap to measure. A hand-held laser ta-
 N ⋅ m ⋅ sec   rad  chometer is less expensive than most engineering students’
T0 = 1.39  175  = 243 N ⋅ m
 rad   sec  calculators.

We know that these estimates are in error to some degree. 4.7.3.3 Estimating Mass Moment of Inertia,
Plotting the ideal, linear step response calculated using these Damping, and Coulomb Friction from
estimates will provide a sense of how large the error is, Step Response Data
Fig. 4.82. The step response function is Energy dissipation in mechanical systems is due to shear
displacement between surfaces or within a plastically de-
 −t
 T0  − t
b
forming material. The linear model of ideal viscous fric-
Ω1g (t ) = Ω1g ( ss ) 1 − e τ  → Ω1g (t ) =  1 − e J

  b tion is used to approximate velocity-dependent shear force
or torque. Often, a shear force or torque which is not ve-
242 4  Mechanical Systems

Winding in a helical slot. Rotor friction force or torque applied by the source drives the
constructed of a stack of stamped
plates of electrical (silicon) steel system. This nonsensical behavior is predicted, when an
-i Graphite brush initially energized system is decelerated under a constant
friction force or torque. The model’s results do not remain
at zero when the system reaches zero velocity, since we
cannot control the input as a function of the output in a
linear model. We shall see these results produced by the
model in the following example.
Commutator
A DC servomotor is a motor used in robotics and industrial
φ +i
automation. We shall investigate electric motors in Chaps. 6
Magnetic flux φ emerging Energizing and 11. Briefly, “servo” derives from the Latin for “slave”
from the rotor’s north pole current i and refers to a device or a system under feedback control. A
servomotor has a tachometer and, often, a position encoder
Fig. 4.84   Schematic of the rotor, commutator, and brushes of a DC
motor
built in to provide feedback from the motor to the control-
ler. DC motors and series wound AC motors, which are used
in power tools, have graphite brushes, which are pressed
locity dependent must be included for the energetic model against a rotating contact, the commutator, to progressively
to reasonably approximate the system’s behavior. One energize the windings on the motor’s shaft, Fig. 4.84. The
mechanism is Coulomb friction due to surface-to-surface motor rotates to align the magnetic field of rotor with the
contact. Other phenomena which contribute a constant stationary magnetic field on the motor’s frame or stator. As
shear or torque force are Bingham fluids, which have shear the rotor approaches the equilibrium position, contact is bro-
strength at zero shear velocity, and electromagnetic forces ken with the energized winding and the adjacent winding is
which may act on mechanical elements in electrical ma- energized, repeating the process.
chinery, i.e., magnetic “cogging” in motors due to residual We will derive a linear model using a linear graph, in
magnetic fields. which the kinetic Coulomb friction TC acting on the motor’s
Regardless of the physical mechanism, which creates the shaft is represented as a constant torque applied by a source,
shear force or torque required to displace two solid surfac- Fig. 4.85. We wish to estimate the mass moment of inertia of
es, it is modeled as if it were Coulomb friction, Fig. 4.17. A the motor’s rotor, JM, the linear viscous friction (damping)
source must be used to exert a constant force or torque on coefficient, bM, and kinetic Coulomb friction acting on the
a system. The need to use a source to apply a shear force motor’s shaft, TC. If the motor is energized and brought up
or torque limits the operating range to one direction of mo- to speed, storing kinetic energy in the rotor’s mass moment
tion, since friction acts to oppose motion. If there is an os- of inertia, and the electric circuit opened (the power turned
cillation or vibration, and the motion of a mass or rotational off), then only the mechanical attributes of the motor can
inertia reverses direction, then the friction force must also contribute to the response, since the electrical subsystems is
reverse direction. This non-linearity of behavior is impos- de-energized. The motor coasts down and stops. If the motor
sible to include in a linear model. We shall see, however, were comprised of ideal linear mechanical elements, the
that it is easy to include such action, when we work with change in angular velocity, as it coasts down from an initial
numerical methods in Chap. 8. For the meantime, we must angular velocity, would be an exponential decay. However,
restrict models with Coulomb friction to motion in one di- the Coulomb friction acting the shaft changes the shape of
rection. Further, we must not accept results in which the

Fig. 4.85  a Schematic cross-


section of the DC motor’s rotor
a Motor’s rotor mass
b
with mass moment of inertia moment of inertia JM
1
JM. The shaft is supported on Rolling-contact bearing
rolling-contact bearings. The linear damping, bM
bearings and air resistance of the
irregular shape of the motor’s
1 TC JM bM
rotor are combined as damping TC
bM. Kinetic Coulomb friction TC
acts on the shaft. b Linear graph g
of the mechanical aspect of the
DC motor
Kinetic Coulomb g
friction torque
g
4.7  Dynamic Tests 243

300 Ω1g (0)

0.632 ∆Ω1g
Ω(t) 200 Ω1g (t)
rad
___ ∆Ω1g
sec Ω1g (τ)
100
0
Ω1g (ss)
0
-0.2 0 0.2 0.4 0.6 0.8 0 τ 2τ 3τ 4τ 5τ
t, sec t, time constants
Fig. 4.86   Angular velocity of a DC servomotor after the electric cir- Fig. 4.87   The linear model of the coast-down of a DC motor due to
cuit is opened as the motor coasts to a stop kinetic Coulomb friction and a linear viscous friction predicts a nonsen-
sical negative steady-state angular velocity

the observed response, Fig. 4.86. The abrupt stop is due to The response function is a decay from a positive initial value
Coulomb friction. to a negative steady-state value
Energetic Equations −t
Ω 1g (t ) = ∆Ω 1g e τ + Ω 1g ( ss )
Node: TC = TJ M + TbM
bM

( )
− t
Loop: Ω 1g = Ω 1g Ω 1g (t ) = Ω 1g ( 0) − Ω 1g ( ss ) e JM
+ Ω 1g ( ss )
d Ω 1g
Elements: TbM = bM Ω 1g TJ M = J M
dt The model’s response, plotted in Fig. 4.87, predicts that the
1
Energy: E system = E J EJ = J M Ω12g rotational inertia of the DC motor will coast-down to a ficti-
M M
2 tious negative steady-state velocity Ω1g( ss) due to the con-
Reduction: Input: TC, Output: Ω1g stant Coulomb friction torque TC. Consequently, the results
are only valid for Ω1g > 0. We will not know the fictitious
d Ω 1g steady-state value Ω1g( ss). Consequently, we cannot calculate
TC = TJ M + TbM → TC = J M + bM Ω 1g
dt the dynamic range of the model ∆Ω 1g = Ω 1g (0) − Ω 1g ( ss ).
Without the dynamic range, we cannot estimate the time
TC J d Ω 1g constant τ. At best, knowing that ∆Ω 1g > Ω 1g (0), we can
= M + Ω1g establish a lower limit for the time constant.
bM bM dt
We have three unknowns, TC, JM, and bM,which appear in
The motor’s time constant is ratios in the response function.

TC J d Ω 1g JM JM T
= M + Ω1g → τ= τ= and C = Ω 1g ( ss )
bM bM dt bM bM bM

τ
Having two equations and three unknowns, there is not
Although the Coulomb friction torque TC is the input to this enough information from the coast-down test of the motor.
model, the value of TC is unknown. We know the signs of the Changing the initial angular velocity and repeating the
unknowns. Parameter values, JM and bM, are always positive. coast-down test would not yield new information. We must
Coulomb friction acts to oppose motion. If we define the ini- change the system in a quantifiable way. The most practical
tial angular velocity of the system, Ω 1g (0), to be positive, approach is to couple to the motor’s shaft a known mass mo-
then TC is negative. The steady-state velocity predicted by ment of inertia, a load JL, supported on a shaft with bearings
the model is with unknown damping bL, Fig. 4.88. When we repeat the
test, the time constant of the modified system is
TC  J M  d Ω 1g ( ss ) TC
= + Ω 1g ( ss ) → = Ω 1g ( ss ) JM + JL
bM  bM  dt 0 bM τ=
bM + b L
244 4  Mechanical Systems

Fig. 4.88   Load inertia, JL, at- Load mass moment


tached to the motor’s shaft with Motor’s rotor mass of inertia J L
a flexible coupling. The load
moment of inertia JM
Motor bearings
inertia’s shaft is supported on damping bM
bearings with damping bL Flexible coupling

TC 1
g
g
Load bearings
damping bL

We know JL. We can set an upper limit for the load inertia’s
damping bL with the reasonable assumption that the load iner- 400
tia’s damping is not greater than the motor’s damping, bL ≤ bM , Estimate the slopes at
Ω(0) and Ω(t)=0
and, for the initial curve fitting, that they are equal, bL = bM .
Ω(t) 300
Manual calculations. The second approach is to approxi-
mate the derivative of the motor’s angular velocity in the rad
___
sec 200
system equation by a finite difference,
Motor Motor with Load
100
d Ω 1g ΔΩ 1g

dt Δt
0
The finite difference approximation must be made over 0 0.2 0.4 0.6 0.8 1.0
as small an increment of ∆t as practicable to reasonably t, sec
approximate the derivative. If the measurement of angular Fig. 4.89   Locations for the finite difference approximations of coast-
velocity were ideal, we would use ∆t equal to one time step down test data
in the data. However, there will be noise on the tachometer
signal. We may need to expand ∆t slightly, so that velocity The motor coast-down data yields two useful finite
values at the two end points of the interval do not appear to difference equations
be distorted by noise. Otherwise, it is essential that ∆t be as
small as possible.  ∆Ω1 
TC = J M  + bM Ω 1g ( 0)
Having approximated the derivative with a finite differ-  ∆t1 
ence, we can then evaluate the system equation since we also
know Ω 1g and

d Ω 1g  ΔΩ 1g   ∆Ω 2 
+ bM Ω 1g → TC ≈ J M  + bM Ω 1g TC = J M  + bM Ω 1g ( 0.15)
TC = J M
dt  Δt   ∆t2  0

The very top and very bottom of the coast-down curve are Since Ω 4 g (0.15) = 0, the second equation simplifies to
the most convenient segments to use for the finite differ-
ence equation, Fig. 4.89. The angular velocity is a maxi-  ΔΩ 2 
TC = J M 
mum at the top, so the effect of the velocity dependent,  Δt2 
viscous friction on the deceleration is greatest at the top
of the curve. At the very bottom of the curve, the angular The coast-down test performed with the known load inertia
velocity is zero, so the viscous friction should have no ef- JL coupled to the motor shaft yields a second set of equations
fect on the final deceleration of the motor. Fit tangents to
the curve at the top and bottom of the data to estimate the  ΔΩ 3 
TC = ( J M + J L )  + (bM + bL ) Ω 1g ( 0)
derivative of the angular velocity. Use the smallest value  Δt3 
∆t practicable to increase the precision of the finite differ-
and
ence approximation. The estimated derivatives will have
substantial uncertainty.  ΔΩ 4 
TC = ( J M + J L )  + (bM + bL ) Ω 1g ( 0.90)
 Δt4  0
4.7  Dynamic Tests 245

14.5 2.0
Time Step
ζ = 0.1
∆t = 1´10-4 sec
14.0 0.2
1.5
0.3
Data 0.4
13.5 Data Quantification
rad
___ Negative Error 0.5
Model ∆Ω = 0.6 sec
Model - Data f(t)
___ 1.0
rad 13.0
___ f(ss) ζ = 0.9
sec
ζ = 0.707
12.5 Positive Error 0.5
Positive Error ≡ Model > Data Model - Data
Negative Error ≡ Model < Data
12.0
0.200 0.201 0.202 0.203 0
t, sec 0 0.5 1.0 1.5 2.0
t, sec
Fig. 4.90   Coast-down model plotted against experimental data. Note
the coarser quantification of the vertical axis of the data relative to the Fig. 4.91   A family of underdamped second-order step responses nor-
size of the time step. A positive error occurs when the model’s value is malized to the steady-state value and parameterized by the damping
greater than the data ratio ζ

Equating the damping bL of the load bearings with the damp- minimize the error function. A typical algorithm may have
ing bM of the motor yields 10 or 20 sets of parameter values. The parameter values
tested in an iteration include (1) the parameter set which
 ΔΩ 3  minimized the error function in the previous iteration, (2)
TC = ( J M + J L )  + 2bM Ω 1g ( 0)
 Δt3  parameter sets created by making random changes to the
and that set, and (3) parameter sets created from random values
within broad limits. The algorithm is illustrated in Chap. 8.
 ΔΩ 4 
TC = ( J M + J L )  MATLAB’s optimization tool box also contains a genetic
 Δt4  algorithm function, ga().

We now have four equations with three unknowns. The typi-


cal result of having more equations than we need and un- 4.7.4  Second-Order System Step Responses
certain approximations of the finite differences is a range of
values for TC, JM, and bM. The techniques characterizing a second-order system’s step
The estimates of TC, JM, bM, and bL are refined by curve responses, introduced in Sects. 2.9.2 and 2.9.3, depend
fitting the coast-down models to the data. The curve fitting on the damping ratio of the system and the type of input.
can be performed either manually by iterating the parameter Overdamped systems have two real eigenvalues, s1 = σ1
values, until the fit between the model and the data is satis- and s2 = σ2, a damping ratio ζ ≥ 1, and non-oscillatory im-
factorily fit. One could also do so automatically by use of pulse and step responses. The eigenvalues of underdamped,
a minimization routine in Mathcad or MATLAB. Minimi- oscillatory second-order systems, with ζ < 1, are a complex
zation calculations reduce error between the model and the conjugate pair, s1 , s2 = σ ± jw . Overdamped and under-
data, Fig. 4.90. To prevent positive and negative errors from damped second-order systems with damping ratios ζ greater
canceling each other, the absolute value of the error or the than approximately 0.7 are more difficult to characterize
square of the error is summed than less damped systems, because their impulse and step
responses oscillations which are distinctive and yield the
 SumAbsoluteValueError = ∑ 0
N −1
Model p − Data p (4.50) system’s eigenvalues, Fig. 4.91.
Characterization of second-order systems impulse and
( )
2
SumErrorSquared = ∑ 0
N −1
Model p − Data p (4.51) step responses is easiest using Laplace transformations and
 transfer functions, Sect. 2.10 and Table 2.3. The relevant
Mathcad has a minimization function, which is used inside transform pairs of Table 2.3 are reproduced as Table 4.5, for
a “solve” block. MATLAB’s minimization function, fmin- convenience. Note that the Laplace transformation of the
con(), is part of the add-on “optimization toolbox.” unit impulse is the number one.
A simple and surprisingly effective minimization A transfer function is a multiplicative linear operator. We
technique uses a “genetic algorithm,” in which sets of are familiar with multiplicative unit conversion. For exam-
parameters are randomly changed, retaining those which ple, the conversion from inches to centimeters is performed
by multiplying by the factor 2.54 cm/in, as shown below.
246 4  Mechanical Systems

Table 4.5   Selected Laplace transform pairs 1 1 1


=
f(t) F(s) ( s + a
s  )( s + b) 
s ( s + a )( s + b )

δ (t ) = 0 for t ≠ 0 Input Transfer Output
1 Function
δ (t ) = ∞ for t = 0
Unit step: us (t ) = 0 for t < 0 1 The output could be either the step response of the system
us (t ) = 1 for t > 0 s represented by the transfer function, or the product of two
1 1
operators, integration with respect to time and the transfer
(e − at − e − bt ) function. The former would transform to the time-domain as
b−a ( s + a )( s + b)
a response function. The latter would transform to the time-
1  1  1 domain as the integration of the system equation.
1+
ab  a − b
(
be − at − ae − bt 

) s ( s + a )( s + b ) The ambiguity of the Laplace-domain is more than off-
set by its advantages, but it does require a change in many
ωn
1−ζ 2 (
e −ζω nt sin ω n 1 − ζ 2 t ) ω n2
s + 2ζω n s + ω n2
2 students’ perspective. The result, by itself, is meaningless. A
Laplace-domain expression has meaning only if the entire
logic of the computation is known (and understood).
Transfer functions are multiplicative operators but in the The example second-order system is the force source-
Laplace-domain, Eq. 4.52, spring-mass-damper system of Sect. 3.2.4.3, shown as an
energetic schematic and a linear graph in Fig. 3.31, repro-
2.54 cm duced below for reference, with the energetic equations and
x in = y cm
in the system equations for the spring force, FK, Eq. 3.33, and
the velocity of the mass, v1g, Eq. 3.34.
Output ( s )
 = Transfer Function ( s ) ≡ G ( s ) (4.52)
Input ( s ) Energetic Equations

Output ( s ) Continuity or Node Eq: F (t ) = Fb + FK + FM


Input ( s ) = Input ( s ) G ( s ) = Output ( s )
Input ( s )
Compatibility or Path Eq: v1g = v1g
There is an ambiguity in the Laplace-domain which results
from the form of the singularity inputs: the impulse, step and dFK dv1g
Element Eqs: Fb = bv1g = Kv1g FM = M
ramp. Without knowledge of the history of a calculation, it is dt dt
impossible to identify a term as either a transfer function or F 2
1
a transformed variable, a signal. The impulse input yields an Energy Eqs: E sys = E K + E M EK = K
EM = Mv12g
2K 2
output, the Laplace transformation of the response function,
which is identical to the transfer function. For example,
K d 2 FK b dFK K
1 1 F (t ) = 2
+ + FK (3.33)
1 =  M dt M dt M
Input ( s + a
 )( s + b) ( s + a
 )( s + b)
1 dF (t ) d v1g b dv1g K
2
Transfer Output
Function = + + v1g (3.34)
 M dt dt 2 M dt M
The Laplace transformation of a unit step input is identical
to the Laplace transformation of the operation of integration The objective is to determine the eigenvalues of the system
with respect to time. For example, and, possibly, the magnitude of the input to the system, from

Fig. 3.31  a Force source spring-


mass-damper system of Fig. 3.23.
a b 1
b Linear graph, Fig. 3.24b
x,v

1 F(t) b K M
F(t) K
M g
g

g g
Lubricating fluid
Damping b
4.7  Dynamic Tests 247

an observed impulse or step response, in order to calculate 3.5


the values of unknown parameters of the system. There are 3.0
two approaches. When the eigenvalues are known, they are
2.5
then equated with the expressions for the eigenvalues for- v(t)
mulated from the characteristic equation derived from the 2.0
cm
___
system equation. Alternatively, features and values of the ob- sec 1.5
t3 t1
served response are equated with the corresponding Laplace- 1.0
domain expressions. In the case of underdamped, oscillatory t2
0.5
systems, the coefficients of the system’s characteristic equa-
tion are equated with the coefficients of the corresponding 0.0
0 2 4 6 8 10 12
Laplace-domain expression, which is written in terms of the t, sec
damping ratio ζ and the ideal, undamped natural frequency
ωn, which are calculated from the observed response. Fig. 4.92    Example overdamped second-order impulse response
When the eigenvalues are known, it is always possible  −t −t 

to solve for two unknown parameters by formulating the ei- v (t ) = −10  e 0.67 − e 1.67  . Times, t1, t2, and t3, indicate the locations
genvalues in terms of the systems parameters by solving the of data used in the characterization of the response
characteristic equation with the quadratic formula. Using the
system equation, Eq. 3.34, if one of the parameters is known,
2 2
then the other two can be calculated. For example, if M is  b K  b
2 K
s1 − s2 =   − 4 → ( s1 − s2 ) =   − 4
known, find b and K. M M M M

1 dF (t ) d v1g b dv1g K
2
b K
M  b  2
2 2
= + + v1g → s 2 + s+ =0 K  b
=   − ( s1 − s2 ) → K = − ( s1 − s2 ) 
2
M dt dt 2
M dt M M M 4   
M M 4  M  
 b
2
 b K
±   −4 (4.53) The alternative approach using the Laplace-domain expres-
M M M
s1 , s2 = sions is often easier, as will be illustrated in the following
2 sections.
Solve for b
4.7.4.1  Second-Order System Impulse Responses
2
b 1  b K An impulse response begins and ends at zero. Impulse re-
s1 = +   −4 sponses are created by either an impulsive input, such as a
2M 2  M  M
hammer striking an elastic object, or by a step input acting
and
on a system, in which a system equation has the derivative of
2
b 1  b K the input as the only input term, i.e., Eq. 3.34.
s2 = −   − 4
2M 2 M M
Overdamped Second-Order Impulse Response:  An over-
damped system’s impulse response, Fig. 4.92 and Eq. 4.54,
s1 + s2
is the superposition of two real exponential decays. The con-
 2   2  stant C is the magnitude of a possibly unknown input
b 1  b K b 1  b K
= +   − 4  + −   − 4 
 2 M 2 M M   2M 2 M
 
M 
  C
f impulse (t ) =
b−a
(
e − at − e − bt ) (4.54)
b
s1 + s2 = To avoid the potential confusion between the smaller eigen-
M
value, which is more negative, and the larger eigenvalue,
b = M ( s1 + s2 ) which has the smaller magnitude, we will work with the
response function written in terms of time constants
Solve for K
s1 − s2 ( −t

vest (t ) = A e τ1 − e τ 2
−t
)
 b 1  b
2
K  b 1  b
2
K
= + − 4 −
  − −4  1
 
 2 M 2  M    where τ =
M   2 M 2  M  M  σ
248 4  Mechanical Systems

Characterization begins by finding the larger, slower time 3.5


( )
-t
____ -t
____
constant which dominates the decay portion of the response; 3.0 v(t) = 10 e 1.67 - e 0.67
the faster exponential, with the smaller time constant, has, 2.5
hopefully, decayed to a negligible value. Two data pairs, t1,
v(t1) and t2, v(t2), are selected near the end of the decay portion
v(t) 2.0
cm 1.5
___
( -t
____
vest(t) = 9.28 e 1.70 - e 0.64
-t
____
)
of the response to estimate the first time constant, τ1. sec
1.0
− t1 − t2

v (t1 ) = A e τ1 and v ( t 2 ) = A e τ1 0.5


0.0
Taking the natural logarithm of both expressions 0 2 4 6 8 10
t, sec
 − t1   − t2 
( )
ln v (t1 ) = ln  A e τ1  ( )
ln v (t2 ) = ln  A e τ1    Example overdamped second-order impulse response
Fig. 4.93 
 −t −t 
v (t ) = −10  e 0.67 − e 1.67  and the estimate of the response function
yields
 −t −t 

  − t1
  − t2 vest (t ) = −9.28  e 0.64 − e 1.70 
( )
ln v (t1 ) = ln ( A) + ln e τ1  ln v (t2 ) = ln ( A) + ln e τ1  ( )
Logarithms of negative numbers do not exist. Multiply
−t1 −t2
( )
ln v (t1 ) = ln ( A) +
τ1
( )
ln v (t2 ) = ln ( A) +
τ1
both sides by negative one and then take the natural loga-
rithm

 v (t )  v t − t3  
( 3)
− t3  − t3
Subtracting the second equation from the first eliminates the −t
3
unknown magnitude A − − e  = e τ2
τ1
→ ln  −  − e τ1   = 3
 Aest    A   τ 2
  est 
−t1 −t2
( ) ( )
ln v (t1 ) − ln v (t2 ) = ln ( A) − ln ( A) +
τ1

τ1  − t3
τ2 = → τ 2 = 0.64
  v (t ) − t3   (4.57)
−t1 + t2 ln  −  3 − e τ1  
( )
ln v (t1 ) − ln v (t2 ) = ( ) τ1   Aest 

 −t1 + t2 The unknown input impulse magnitude C is


τ1 = → τ1 = 1.70 (4.55)
( )
ln v (t1 ) − ln v (t2 ) ( ) C
A= → C = A (b − a )
b−a
We now use the same data to estimate A
 1 1
C = A −  → C = 9.07
 v (t1 )  τ 2 τ1 
−t 1
=A → A = 9.28 (4.56)
e τ1 The function used to create the “data” and the estimated re-
sponse function are plotted in Fig. 4.93. The accuracy of the
To estimate the faster, smaller, time constant, select a data estimated values would be improved by iteration or use of a
pair near the beginning of the growth of the response. Both minimization routine, if necessary.
exponentials in the response function contribute to the initial The eigenvalues are used with the solution formulated in
portion of the response terms of the system’s parameters, M, K, and b, Eq. 4.53, to
solve for two if the third is known
 − t3 − t3

v (t3 ) = A  e − e τ 2 
τ1
1 1
s1 = σ 1 = − → s1 = − = −0.58
τ1 1.7
Move the known quantities to the left side
1 1
s2 = σ 2 = − → s2 = − = −1.56
τ2 0.64
v (t3 ) v (t3 )
− t3 − t3 − t3 − t3
τ1 τ2 τ1 τ2
=e −e → −e = −e
A A
4.7  Dynamic Tests 249

t1 n=1
2
n=2 n=3
y1 n=4
y2
v(t) 1 f(t)
___ y3 y4
1.0
m
___ t2 f(ss)
sec
0

-1
0 0.2 0.4 0.6
t, sec 0
0 2Td 4Td 6Td
Fig. 4.94   Underdamped impulse response. Times, t1 and t2, indicate the t, damped periods
peaks used to calculate the damped period Td and the damping ratio ζ
Fig. 4.95   Peak numbering n and amplitudes yn, relative to steady-state
used with the log-decrement formula, Eq. 4.53
Underdamped Second-Order Impulse Response: An
underdamped impulse response, Fig. 4.94, can characterized
using the methods presented in Sect. 2.9.3 to find the eigen- The times and amplitudes of the peaks are t1,
values directly. An alternative approach is to determine the v(0.041) = 2.09 m/sec , and t2, v(0.251) = 0.26 m/sec
damping ratio ζ and the ideal, undamped natural frequency
ωn, from the response. The Laplace transform pair for a unit 1  y1 
ln
impulse response of a second-order underdamped system is n − 1  yn 
written in terms of the damping ratio and the undamped natu- ζ=
2
ral frequency ωn  1  y1  
2
4π +   ln  
 n − 1  yn  
  
 wn  2 
 f (t ) =
−ζw n t
e sin w n 1 − ζ t 1  2.09 
  
 1−ζ 2  wd   ln 
 2 − 1  0.26 
 ζ= → ζ = 0.32
F (s) = w n2 2  1  2.09  
2

4π +   ln 

 s 2 + 2ζw n s + w n2  2 − 1  0.26  

where the observed or damped frequency ωd of the response The damped period, which is the observed or actual period,
is the factor multiply time t. is approximately equal to the duration between adjacent
The general case has an additional coefficient, C, which peaks or troughs. The true damped period is calculated
scales the response and carries units using the zero crossings, but the likely error locating the
zero crossings is great enough to justify using the peak
 f (t ) = C
wn
1−ζ 2 (
e −ζw nt sin w n 1 − ζ 2 t ) (4.58) values

Td = t2 − t1 = 0.251 sec − 0.041 sec → Td = 0.210 sec


The damping ratio ζ is calculated from the response with the
log-decrement formula, Eq. 4.57 and Fig. 4.95. Note that the The observed or actual angular frequency is the time required
peak numbers begin at one and that the amplitude values yn to progress through a full cycle (circle) of 2π rad
are relative to steady-state. The steady-state value of an im-
pulse response is zero, so the amplitude values can be used 2 π rad 2 π rad rad
wd = → wd = → w d = 29.9
directly Td 0.210 sec sec

 1  y1  The relationship between the ideal, undamped natural fre-


ln
n − 1  yn  quency, ωn, and the actual, damped frequency, ωd, is found
ζ= (4.59)
 1  y1  
2 in the sine of the impulse response function:
4π 2 +   ln  
 n − 1  yn  
wd = wn 1 − ζ 2
250 4  Mechanical Systems

The ideal undamped natural frequency ωn is calculated from


L  dF (t )  = L
1  d 2 v1g  b  dv1g  K
the damped frequency ωd:
M
 2 + L   + L v1g { }
 dt   dt  M  dt  M
rad
29.9 1 b K
wd sec rad sF ( s ) = s 2V1g ( s ) + sV1g ( s ) + V1g ( s )
wn = → wn = → w n = 31.6 M M M
1−ζ 2 0.32 sec
1  b K
The eigenvalues of an underdamped second-order system are: sF ( s ) =  s 2 + s +  V1g ( s )
M  M M

s1 , s2 = σ ± jw d 1
V1g ( s ) s
= M
or, written in terms of the damping coefficient, ζ, and the F (s) b K
ideal, undamped natural frequency, ωn: s2 + s+
M M

s1 , s2 = −ζw n ± jw n 1 − ζ 2
(4.60) The parameter values are calculated by equating the char-
acteristic function, the denominator of the transfer function,
The coefficient C, the magnitude of the input which scales with the denominator of the Laplace-domain expression of
the response function, is found by evaluating the impulse re- the Laplace transform pair.
sponse function for a time t other than that of a zero crossing.
A time corresponding to a peak value will minimize error 1
s w n2
due to signal noise. Evaluating the impulse response func- M = 2
tion for time t1 = 0.041sec b K s + 2ζw n s + w n2
s2 + s+
M M
 
wn  2  b K
f (t ) = e −ζw n t
sin w n 1 − ζ t s2 + s+ = s 2 + 2ζw n s + w n2
2    M M
1−ζ  wd 

v ( 0.041 sec ) and then equating coefficients of like powers of s


rad
31.6 b K
= sec e −(0.32)(31.6)(0.041)sin  29.9 rad ⋅ 0.041sec = 2ζw n and = w n2
  M M
1 − 0.322 sec

rad We will again assume that the mass M is known. The spring
v ( 0.041sec ) = 20.7 constant, K, is found from the ideal, undamped natural fre-
sec
quency, ωn,
The ratio of the observed value over this value yields the K
coefficient C: w n2 = → K = M w n2
M
m
2.08 The damping coefficient b is found from the product of the
sec m
C= = 0.10 damping ratio, ζ, and the ideal, undamped natural frequency,
rad rad
20.7 ωn,
sec
b
The response function is: = 2ζw n → b = 2 M ζw n
M
rad
31.6
m sec e − (0.32)(31.6)(t )sin  29.9 rad t 
v (t ) = 0.1  
4.7.4.2  Second-Order System Step Responses
rad 1 − 0.322 sec  Second-order step responses contain more information than
impulse responses since they have either an initial value or
m
−10.1t
rad  a steady-state value. For example, the steady-state force car-
v (t ) = 3.35 e sec sin  29.9 t
 ried by the spring of the example spring-mass-damper sys-
sec sec 
tem can be found from the system equation, Eq. 3.33,
Two of the three parameter values can be calculated if one
is known. The easiest method uses the transfer function de- K d 2 FK ( ss ) b dFK ( ss ) K
F0 us ( ss ) = + + FK ( ss )
rived from the system equation, Eq. 3.34, M 1 dt 2 0 M dt 0 M

F0 = FK ( ss )
4.7  Dynamic Tests 251

The transient portion of a second-order step response is ana- v(ss)-v(t 1) v(ss)-v(t 2)


lyzed using the same techniques employed with a second-or- 10
der impulse response, recognizing that the second-order step t2
response decays to its steady-state value rather than zero. 8
This fact is incorporated in the log-decrement formula. It t1
must be remembered when characterizing an underdamped v(t) 6
second-order step response. cm
___
sec 4
Overdamped Second-Order Step Response:  The Laplace 2
transform pair yields the unit step response function
0
1  1  0 2 4 6 8 10
f unit (t ) =
step
1+
ab  a − b
(
be − at − ae − bt 

) t, sec

  Example
Fig. 4.96  overdamped second-order step response,
The general case for a step input of an unknown magnitude is −t −t
v (t ) = 10 − 16.7e 0.67 + 6.7e1.67. Times t1 and t2 indicate the locations
 C  1  of the data used to characterization of the response
f step (t ) = 1+
ab  a − b
(
be − at − ae − bt 

) (4.61)

which, when written in terms of the known steady-state step response has two. The steady-state value and the coef-
value and the unknown transient terms is ficients, A1 and A2, are expressed in terms of the magnitude
of the step input, C, and the inverses of the time constants,
 f step (t ) = f step ( ss ) + A1e − at − A2 e − bt (4.62) a and b, in Eq. 4.60, giving us three equations in three un-
knowns. We know the steady-state value and we have esti-
where mated τ1 and A1, where

C 1
f step ( ss ) = a=
ab τ1

C  b  C C  a  C C τ1C b
A1= A2 = = v ( ss ) = → v ( ss ) = → C= v ( ss )
ab  a − b  = a ( a − b ) 
ab  a − b 
 b ( − b)
a ab b τ1

b
As with the impulse response, the characterization of an un- v ( ss )
derdamped second-order step response begins with estima- C τ1 bv ( ss )
A1 = → A1 = → A1 =
tion of the larger, slower time constant. Two data pairs are a ( a − b) 11  1 
− b  τ − b
selected in the second half of the transient period, where the τ1  τ1  1
response approaches but has not yet reached steady-state,
Fig.  4.96. The difference between the steady-state value, 1 bv ( ss ) 1 v ( ss ) 1 v ( ss )
v( ss), and the value of the response at time, t1, Eq. 4.53, is −b = → −1 = → = +1
τ1 A1 bτ1 A1 bτ1 A1
used to calculate the time constant

1  v ( ss ) 
−t1 + t2 = τ 2 = τ1  + 1 → τ 2 = 0.64
τ1 = → τ1 = 1.68 b  A1 
( ) (
ln v (t1 ) − v ( ss ) − ln v (t2 ) − v ( ss ) )
Back-substitute to calculate the magnitude of the step input C
We now use the same data and the estimate of τ1 to estimate A1
b 1
C= v( ss ) → C= v( ss ) → C = 9.3
v (t1 ) − v ( ss ) τ1 τ1τ 2
− t1
= A1 → A1 = −16.2
e τ1 The coefficient A2 is:

The analysis of this overdamped step response now diverges C C


A2 = → A2 = → C = −5.93
from that of an impulse response. An overdamped impulse b ( a − b) 1 1 1
response has a single coefficient, whereas an overdamped −
τ 2  τ1 τ 2 
252 4  Mechanical Systems

10 400

8
300
v(t) 6
Ω(t)
cm
___ rad 200
___
sec 4
sec
2
100
Error: Initial values less than zero
0
-1 0
0 2 4 6 8 10 0 0.5 1.0 1.5 2.0
t, sec t, sec
Fig. 4.97   Example of an overdamped second-order step response Fig. 4.98   Third-order system’s step response
−t −t
v (t ) = 10 − 16.7e 0.67 + 6.7e1.67 solid line, plotted against estimated
−t −t
response function v (t ) = 10 − 16.2e1.68 + 5.9e 0.64 dotted line ertia separated from the motor’s shaft by a torsion spring,
Fig. 4.99. The system was running in steady-state at an initial
f step (t ) = f step ( ss ) + A1e − at − A 2 e − bt angular velocity Ω0, when the voltage applied to the system
was given a step increase. The initial angular velocity was
−t −t subtracted from the data of Fig. 4.98.
vest (t ) = 10 − 16.2 e 1.68 − ( −5.93) e 0.64 We shall see in Chap. 6 how to couple the electrical and
mechanical subsystems into a single model, but for the time
Note the estimated response function, vest( t), has an initial being, we shall model the electromechanical system as a
error; its initial values are less than zero. The error between mechanical system, by representing the electrical aspect of
an estimated response function and genuine experimental the motor as a torque source. The system is a third-order
data, as opposed to these computer generated “data,” is re- system, because the motor’s rotor mass moment of iner-
duced though iteration, preferably with a curve fitting mini- tia and the load inertia can have different velocities, since
mization routine, Fig. 4.97. they are separated by the torsion spring. When the system is
running in steady-state under a constant input voltage, both
the motor and the load rotate at the same speed. However,
4.7.5  Higher-Order System Responses when the voltage to the motor is changed, the motor’s iner-
tia begins to change speed before the load inertia. The dif-
A “higher-order” system is a system above second-order. The ference in speed between the motor and the load changes
responses of higher-order linear systems can be constructed, the amount of twist in the torsion spring and, consequently,
by superposing the responses of first and second-order sys- the torque through the spring. The changed spring torque
tems. We will defer discussion of higher-order systems until acts to decelerate the motor’s inertia and accelerate the load
the introduction of state-space in Chap. 7, because the reduc- inertia. The variation in motor speed creates the oscillation
tion of the energetic equations of a higher-order system to a in the tachometer data.
system equation can be time consuming. However, we will We will characterize the first-order step response, and
illustrate that higher-order system’s responses are comprised then remove it from the third-order step response to reveal
of superposed first and second-order responses, by “de- the second-order impulse response. The Laplace transform
composing” a third-order rotational system’s step response, pair for a first-order unit step response from Table 2.3 is
Fig. 4.98.
 1
Decomposition is the term for reversing the process of su-
perposition. Decomposition is accomplished by characteriz-
(
 f (t ) = a 1 − e
− at
)
ing the lower-order responses which superpose to create the  1
F (s) =
higher-order response, creating the corresponding impulse  s (s + a)
or step response function, and then subtracting it from the
higher-order response. This is only practical, if the higher- Multiplying both the time-domain and the Laplace-domain
order response data are available as a computer file. expressions by the factor C creates the general case of a step
The step response data shown in Fig. 4.98 is from the ta- input of any magnitude C
chometer of the DC servomotor used in the coast-down tests,
C C
Sect. 4.7.3.3. The system was configured with the load in- Ω 1st (t ) =
a
(
1 − e − at ) ↔ Ω 1st ( s ) =
s (s + a)
4.7  Dynamic Tests 253

Fig. 4.99   Test system consisting Load mass moment


Motor’s rotor mass of inertia JL
of the mechanical aspects of a moment of inertia JM
DC motor. The system is driven Torsion spring
by the torque T( t) created by spring constant K
the electrical aspect of the motor,
the mass moment of inertia of
the motor’s rotor, JM, a torsion T(t) 1 2
spring with spring constant K,
and an inertial load, JL
g
g
Coupling Coupling
Motor bearings
damping b M Load bearings
damping bL

K 400
1 2
300

Ω(t)
T(t) bM JM bL JL rad 200
___ ∆Ω = 393
sec
0.632 ∆Ω = 248
100

g g
0
0 τ 0.5 1.0 1.5 2.0
Fig. 4.100   Linear graph of the third-order mechanical subsystem
τ ≈ 0.32 sec t, sec
model

Fig. 4.101   Construction to estimate the time constant of the first-order


We need the time constant and the steady-state value to calcu- aspect of the third-order step response
late the eigenvalue, s = − a , and the magnitude factor, C. From
Fig. 4.94, we estimate the steady-state angular velocity as We will estimate the time constant the time corresponding
to 63.2 % progression through the dynamic range, ∆Ω,
rad
Ω ( ss ) = 393
sec rad rad rad
ΔΩ = Ω ( ss ) − Ω ( 0) → ΔΩ = 393 −0 = 393
We will work with the time constant τ where, if the eigen- sec sec sec
value is purely real then s = σ and
 rad  rad
Ω (τ ) = ( 0.632)  393  = 248
1 1  sec  sec
τ= =
σ a
From Fig. 4.101, the time constant is estimated to be
We have two methods for estimating the time constant τ. τ = 0.32  sec
Recall that a first-order step response progresses through
C  −t

63.2 % of its dynamic range during each time constant τ.
We can estimate the time constant by finding the time when
Ω 1st (t ) =
a
(
1 − e − at ) → Ω 1st (t ) = ∆Ω 1 − e τ 
 
the system has progressed 63.2 % of 393 rad/sec. We need
to “eyeball” (visually approximate) the first-order step re- rad  −t

Ω 1st (t ) = 393 1 − e 0.32
sponse curve due to the oscillation, if we use this method. sec  
An alternative technique is to choose a point on the trace
half way between the peaks and troughs of the oscillation. We wish to subtract the value of the first-order step response
That point should lie on the first-order step response, and a function Ω first (t ) from the third-order step response data to
visual approximation of the curve is not needed. We would yield the second-order impulse response superposed on the
then calculate the time constant using the response function, first-order step response. The data were acquired digitally at
since we know the value of the response and the time at each a sampling period Tsample. We must rewrite Ω first (t ) as a dis-
point, Fig. 4.100. crete time function using a time step ∆t equal to the data ac-
254 4  Mechanical Systems

40 40
n=1
20 20
n=4
Ω2nd(t) Ω2nd(t)
0
rad
___ rad
___
0
sec sec
-20 -20
3Td
-40 -40
0 0.2 0.4 0.6 0.8 0 0.1 0.2 0.3 0.4
t, sec t, sec
Fig. 4.102   Second-order impulse response decomposed from the
Fig. 4.103   Construction to determine the damped period Td and the
third-order response of Fig. 4.94. If the impulse response was not sym-
damping ratio ζ of the second-order impulse response
metric about zero, then the estimated time constant would be iterated
to make it so
1  y1 
ln
n − 1  yn 
ζ=
quisition system’s sampling period Tsample, so that the values  1  y1  
2

calculated from the model are synchronized with the data. 4π + 2


 ln  
Otherwise, the subtraction of the first-order model from the  n − 1  yn  
third-order data will yield nonsense. Define the integer coun-
ter variable m such that 1  26.1
 ln 
4 − 1  9.4 
ζ= = 0.054
mmax Δt = tmax  1  26.1 
2

4π 2 +   ln 
 4 − 1  9.4  
The first-order step response model written in discrete time is
Calculate the ideal, undamped natural frequency ωn from the

Ω first ( m∆t ) =
A
a
( )
1 − e − a m∆t = ∆Ω 1 − e ( − m ∆t
τ ) observed, damped natural frequency, ωd, and the damping
ratio ζ,
wd
( ) wd = wn 1 − ζ 2 → wn =
− m ∆t
Ω first ( m∆t ) = 393 1 − e 0.32 1−ζ 2

We now use the counter variable m to perform the subtraction rad


86.1
sec rad
wn = = 86.2
( )
− m ∆t
1 − 0.054 2 sec
Ω second = Ω m − 393 1 − e 0.32
m
We check the estimated damping ratio ζ and ideal, undamped
The subtraction yields the second-order impulse response natural frequency ωn by plotting an oscillatory second-order
shown in Fig. 4.102. The time scale is expanded to show the impulse response against the data. The oscillatory second-
initial portion of the response in Fig. 4.103. order unit impulse response is
The damped frequency, ωd, is calculated from the damper
period, Td. f unit
impulse
(t ) =
wn
1−ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t )
3Td = 0.22 sec → Td = 0.073 sec
Our undamped, natural frequency of 86.1 rad/sec is much
2π rad rad larger than the maximum peak value of the observed impulse
wd = = 86.1
0.073sec sec response, so we must attenuate the unit impulse response.
We will include a magnitude factor, B, in the numerator to
The damping ratio ζ is calculated by means of the log-decre- scale the response. The impulse function, including the scal-
ment formula, Eq. 4.59, ing factor B, is

rad rad wn  
Ω 1 = 26.1 and Ω 4 = 9.4 Ω impulse (t ) = B e −ζw n t sin  w n 1 − ζ 2 t 
sec sec 1−ζ 2  w 
d
4.8  Equivalent Elements 255

40 the relevant equivalent behavior of an equivalent element is


equal energy storage or dissipation. An equivalent element
20 must have the same power flow as the combined power
flows of the original elements to be energetically equivalent
Ω2nd(t)
rad 0 POriginal + POriginal = PEquivalent
___
sec Element A Element B Element

-20 If two elements are replaced by one equivalent element,


then either the across variable or the through variable in the
equivalent element will be the sum of the corresponding
-40
0 0.2 0.4 0.6 0.8 variables in the original elements, Figs. 4.105 and 4.106.
t, sec Elements in series have the same through variable,
Fig. 4.108. Using F and v as generic through and across vari-
Fig. 4.104   Second-order impulse model, dotted line, plotted against
the data, solid line ables, and equating power dissipated in the two dampers in
series in the original system with the single damper in the
equivalent system yields
We guess an initial value for the scaling factor, B = 0.5. We
then plot the impulse response against the data, expecting FA v12 + FB v23 = Psum → Fv12 + Fv23 = Psum
series series
the need to adjust the scaling factor, B, and discover that, elements elements

excluding the first peak, we have an excellent fit to the data,


Fig. 4.104. F (v12 + v23 ) = Psum → Fv13 = Psum
series series
Having validated the first and second-order models by fit- elements elements

ting the data, we now calculate the eigenvalues of the system


from the time constant, τ, of the first-order model and the Conversely, elements in parallel have the same across vari-
damping ratio, ζ, and ideal, undamped natural frequency, ωn, able, Fig. 4.107. Again summing the power dissipated in the
of the second-order model, dampers in parallel and the power dissipated in the single
damper in the equivalent system
1 1
s1 = − → s1 = − → s1 = −3.13
τ 0.32 FA v12 + FB v12 = Psum → ( FA + FB ) v12 = Psum
and s2 , s3 = −ζw n ± jw n 1 − ζ 2 parallel
elements
parallel
elements

The values of the equivalent parameters are determined by


s2 , s3 = − ( 0.054)(86.2) ± j86.2 1 − 0.0542 writing the continuity, compatibility, and elemental equa-
tions for the portion of the linear graph, which contains
s2 , s3 = −4.65 ± j86.1 the similar elements to be replaced. Either the continuity
or the compatibility equations will contain a sum. The set
of equations is reduced to the form of the elemental equa-
4.8  Equivalent Elements tion written in terms of the through and across variables for
the equivalent element. Although you will likely remember
We often wish to simplify a system by replacing two simi- some equivalent element equations, be careful if you wish to
lar elements which are in parallel or in series with a single generalize the parallel and series relationships. The equiva-
“equivalent” element. What does equivalent mean in this lent parameter expressions depend on both these condi-
context? We are dealing with energetic systems. Therefore, tions: whether the elements are in parallel or in series; and

Fig. 4.105  a Linear graph of


a mass damper system with
a b1 b2 b b equiv
dampers, b1 and b2, in series. b 1 2 3 1 3
Linear graph of the equivalent
system with a single equivalent
damper in place series dampers,
b1 and b2 F(t) M F(t) M

g g
256 4  Mechanical Systems

Fig. 4.106  a Linear graph of


a mass-damper system with a b1 b b equiv
damper, b1 and b2, in parallel. b 1 2 1 2
Linear graph of the equivalent
system with a single equivalent
damper in place of parallel damp- b2
ers, b1 and b2
F(t) M F(t) M

g g

Fig. 4.107  a Linear graph of


a b bequiv
dampers, b1 and b2, in parallel b1
between nodes 1 and 2. b Equiv-
alent single damper replacing the 1 2 1 2
two parallel dampers between
nodes 1 and 2, bequiv = b1 + b 2
parallel

b2

Fig. 4.108  a Linear graph of


dampers, b1 and b2, in series a b
between nodes 1 and 3. b b1 b2 bequiv
Equivalent single damper,
1 1 b1b 2 1 2 3 1 3
bequiv = + =
series b1 b 2 b1 + b 2

whether the parameter is multiplying the across variable or


the through variable. 4.8.2  Dampers in Series
Translational and rotational mechanical elements com-
bine identically, but the terminology describing their con- Elements in series have the same through variable, Fig. 4.109.
figuration differs slightly. Rotational dampers or rotational The sum of their across variables equals the across variable
(torsion) springs which have the same angular velocity nodes of the equivalent element, Fig. 4.108.
are said to be coaxial rather than in parallel. The calculations
of equivalent translational dampers and springs below also Continuity: Fb1 = Fb2 = Fbequiv
apply to the rotational analog.
Compatibility: v12 + v23 = v13

Elemental: Fb1 = b1v12 Fb2 = b2 v23 Fbequiv = bequiv v13


4.8.1  Dampers in Parallel

Elements in parallel have the same across variable. The sum Reduction:
of parallel elements’ through variables equals the through
Fb1 Fb 2 Fb Fb
variable of the equivalent element, Fig. 4.107. v12 + v23 = v13 → + = v13 →
equiv
+
equiv
= v13
b1 b2 b1 b2
Continuity: Fb1 + Fb2 = Fbequiv
1 1
Compatibility: v12 = v12  +  Fbequiv = v13
 b1 b 2  series
Elemental: Fb1 = b1v12 Fb2 = b2 v12 Fequiv = bequiv v12 
bequiv
series
Reduction:
Algebra with a sum of ratios often leads to error. It is often
Fb1 + Fb 2 = Fbequiv → b1v12 + b 2 v12 = Fbequiv clearer and less error-prone to avoid dividing by a sum of
ratios. We will use both methods to illustrate. First, we will
(
Fbequiv = b1 + b 2 v12

) divide by a sum of ratios and then clear improper ratios
bequiv
parallel
4.8  Equivalent Elements 257

Fig. 4.109  a Two springs in


a b K equiv
parallel between nodes 1 and K1
2. b Equivalent single spring
replacing the two springs in 1 2 1 2
parallel between nodes 1 and 2
K equiv = K1 + K 2
parallel
K2

Fig. 4.110  a Springs, K1 and K2,


in series between nodes 1 and
a b
3. b Equivalent single spring K1 K2 K equiv
replacing the two springs in
series between nodes 1 and 3 1 2 3 1 3
K1K 2
K equiv =
series K1 + K2

1 1 1
 +  Fbequiv = v13 → Fb = v Continuity: FK1 + FK2 = FKequiv
 b1 b 2  series equiv 1 1 13
series
+
b1 b 2 Compatibility: v12 = v12

b1b 2 b1b 2 dFK1 dFK2


1 Elemental: = K1v12 = K 2 v12
Fb = v13 → Fb = v13 dt dt
equiv
b1b 2 1 1 equiv
b1 b 2 b1 b 2
series
+ series

+ dFKequiv
b1 b 2
b1 b2 = K equiv v12
dt
b1b 2 b1b 2
Fb = v13 → b = Reduction:
equiv
series b1 + b 2 equiv
series
b1 + b 2
dFK1 dFK2 dFKequiv
FK1 + FK2 = FKequiv → + =
Repeating the calculation, we can avoid division by com- dt dt dt
bining the sum of ratios as a single ratio over a common
denominator and then multiply both sides by the inverse K1v12 + K 2 v12 = K equiv v12 → (K1 
+ K 2 ) v12 = K equiv

v12
parallel parallel
ratio K equiv
parallel

1 1  b2 b1 
 +  Fbequiv = v13 →  +  Fb = v13 4.8.4  Springs in Series
 b1 b 2   b1b 2 b1b 2  equiv
Elements in series have the same through variable. The sum
b1 + b 2 b1b 2 b1 + b 2 b1b 2 of their across variables equals the across variable of the
Fb = v13 → Fb = v13
b1b 2 equiv
b1 + b 2 b1b 2 equiv
b1 + b 2 equivalent element, Fig. 4.110.

b1b 2 b1b 2 Continuity: FK1 = FK2 = FKequiv


Fb = v13 → bequiv =
equiv
b1 + b 2 b1 + b 2
Compatibility: v12 + v23 = v13

4.8.3  Springs in Parallel dFK1 dFK2


Elemental: = K1v12 = K 2 v23
dt dt
Elements in parallel have the same across variable,
dFKequiv
Fig. 4.109. The sum of their through variables equals the = K equiv v13
through variable of the equivalent element. dt series
258 4  Mechanical Systems

Reduction: a 1 b 1
1 dFK1 1 dFK2
v13 = v12 + v23 → v13 = +
K1 dt K 2 dt
M1 M2 M equiv
1 dFKequiv 1 dFKequiv  1 1  dFKequiv
v13 = + → v13 =  +
K1 dt K 2 dt  K1 K 2  dt

dFKequiv 1
= v g g
dt 1 1 13
+
K1 K 2 Fig. 4.111  a Masses in parallel. b Equivalent single mass,
  
K equiv M equiv = M1 + M 2
series

Fig. 4.112   Mass and mass 1


Clear fractions moment of inertia elements can
never be in series, since every
mass or inertia must be refer-
enced to ground M1 Impossible
1 KK 1 K1 K 2 K1 K 2
= 1 2 = =
1 1 K1 K 2 1 1 K1 K 2 K1 K 2 K1 + K 2 2
+ + +
K1 K 2 K1 K 2 K1 K2

M2
dFKequiv 1 dFKequiv K1 K 2
= v13 → = v13
dt 1 1 dt K + K
+ 
1 2
g
K K
12 K equiv
K equiv series

series
rigid. The element can have no compliance, or it would also
K1 K 2 store strain energy or dissipate energy. Consequently, if two
K equiv =
series K1 + K 2 masses share the same velocity node, other than ground, then
they are not two masses but a single mass. The equivalent
mass is the sum of the individual masses, Figs. 4.111, 4.112.
4.8.5 Equivalent Mass and Mass Moment Conversely, if the two masses are separated by a spring
of Inertia or damper then they cannot be combined into a single mass,
since they have independent velocities. The test for indepen-
The linear graph system incorporates inertia elements, trans- dence between power variables is whether we can imagine
lational masses, or rotational mass moments of inertia, as an instant in which the power variable has a value of zero
special cases. They are, of course, very common energetic in one element and a non-zero value. We choose an instant,
elements. They are special cases in that they are not the because we do not expect a system to remain in this state.
general case of a linear graph element. There are two restric- We choose zero as one of the values, because this eliminates
tions which apply to inertia elements. First, a branch repre- the possibility that power variables are different but linearly
senting either a force accelerating a mass must be connected related, as we shall is the case when levers, belt drives, and
(referenced) to ground. Second, there is no reaction force gear sets are introduced in Chap. 6.
from ground. The lack of a reaction from the ground node The only way to combine masses separated by a spring
is why the portion of the symbol connected to ground is a or a mass is to revise the model, by removing the element
dashed line rather than a solid line. Although a force acts separating the masses. Revising the model to simplify it is
through a spring or a damper, a force does not act through a justified in the case that the velocity difference between the
mass to the ground node. two masses is judged to be insignificant, or, equivalently, the
Mass and mass moment of inertia elements must either amount of energy stored or dissipated by the elements sepa-
be in parallel, sharing a velocity node, or separated by a rating the two masses is judged to be insignificant.
spring or damping element. Mass and inertia elements are
Problems 259

Summary direction of the drop in the across variable, except for


sources. Sources increase the across variable in the direc-
Table 4.6   Translational mechanical systems E = Fx P = Fv tion of the through variable “flow.”
Energetic Attribute Element Eqs. Energy Eqs. 3. Check the units of the system equation in terms of the
Customary Power variable power variables of the system and time, not fundamental
form form
units or conventional units.
Mass
4. Use superposition to create a pulse input from scaled and
Kinetic energy dv1g 1 2
storage F = ma Fm = m Em = mv1g time shifted Heaviside step functions. Use superposi-
dt 2
tion to create the response function, by scaling and time-
Spring
shifting the unit step response with the same factors and
Strain energy F = K ∆x dFK FK2 time shifts used to create the input function. Each step re-
= Kv12 EK =
storage F = K x12 dt 2K sponse in the response function must be multiplied by the
Damper corresponding time shifted Heaviside unit step function
d x12
to zero-out the response of that term unit its correspond-
Ideal viscous F=c Fb = bv12 ing input acts on the system.
friction dt
Note use of “c” 5. A dynamic system has only one characteristic equa-
tion, independent of our choice of the output variable.
Consequently, there is only one time constant of a first-
order system. All of the power variables in the system
Table 4.7   Rotational mechanical systems E = T θ P = T Ω
vary at the same rate, although their step responses
Energetic Attribute Element Eqs. Energy Eqs.
will differ. Some variables will grow, while others will
Customary Power variable
form form decay, but they do so in synchronization. Second-order
Mass systems have two eigenvalues. If the eigenvalues are
d Ω 1g 1 real, then the system is overdamped and does not have
Kinetic energy J Ω 12g
T = Iα TJ = J EJ = an oscillatory homogeneous or step response. If the ei-
storage dt 2
genvalues are complex conjugates, then the system is
Spring
underdamped and the homogeneous and step responses
Strain energy T = K ∆θ dTK TK2 are oscillatory.
= K Ω 12 EK =
storage T = Kθ12 dt 2K
Damper Problem 4.1 A rotational mechanical which consists of
Ideal viscous a angular velocity source, a drag cup, with damping con-
Tb = bΩ 12
friction stant, b = 8 N ⋅ m ⋅ sec/rad , and a torsion spring constant,
K = 60 N ⋅ m/rad is shown in the schematic Fig. P4.1a. The
system had reached steady-state under the input angular ve-
Problems locity − 20 rad/sec, applied at an unknown time, t < 0, before
a step input of ∆ Ω (t) = 50 rad /sec us (t) was applied at time,
Reminders t = 0 , as plotted in Fig. P4.1b.
1. Draw a linear graph by identifying nodes of distinct val- 4.1.a Derive the system equation that relates the applied
ues of the across variable, velocity for mechanical sys- velocity input to
tems, on the schematic. A rigid object has a single veloc- i The torque acting through the torsion spring.
ity. Ideal rods, bars, and shafts are rigid and massless. ii The velocity drop across the drag cup.
Identify the two nodes of the across variable associated iii The velocity drop across the torsion spring.
with each element in the schematic by identifying nodes Check the units of the system equations in terms of the power
on either end of the schematic symbol, except for transla- variables and time.
tional masses and rotational inertias, Tables 4.6 and 4.7. 4.1.b The input velocity Ω(t) shown in the plot. Determine
Mass and inertia are rigid. They have a velocity node the unit step responses of the system equations derived
and an inertial (ground) reference. After identifying two in part a. Use superposition to determine the responses
nodes for each element, eliminate the redundant nodes, to the angular velocity input plotted in Fig. 4.1b.
and then label the distinct nodes with numbers, except 4.1.c Plot the responses using Mathcad or MATLAB.
ground, which is identified with a “g.” All ground nodes
are the same velocity. Problem 4.2 A rotational mechanical system consist-
2. Use nodal notation for the across variable drop and indi- ing of an angular velocity source, a drag cup (fluid cou-
cate the positive direction of the through variable in the pling) with damping b and a flywheel with mass moment
260 4  Mechanical Systems

Fig. P4.1  a System with an


angular velocity source, drag a b Ω(t)
cup (fluid coupling) and torsion
spring. b Angular velocity input rad
___
Ω(0+) = 30 sec
applied to the system b
Ω(t) K

0 t, sec
rad
Ω(0-) = -20 ___
sec

Fig. P4.2  a System with an


angular velocity source, drag cup
a Flywheel b Ω(t)
(fluid coupling), and flywheel, Rotational Inertia J
rad
b Angular velocity input applied Ω(0-) = 40 ___
sec
to the system
Fluid Coupling
Damping b

0 t,sec
Angular Velocity
rad
Input Ω(t)
Ω(0+) = -20 ___
sec

Fig. P4.3  a Rotational mechani-


cal system comprised of a torque
a b
source, mass moment of inertia T(t) T(t), N m

J and damping b. b Torque input J 150


applied to the system
b
Hydrodynamic 50
Bearing

0 t, sec

of inertia J is shown in Fig. P4.2a. The damping constant 4.2.c Plot the responses using Mathcad or MATLAB.
is b = 80 N ⋅ m ⋅ sec/rad . The mass moment of inertia is
J = 12 kg ⋅ m 2. The system had reached steady-state under Problem 4.3 A rotational mechanical which consists of a
a step input in angular velocity of 40 rad/sec applied at an torque source T( t), a rotational inertia, J = 8 kg · m 2, and a
unknown time, t < 0 , before a step input was applied at time, hydrodynamic bearing with b = 5 N · m ·sec/rad is shown in
t = 0, reducing the angular velocity input to − 20 rad/sec, as Fig. P4.3a. The system had reached steady-state under the
shown in Fig. P4.2b. torque of 50 N · m applied at a unknown time, t < 0, when
4.2.a Derive the system equation that relates the applied at time, t = 0, the input torque is increased to 150 N · m,
velocity input to as shown in Fig. P4.3b.
i The torque acting to accelerate the flywheel. 4.3.a Reduce the equation list to the differential system
ii The angular velocity of the flywheel. equation which relates the input torque T(t) to
Check the units of the system equations in terms of the power i The angular velocity of inertia J.
variables and time. ii The torque TJ acting to accelerate the inertia J.
4.2.b Determine the unit step responses of the system equa- iii The torque acting shear the fluid of the hydrody-
tions derived in part a. Use superposition to solve the namic bearing, Tb.
system equations for the angular velocity input shown Check the units of the system equations in terms of the power
in Fig. P4.2b. variables and time.
Problems 261

Fig. P4.4   a System with an


angular velocity source, drag a End of Shaft b Ω(t)
Rigidly Attached
cup (fluid coupling) and torsion
rad
spring. b Angular velocity input Compliant Shaft Ω(0+) = 40 ___
sec
applied to the system Torsional Spring K

Fluid Coupling
Damping b

Angular Velocity
0 t,sec
Input Ω(t) rad
Ω(0-) = -20 ___
sec

Fig. P4.5  a Rotational mechani-


cal system comprised of a torque
a Drag Cup Shaft rigidly b
source, mass moment of inertia, Damping b attached T(t),N m●

Flywheel 150
J, and damping, b. b Torque Rotational Inertia J
input applied to the system

50

Input Torque 0 1 2 t, sec


T(t)

-100

4.3.b Solve the system equations of part 4.3a for the input Problem 4.5 A rotational mechanical which consists of a
plotted in Fig. P4.3b. torque source T( t), a rotational inertia, J = 5 kg · m 2, and a
4.3.c Plot the response using Mathcad or MATLAB. hydrodynamic bearing with b = 8 N · m ·sec/rad is shown in
Fig. P4.5a. The system had reached steady-state under the
Problem 4.4 A rotational mechanical system consisting of torque of 50 N · m applied at a unknown time, t < 0, when
an angular velocity source, a drag cup (fluid coupling) with at time, t = 0, the input torque pulse shown in Fig. P4.5b is
damping coefficient, b = 8 N · m ·sec/rad , and a torsional applied to the system.
spring with spring constant, K = 60 N · m/rad, is shown in 4.5.a Reduce the equation list to the differential system
Fig. P4.4a. The system has reached steady-state under the equation which relates the input torque T(t) to
input of Ω (t ) = −20 rad/sec applied at an unknown time, i The angular velocity of inertia J.
t < 0, when at time, t = 0, the input angular velocity is ii The torque TJ acting to accelerate the inertia J.
increased to Ω (t ) = 30 rad/sec, as shown in Fig. P4.4b. iii The torque acting shear the fluid of the hydrody-
4.4.a Derive the system equation that relates the input angu- namic bearing, Tb.
lar velocity to Check the units of the system equations in terms of the power
i The angular velocity of the across the spring, Ω2g. variables and time.
ii The torque acting through the spring, TK. 4.5.b Determine the unit step response and the use superpo-
iii The velocity drop across the drag cup (fluid cou- sition to solve the system equations of part a for the
pling), Ω12. input shown in Fig. P4.5b.
Check the units of the system equations in terms of the power 4.5.c Plot the response using Mathcad or MATLAB.
variables and time.
4.4.b Determine the unit step response and the use superpo- Problem 4.6 A rotational mechanical system modeled
sition to solve the system equations of part a for the as consisting of an angular velocity source, two dampers,
velocity input Ω(t) shown in Fig. P4.4b. b1 = 60 N · m ·sec and b2 = 5 N · m ·sec, and torsional
4.4.c Plot the response using Mathcad or MATLAB. spring, K = 40 N · m/rad, is shown schematically in Fig. P4.6a.
262 4  Mechanical Systems

Fig. P4.6  a Angular velocity


source Ω( t) is connected to the
a b
input shaft of a fluid coupling with End of Shaft
Rigidly Attached
damping, b1. The output shaft of Ω(t) 800
the fluid coupling connects to the Fluid Bearing rad
Damping b 2 ___ 600
torsion spring, K, the end of which sec
is free to rotate on bearings with 400
Compliant Shaft
damping, b2, b Angular velocity Torsional Spring K 200
applied to the system by the source
Ω( t) Fluid Coupling
Damping b1 2 4 6 8 10
t, sec
-400
Angular
Velocity
Input Ω(t)

Fig. P4.7  a Force source F(t)


acts the bar which is attached to a x,v
b
dampers, b1 and b2. Damper b1 Ideal massless
is connected directly to ground. and rigid bar b1 4
Damper b2 is connected to spring
K which is attached to ground. F(t)
b Force pulse applied to the F(t) kN 2
system b2
K 0
0 0.5 1.0 1.5
t, sec
-2

4.6.a Derive the system equations for and check their units.
i The difference in the angular velocity of the two The system de-energized when the pulse input of force
ends of spring K. shown in Fig. P4.7b is applied at time, t = 0.
ii The torque in damper b1. 4.7.b Solve the system equations.
iii The angular velocity across damper b2. 4.7.c Plot the responses using Mathcad or MATLAB.
and check their units. 4.7.d Plot the power flow from the source using Mathcad or
A step input of angular velocity Ω (t ) = −20us (t ) was ap- MATLAB
plied to the system at a time, t < 0 , sufficient for the system
to reach steady-state prior to when a second step input is ap- Problem 4.8 A rotational mechanical system modeled
plied at time, t = 0, as plotted in Fig. P4.6b. as consisting of an angular velocity source, two dampers,
4.6.b Solve the system equations. b1 = 60 N · m ·sec and b2 = 5 N ⋅ m ⋅ sec, and a rotational iner-
4.6.c Plot the responses using Mathcad or MATLAB tia, J = 12 kg · m 2, is shown schematically in Fig. P4.8a.
4.6.d Plot the power flow from the source using Mathcad or 4.8.a Derive the system equations for
MATLAB i The angular velocity of the inertia.
ii The torque in damper b1.
Problem 4.7 A translational mechanical system is mod- iii The torque in damper b2.
eled as consisting of a force source, two dampers, and check their units.
b1 = 5.0 kN ·sec /m and b2 = 3.0 kN ·sec/ m , and a spring, The system was in steady-state at time, t = 0, under a
K = 4.0 kN/m, is shown in the schematic, Fig. P4.7a. The previously applied input when the angular velocity pulses
bar is modeled as rigid, massless, and constrained to hori- shown Fig. 4.8b were applied to the system.
zontal motion. 4.8.b Solve the system equations.
4.7.a Derive the system equations for 4.8.c Plot the responses using Mathcad or MATLAB.
i The force in the spring. 4.8.d Plot the power flow from the source using Mathcad
ii The force in damper b1. MATLAB.
iii The velocity of the bar.
Problems 263

Flywheel mass Shaft rigidly attached 4.9.a Derive the system equations for
moment of inerita J to support i The force in the spring.
ii The force in damper b1.
Fluid coupling iii The velocity of the force spreader bar.
damping b1 and check their units.
The system was at rest and relaxed before the velocity
Angular pulse plotted in Fig. P4.9b acted on the system.
Fluid coupling
Velocity damping b2 4.9.b Solve the system equations.
Input Ω(t) 4.9.c Plot the responses using Mathcad or MATLAB.
4.9.d Plot the power flow from the source using Mathcad or
MATLAB
Fig. P4.8  a Angular velocity source Ω( t) is connected to the input
shaft of a fluid coupling with damping b1. The output shaft of the fluid Problem 4.10 The translational mechanical system shown in
coupling connects to the mass moment of inertia J which is supported
on bearings with damping b2 Fig. P4.10a consists of a velocity source which acts on damper
b1 which is connected to mass M. The mass slides on a fluid film
with damping b2. The parameter values are b1 = 300 N ·sec /m,
Ω(t) 800 b2 = 10 N ·sec /m , and M = 100 kg.
rad
___ 4.10.a Derive the system equations for
sec 600 i The force from the velocity source.
ii The velocity of the mass.
iii The force acting to accelerate the mass.
200 and check their units.
The system was in steady-state under the previously ap-
2 4 6 8 10 12 plied step of 100 m/sec at time, t = 0 when the velocity input
-200 t, sec plotted in Fig. P4.10b was applied to the system.
4.10.b Solve the system equations.
-400 4.10.c Plot the responses using Mathcad or MATLAB.
Fig. P4.8  b Velocity pulses applied to the system by the angular veloc-
ity source Ω (t) Problem 4.11 A rotational mechanical system is shown in
Fig. P4.11a.
4.11.a Derive the system equations for
Problem 4.9 A translational mechanical system is mod- i The velocity of the flywheel J.
eled as consisting of a velocity source, two dampers, ii The torque acting through spring K.
b1 = 5.0 N ·sec /m and b2 = 3.0 N ·sec /m , a spring, iii The velocity difference across damper b.
K = 4.0 N/m , and a rigid, massless bar (a force spreader) is and check their units.
shown in Fig. P4.9a. The bar is constrained to horizontal trans-
lation.

Fig. P4.9  a Velocity source v( t)


acts on damper b1 which is at-
a x,v b
Ideal massless 4
tached to the force spreader bar.
Attached between the bar and
and rigid bar b2
ground are spring K and damper 2
b2. b Velocity pulse applied to b1
the system v(t) v(t)
m 0 0
___ 2 4 6
K sec t, sec
-2

-4
264 4  Mechanical Systems

Fig. P4.10  a Schematic of the


translational mechanical system,
a b
b Velocity input applied to the v(t), ___
m
system x,v sec 3
b1 Lubricating 2
v(t) fluid film
Damping b2 1
M

0 5 10 15
-1 t, sec

-2

Fig. P4.11  a A rotational system


consisting of torque source T( t),
a Shaft rigidly attached b
Hydrodynamic
a flywheel with mass moment bearing damping b
of inertia J, a compliant shaft
Compliant shaft 400
represented schematically as a torsion spring K T(t),N m

torsion spring K. The shaft turns Flywheel


rotational inertia J 300
a drag cup with its output shaft
rigidly fixed. b Torque applied 200
by the source to the rotational 100
mechanical system
Torque 5 10 15 20 25
input T(t) -100 t, sec

The system was in steady-state under the torque input of + 100 rad/sec at time t = 0, then to + 400 rad/sec at t = 5  sec,
− 200 N·m before the torque was increased to + 200 N·m at and then to 0 rad/sec at t = 10 sec. as shown in Fig. P4.12b.
time, t = 0, and then to + 400 N·m at t = 10 sec, as shown in 4.12.b Solve the system equations for the parameter values:
Fig. P4.11b.
4.11.b Solve the system equations for the parameter values: b1 ,
N · m · sec
b2 ,
N · m · sec N·m J , kg · m 2
K,
rad rad rad
N·m·sec N ·m J , kg · m 2 i 1,000 0.5 3,000 50
b, K,
rad rad ii 100 0.5 3,000 50
i 10 1,500 25 iii 1,000 0.5 1,500 500
ii 1 1,500 30
iii 5 3,000 10
4.12.c Plot the responses using Mathcad or MATLAB.
4.12.d Plot the power flow from the source using Mathcad
4.11.c Plot the responses using Mathcad or MATLAB. or MATLAB.
4.11.d Plot the power flow from the source using Mathcad
or MATLAB. Problem 4.13 A translational mechanical system is modeled
as consisting of a force source, two masses, M = 400 kg
Problem 4.12 A rotational mechanical system is shown in and M = 800 kg , and two dampers, b1 = 1, 000 N ·sec /m
Fig. P4.12a. and b2 = 2, 000 N ·sec /m , as shown schematically in Fig.
4.12.a Derive the system equations for P4.13a.
i The torque acting through spring K. 4.13.a Derive the system equations for
ii The velocity difference across damper b1. i The force acting through damper b1.
iii The velocity of the Flywheel J. ii The force acting through damper b2.
iv The torque acting to accelerate the flywheel. iii The velocity of mass M1.
and check their units. iv The velocity of mass M2.
The system was in steady-state under the angular veloc- v The force acting to accelerate mass M2.
ity input of − 100 rad/sec before the velocity was changed to and check their units.
Problems 265

Fig. P4.12  a A rotational system


consisting of velocity source
a Hydrodynamic
Shaft rigidly attached b
Ω(t) driving a fluid coupling bearing damping b2
with damping b1, a torsion spring Flywheel
rotational inertia J
K, and a flywheel supported on Ω(t) 400
a bearing with damping b2. The Compliant shaft rad 300
___
hydrodynamic bearing is repre- torsion spring K sec
200
sented schematically as a drag Fluid coupling
cup. b Angular velocity applied damping b1 100
by the source Ω( t) driving the
rotational mechanical system Angular velocity 2.5 5.0 7.5 10.0 12.5
input Ω(t) t, sec
-200

Fig. P4.13  a Force source F (t )


acts on a systems of two masses
a b
and two dampers. b Force pulse 2
in kilonewtons applied to the x,v
system 1
b1 b2 t, sec
F(t) F(t)
M1 M2 kN 0 0 10 20
-1

-2

Fig. P4.14  a Velocity source


v (t ) acts on a system of two
a b
2
masses and two dampers,
b Velocity pulse applied to the x,v
Lubricating 1
system
K fluid film v(t)
v(t)
M Damping b m 0 0
___ 5 10 15
sec t, sec
-1

-2

The system was at rest when the force pulse shown in The system was at rest when the velocity pulse shown in
Fig. P4.13b was applied to the system. Note that the units of Fig. P4.14b was applied to the system.
force are kilonewtons. 4.14.b Solve the system equations.
4.13.b Solve the system equations. 4.14.c Plot the responses using Mathcad or MATLAB.
4.13.c Plot the responses using Mathcad or MATLAB. 4.14.d Plot the power flow from the source using Mathcad
4.13.d Plot the power flow from the source using Mathcad or MATLAB.
or MATLAB.
Problem 4.15 A translational mechanical system is shown
Problem 4.14 A translational mechanical system is shown schematically in Fig. P4.15a.
schematically in Fig. P4.14a. Parameter values are M = 10 kg, 4.15.a Derive the system equations for
b = 0.4 N ·sec /m , and K = 3,000  N/m. i The force acting through spring K.
4.14.a Derive the system equations for ii The velocity drop across spring K.
i The force acting through spring K. iii The velocity drop across damper b.
ii The force acting through the lubricating film iv The velocity of mass M.
damping b. and check their units.
iii The velocity of mass M. The system was at rest when the velocity pulse shown in
iv The force acting to accelerate the mass M. Fig. P4.15b was applied to the system.
and check their units.
266 4  Mechanical Systems

Fig. P4.15  a Velocity source


v (t ) acts on a systems of two
a b
masses and two dampers. b Ve- 2
locity pulse applied to the system x,v
1
b v(t)
v(t) K
M m 0 0
___ 5 10 15
sec t, sec
-1

-2

Young’s modulus E
Problem 4.16 A cantilevered beam with rectangular cross-
Density ρ
Width b section is shown in Fig. P4.16. Attached to the beam is a
cylinder of mass M. The diameter of the cylinder equals the
M width of the beam b. The height of the cylinder is twice its
diameter.
h
4.16.a Determine the undamped angular frequency of the
L beam and mass system when they are carbon steel
and L = 40  mm, b = 5  mm, and h = 1  mm.
4.16.b Determine the undamped angular frequency of the
beam and mass system when they are carbon steel
Fig. P4.16   A cantilevered beam with a cylindrical mass attached at and L = 50  mm, b = 10  mm, and h = 1  mm.
distance L from the support 4.16.c Determine the undamped angular frequency of the
beam and mass system when they are aluminum
4.15.b Solve the system equations for the parameter values: 6061 and L = 50  mm, b = 10  mm, and h = 1  mm.
i M = 500 kg, b = 2, 000 N ·sec /m, K = 1,000 N/m.
ii M = 300  kg, b = 400 N ·sec /m , K = 4,000  N/m. Problem 4.17 A rotational mechanical system consisting
iii M = 500 kg, b = 1, 000 N ·sec /m , K = 6,000 N/m. of a torque source driving a shaft with a flywheel is shown
4.15.c Plot the responses using Mathcad or MATLAB. in Fig. P4.17. The shaft and flywheel are carbon steel. The
4.15.d Plot the power flow from the source using Mathcad shaft is solid and has a circular cross-section. It is sup­
or MATLAB.

Fig. P4.17  a Rotational


mechanical system. b Plan and a Mass moment of inertia, J
cross-section of the flywheel
Rolling-contact bearing, b

T(t)

L
b

A A e

a
b
c
d
Plan Section A-A
Problems 267

Table P4.17   Flywheel and shaft dimensions Problem 4.18 A test was performed to determine the ener-
a (in.) b (in.) c (in.) d (in.) e (in.) f (in.) L (ft.) getic element parameters of translational mechanical system
i 1.5 3 18 20 0.5 2 10 shown in Fig. P4.18a. The system was de-energized before
ii 2 4 20 24 0.5 3 20 it was subjected to a step input of force, F (t ) = 100.0 N us (t ).
iii 3 5 30 36 0.5 4 30
The velocity of its point of application of the force on the
ideal massless and rigid bar was measured. The data are pre-
sented in Fig. P4.18b and Table P4.18.
ported by three rolling-contact bearings, each with damping Determine the damping coefficient b and the spring con-
b = 0.5 N · m ·sec/rad. stant K. Report your results in the correct SI units.
4.17.a Calculate the torsion spring constant K of the shaft
and the mass moment of inertia J of the flywheel for Problem 4.19 A machine’s energetic attributes are modeled
the following dimensions. as a mass moment of inertia J supported on an ideal, rota-
4.17. b Determine the angular velocity of the flywheel when tional damper b, Fig. P4.19a. A test was performed in which
a step input of 200 N-m of torque is applied to the a step input torque of 1,000 N-m was applied to the system
de-energized system. and the angular velocity of the inertia measured, Fig. P4.19b
4.17.c Plot the response of the system in Mathcad or MAT- and Table P4.19. Determine the magnitudes of the inertia J
LAB. and the damping coefficient b. Interpolate if necessary. Re-
port your results in the correct SI units.

Fig. P4.18  a Translational


mechanical system, b Velocity of
a b
x,v 4
the ideal massless and rigid bar

K 3

v(t)
F(t) 2
m
___
b sec
1

Ideal massless 0
and rigid bar
0 2 4 6 8 10
t, sec

Table P4.18   Velocity data Time (sec) Velocity Time (sec) Velocity Time (sec) Velocity
(m/sec) (m/sec) (m/sec)
0.0 4.000 3.5 0.563  6.5 0.105
0.5 3.023 4.0 0.426  7.0 0.079
1.0 2.285 4.5 0.322  7.5 0.060
1.5 1.727 5.0 0.243  8.0 0.045
2.0 1.305 3.5 0.563  8.5 0.034
2.5 0.986 5.5 0.184  9.0 0.026
3.0 0.745 6.0 0.139  9.5 0.020
10.0 0.015
268 4  Mechanical Systems

Fig. P4.19  a Rotational


mechanical system, b Angular
a b
1.5
velocity of the mass moment of
inertia J
T(t) 1.0
J Ω(t)
rad
____
b sec 0.5
Hydrodynamic
Bearing

0.0
0 1 2 3
t, sec

Table P4.19  Angular Time (sec) Ω1g(kRPM) Time (sec) Ω1g(kRPM) Time (sec) Ω1g(kRPM) Time (sec) Ω1g(kRPM)
velocity data
0.00 0.00 0.65 1.03 1.30 1.21 1.95 1.24
0.05 0.16 0.70 1.06 1.35 1.22 2.00 1.24
0.10 0.29 0.75 1.08 1.40 1.22 2.05 1.24
0.15 0.41 0.80 1.10 1.45 1.22 2.10 1.25
0.20 0.52 0.85 1.12 1.50 1.23 2.15 1.25
0.25 0.61 0.90 1.14 1.55 1.23 2.20 1.25
0.30 0.69 0.95 1.15 1.60 1.23 2.25 1.25
0.35 0.76 1.00 1.16 1.65 1.23 2.30 1.25
0.40 0.82 1.05 1.17 1.70 1.24 2.35 1.25
0.45 0.87 1.10 1.18 1.75 1.24 2.40 1.25
0.50 0.92 1.15 1.19 1.80 1.24 2.45 1.25
0.55 0.96 1.20 1.20 1.85 1.24
0.60 1.00 1.25 1.21 1.90 1.24

References and Suggested Reading Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice-Hall, Upper Saddle River
Boyce MC, Arruda EM (2000) Constitutive models of rubber elasticity: Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
a review. Rubber Chem Tech 73:504–523 dynamics. Addison-Wesley, Reading
Budynas RG, Nisbett KJ (2011) Shigley’s mechanical engineering Timoshenko SP, Gere SM (1997) Mechanics of materials, 4th edn.
design, 9th edn. McGraw-Hill, New York PWS-Kent, Boston
Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs
Fluid, Electrical, and Thermal Systems
5

Abstract
Fluid and electrical systems are important means of transmitting, transforming, and con-
verting power in mechanical design. Fluid and electrical systems are networks, naturally
represented by the linear graph method. Everyday experience with fluid flow provides a set
of analogies to aid in understanding electrical systems. The only significant discrepancy
of the analogies is the network representation of fluid and electrical capacitances. Thermal
system phenomena are presented without entropy generation or flow. The focus of these
presentations is to model the removal of waste heat from mechanical, electrical, and fluid
systems. As a result, heat engines are absent. Heat transfer is modeled instead as a potential-
driven flow, similar to low velocity (seepage) fluid flow.

5.1  Fluid Systems the mixture of gases found in air. Vacuum systems are used
in automobile design, exploiting the difference in pressure
Fluids, both liquids and gases, are used in machine design between the intake manifold and the atmosphere.
to transmit power, store energy, and actuate mechanisms. Water and wind power systems are ancient technologies.
Fluids have mass and are compressible. Thus, they store The invention of steam power can be attributed to the In-
both kinetic and strain energy. Further, fluids are generally dustrial Age. Hydraulic and pneumatic systems have long
confined within containers such as tubes, pipes, tanks, and been a means to apply force over a limited displacement,
cylinders. An additional mode of energy storage is the work with vehicle brakes being the most common example. A hy-
that is done by those forces, which are exerted by fluids on draulic piston/cylinder, a linear hydraulic motor, is also the
their containers. One important form of energy storage in most familiar hydraulic component, since it is often visible.
low pressure systems is the elevation of liquid against grav- Hydraulic systems have gained popularity in machine de-
ity. Mechanical engineers work with many different types of sign, due to the diminishing cost of hydraulic components,
fluid systems. Our emphasis will be on fluid power systems, resulting from increased use of computer numerical control
either pneumatic or hydraulic. These analyses, however, can (CNC) machine tools. Rotational hydraulic motors are now
be applied to other types of fluid systems, such as water dis- used to eliminate shafts and gear sets (rotational mechani-
tribution and fire protection systems, as well. cal power transmission elements), to reduce cost and in-
Fluid power systems use both liquids and gases as the crease design flexibility. High-pressure pneumatic systems
working fluid. “Hydraulic” systems typically use oil as are unusual, but should become more common, if current
the working fluid. They operate at pressures ranging from trends continue.
750 psi to 15,000 psi. Fixed industrial equipment typical-
ly operate at pressures from 2,500 to 3,000 psi. A pressure
of 5,500 psi is common in mobile construction equipment. 5.1.1  Fluid Power Variables
Pneumatic pressures range from conventional low-pressure
pneumatic systems of 100–150 psi to high-pressure pneu- Mechanical engineers have an intuitive understanding of the
matic systems with pressure up to 7,000  psi. Some “pneu- basics of fluid systems, because water flow is so familiar.
matic” systems use single gases, such as nitrogen, rather than We know that water flows in the direction of the pressure

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_5, © Springer Science+Business Media New York 2014 269
270 5  Fluid, Electrical, and Thermal Systems

Pipe, tube, or hose a b

Q(t) Q(t)
p p v v
1 2

Fig. 5.1   Pressure nodes p1 and p2 and the positive direction of fluid
volume flow rate Q through a pipe, tube, or hose
Laminar flow Slug flow
Parabolic velocity profile Uniform velocity profile

drop in the pipe. Knowledge of the pressure at one location Fig. 5.2  a Parabolic velocity profile of fully developed laminar flow,
(node) is insufficient information to determine the direction b Uniform velocity profile of slug flow
of flow. We must be able to calculate the difference in pres-
sure between two nodes to determine the flow direction with   m3 
any accuracy. Q =  (5.3)
Pressure drop in the direction of flow, p12, is the across  sec 
variable in fluid systems. We will measure flow as Vol-
ume Flow Rate, Q. Q is the through variable, which flows A cubic meter is an awkwardly large volume to use in ma-
through the pipe, Fig. 5.1. Fluid power is the product of the chine design. For a sense of scale, the mass of a cubic meter
across variable times the through variable of water is 1,000 kg per metric ton. So, why use SI? In US
customary units, volume flow rate is expressed in units of
 P = p12 Q (5.1) gallons per minute, abbreviated as GPM. However, US cus-
tomary units for pressure, whether as pounds per square inch
where p12 is the pressure drop in the direction of the flow, (psi) or pounds per square foot (psf), are not derived from
and Q is the volume flow rate. gallons. One must convert gallons to either cubic inches or
We will simplify the flow of fluid through pipes, tubes, cubic feet, for most calculations. For example, in the USA,
and hoses by modeling the flow as “slug” flow, Fig. 5.2. In fluid power pumps are described by their flow rate in GPM
slug flow, the fluid velocity profile is uniform. The veloc- and outlet pressure in psi. In order to calculate a pump’s
ity of real fluid flowing down a pipe varies with radial and power in US customary units, one must convert gallons to
longitudinal positions. Factors include the fluid’s properties, cubic feet and psi to psf, in order to yield power in units of
the velocity of the flow, as well as the pipe’s geometry and foot pounds per minute. The calculation is much more direct
surface roughness. Assume the velocity is uniform across the in SI units
pipe, to visualize the fluid flowing down the pipe as a cylin-
drical slug, with velocity v. Although slug flow is a gross 
 3

simplification for some types of flows, it is a very accurate [ P ] = [ pQ ] =  N2   m  =  N·m  =  Joule  = Watt
 m   sec   sec   sec 
approximation for turbulent flow, which is the typical flow
in fluid power systems. Turbulent flow occurs when fluid (5.4)
passes through pumps with changes in pipe or tube diame-
ters. Such flow also occurs at the sites of restrictions, such as SI also eliminates the need to calculate mass from weight.
valves or the orifices of hydraulic cylinders. These compo- We can dispense with having to remember the gravitational
nents are spaced closely enough in typical fluid power sys- constant in a variety of combinations of US customary units.
tems to prevent the flow from ever developing the parabolic Finally, in the case of a hybrid system with an electrical sub-
velocity profile characteristic of laminar flow. system, conversion to SI is required, because electrical units
An assumption of slug flow (with its uniform fluid ve- of volts and amperes are SI.
locity profile) greatly simplifies calculation of volume flow Most mechanical engineers understand low-frequency
rates. The volume flow rate, Q, which passes through a electrical systems by the fluid analogies. Fluid pressure is
cross-section of the pipe with area, A, at time, t, is analogous to electrical voltage. Fluid volume flow rate is
analogous to electrical current. The only significant differ-
 Q(t ) = Av(t ) (5.2) ence between network representations of electrical and fluid
systems is that fluid capacitors must be referenced to atmo-
We will work in SI units, with volume flow rate expressed in spheric pressure, or some other “ground” pressure, whereas
cubic meters per second. electrical capacitors do not.
5.2  Fluid Elemental and Energy Equations 271

Pressure Pressure Operating Range


+/-∆p R
Drop p12 Drop, p
1
Operating Point
R p0 Q 0, p0
Operating Range
1 +/-∆Q

Q0
Volume Flow Rate, Q R Volume Flow Rate, Q

Fig. 5.3   Linear fluid resistance, p12 = RQR Fig. 5.4   Linearization of a non-linear fluid resistance

If we need to improve the accuracy of our model of fluid


5.2  Fluid Elemental and Energy Equations resistance, we will linearize the model about an operating
point. The process, described in Sect. 4.2.4.1, approximates
Fluid systems are mechanical systems, since they obey New- the pressure–volume flow rate curve by a straight-line tan-
ton’s laws. They store energy as kinetic energy (energy of gent of the curve at the “operating point”of the flow rate, Q0,
motion) or strain energy (energy of elastic deformation). and pressure drop, p0, which is the expected operating condi-
Kinetic energy is dissipated as heat through viscous fric- tion of the system Fig. 5.4. If the pressure–volume flow rate
tion, which is lost from the system. One difference between curve is available as a mathematical function, the function is
fluid systems and our treatment of translational mechanical linearized as a two-term Taylor series approximation, which
systems is that we will represent gravity as potential energy is equivalent to the graphical procedure.
storage in fluid systems, rather than as a force source. The tangent to the operating point, Q0, p0, is

 p12 = R ∆QR + p0 (5.7)


5.2.1  Fluid Energy Dissipation: Fluid Resistance
where the term, ΔQR, is a deviation above or below the op-
Energy is dissipated as heat in fluid flow by viscous shear erating point, Q0. Notice that this relationship cannot be in-
in components, such as filters, orifices, and valves, due to verted to calculate QR, since the expression does not contain
changes in diameters and direction of pipes and the rough- Q0. The expression yielding QR is
ness of their inner surface. In the simplest case, the pressure
drop in the direction of the flow is proportional to the flow 
∆ p12
rate, Fig. 5.3. QR = + Q0 (5.8)
R
 p12 = RQR (5.5)
5.2.2  Kinetic Energy Storage Fluid Inertance
This relationship has the same form as electrical resistance,
v12 = RiR. Fluid has mass. Consequently, flowing fluid possesses ki-
Unfortunately, the simplest case of fluid resistance does netic energy. We will derive the Elemental and Energy equa-
not describe many real-world situations. It is an accurate tions for the dynamic element, which represents mass in a
model for slow flows, such as flow of oil through a filter, fluid system, by expressing the corresponding equations for
or water permeating through soil. The model loses accuracy, a translational mechanical system in terms of the fluid power
however, as the fluid volume flow rate increases, because variables, p and Q. Consider the mass of fluid with density
fluid viscosity is not constant. Viscosity is a function of the ρ in a pipe of length L and cross-sectional area A, shown in
velocity of the fluid flow, more accurately described by a Fig. 5.5. The cylindrical lump of fluid has mass M
“power law” viscosity function
 M = ρ AL (5.9)
 µ ( v ) = µ0 v n (5.6)
To accelerate or decelerate, a force must act on the mass of
We must use linear elemental equations to produce linear fluid. Substituting M = ρ AL for the mass of the fluid, and
system equations. The simplest model for flow through a expressing acceleration as the derivative of velocity with re-
fluid resistance is illustrated with a pressure drop across the spect to time
resistance that is linearly related to the volume flow rate, Q.
272 5  Fluid, Electrical, and Thermal Systems

Length L t1 t2

Density ρ Velocity
Q(t) v Q(t)
Volume
Area A p p
1 2
Fig. 5.5   Cylindrical fluid mass, a fluid “lump” Area A
Fig. 5.7   Volume flow rate in slug flow
F1 = p1 A F2 = p2 A
M = ρAL
p p This is the elemental expression for fluid inertia, the element
1 2
which stores kinetic energy in a fluid system. The quantity
Fig. 5.6   Net pressure force acting on a fluid lump
 ρL
I= (5.13)
A
dv dv
F = Ma → F=M → F = ρ AL
dt dt is called fluid inertance.
Note an apparent contradiction in the definition. Shouldn’t
The force acting to accelerate the mass of fluid is the net increasing the cross-sectional area of the pipe increase the
pressure force between node 1 and node 2, Fig. 5.6, inertia of the fluid moving down it, by increasing the mass
of fluid in a length of pipe? It does because fluid inertance
FNet = F1 − F2 = Ap1 − Ap2 = A( p1 − p2 ) = Ap12 I is not inertia. Fluid inertance is derived by using volume
Pressure flow rate QI in place of fluid velocity v. Increasing the cross-
sectional area of a pipe results in decrease in the fluid veloc-
The power variable, Q, volume flow rate, is related to the as- ity for a given volume flow rate QI. The acceleration of the
sumed uniform velocity of the flow. The volume of fluid flow- fluid is also decreased proportionally. Hence, fluid inertance
ing at velocity v, that passes through a plane, normal to the I is inversely proportional to the cross-sectional area A of
flow of area A, in a period of time, Δt, is volume ≡ V = Av ∆t , the pipe.
Fig. 5.7. Using the above expressions for the mass of fluid in a
Hence, the volume flow rate, QI, is length of pipe and the uniform velocity in terms of QI, we
can write the expression for the kinetic energy stored in a
 Volume flowing fluid
= QI = Av (5.10)
t
1 ρA 2
E Kinetic = QI
The velocity, v, expressed in terms of volume flow rate, Q, is 2 A

QI  1 2
 EI = IQI (5.14)
v= (5.11) 2
A

Now write F = ma, in terms of p12 and QI where fluid inertance I plays the role of mass in the equation.

dv
FNet = Ma → p12 A = ρ AL
Pr essure dt 5.2.3 Pressure-Based Energy Storage Fluid
Capacitance
Q 
d I
 A ρ L dQI
p12 A = ρ AL → p12 = Fluid systems can store energy in the elastic deformation of
dt A dt the pipe or structure containing the fluid under pressure, in
dQI the elastic deformation of the fluid itself, and by raising fluid
 p12 = I (5.12) against gravity. All three of these energy storage modes lead
dt
to elemental and energy equations with the same functional
form as an electrical capacitor. The parameter used in these
5.2  Fluid Elemental and Energy Equations 273

equations is given the analogous name, fluid capacitance, pg


and the same symbol, C, as electrical capacitance.
The elemental equation for a fluid capacitance relates the Area A
volume flow rate into fluid capacitance, QC, to the rate of Density ρ
change of pressure in the capacitance,
x,v
H
 dp2 g p1g A slug p2g A slug
QC = C (5.15) p1 p2
dt

where C is the fluid capacitance. Note that pressure is mea-


sured relative to atmospheric or ground pressure in the el- Fig. 5.8   A standpipe-type fluid capacitance showing the work done
emental equation of a fluid capacitor. The voltage in an pushing a lump of fluid against the pressure in the capacitance
electrical capacitor is measured across the capacitor. As ca-
pacitance C increases, more fluid volume in or out of it is 
required to change pressure at a given rate. W�
= FNet x (5.16)
Pressure
In general, both the fluid and the components which con-
tain the fluid are elastic and can store strain energy. These
strain energies are calculated independently. First consider This can be expressed in terms of the power variables of a
the fluid to be incompressible, and calculate the strain energy fluid system, pressure, and flow rate. Differentiate the me-
stored in the components containing the fluid. Then, consid- chanical work:
er those components as rigid and calculate the strain energy
stored in the fluid. Gravitational potential energy storage is dW � d d
usually insignificant in hydraulic systems in machines be- dt
≡ P�
=
dt
( )
p12 Aslug x = p12 Aslug ( x ) = p12 Aslug v
dt
cause the hydraulic pressures (on the order of 2,500 psi) are
so much larger than the gravitational pressures. Gravitational Then, use the definition of volume flow rate:
potential energy storage is important in water systems, both
within a building and in a municipal distribution system. QC = Aslug v

5.2.3.1  Gravitational Potential Energy We confirm that the rate of mechanical work being done to
Gravity was not included as an energy storage mode in trans- push fluid into the fluid capacitance is fluid power:
lational mechanical systems, because it only acts in one di-
rection. It is easier to include gravity as a force source which  dW �
≡P �
= p12 Aslug v = p12 QC (5.17)
acts on mass. Gravity modeled as a force source still pro- dt
vides energy storage, because power can flow both into and
out of an ideal source. A car battery is a good example of Derive the fluid capacitance, C, of a standpipe. Assume the
a source which can either deliver (source) or accept (sink) fluid is incompressible, and the standpipe is rigid. The cross-
power. sectional area of the standpipe is A. The density of the fluid
In fluid systems, gravity is represented as pressure-based is ρ. Conservation of mass dictates the mass of fluid, which
energy storage, where the pressure is created by the height flows into the port, is stored in the standpipe. Express mass
of a column of fluid. The most familiar example of such a conservation in terms of mass flow rate:
fluid capacitor is a municipal water tower. A water tower in
the form of a cylindrical tank without legs is called a “stand-  d ( MassStored )
Mass Flow Rate In = (5.18)
pipe.” dt
The net pressure force acting on a slug of fluid to push
 dH
it into the standpipe results in mechanical work being done QC ρ = ρ A (5.19)
on the fluid. The net pressure force acting on the fluid slug, dt
Fig. 5.8, is
Introduce the power variable, pressure, by recognizing that
FNet
Pressure
( )
= p1g − p2 g Aslug = p12 Aslug the pressure at the bottom of the tank, relative to atmospheric
(ground) pressure, is the weight of a column of fluid of unit
area, where g on the right side is the acceleration of gravity:
The mechanical work done to push the fluid slug into the
standpipe is  p2 g = ρ gH (5.20)
274 5  Fluid, Electrical, and Thermal Systems

v2 p1g x, v

QC K p1g A
p1 v3 FK

Piston Area A pg
Fig. 5.10   Force balance on piston of spring-loaded fluid accumulator
Fig. 5.9   A spring-loaded fluid accumulator
Nitrogen Charging Valve

Divide and multiply the right side of the mass conservation


equation by g N2 Gas

g dH A dH
QC ρ = ρA → QC ρ = ρg Pressure Cylinder
g dt g dt
Hydraulic Bladder
Oil
  A  dp2g Low pressure
QC =   (5.21) valve
 ρ g  dt
Hydraulic Lines
The coefficient term multiplying the time derivative of pres-
sure fluid is the fluid capacitance, Fig. 5.11   A bladder-type hydraulic accumulator. The compressibility
of the nitrogen gas in the bladder stores energy
 A
C= (5.22)
ρg
  A2  dp1g
QC =   (5.23)
 K  dt
5.2.3.2 Energy Stored in the Elastic Strain
of the Fluid Container
Consider a fluid system component which consists of a The capacitance of a fluid accumulator is
spring-loaded piston in a cylinder. This device is called a
fluid “accumulator”in hydraulic systems, a “pressure relief  A2
C= (5.24)
tank” in hot water systems, or a fluid capacitor in system K
dynamics. Assume the fluid is incompressible, and the con-
tainer is rigid, except for the spring. The volume of the cyl- The spring used in a fluid accumulator may be a conven-
inder which contains the spring is vented to the reference tional compression spring. An alternative design is to use a
(ground) pressure, usually atmospheric pressure. The only diaphragm, rather than a piston, and high pressure gas, typi-
force acting on the right side of the piston is the spring force, cally nitrogen, as the “spring,” Fig. 5.11. A component may
FK. There are two different types of across-variable nodes act as an unintentional fluid capacitor, if it exhibits sufficient
identified on the schematic, pressure, and velocity. elastic deformation when filled with a pressurized fluid. A
The volume flow rate, QC, is related to the velocity of the common example is a garden hose. Many garden hoses ex-
piston relative to the end of the cylinder, v12. pand significantly when pressurized, causing a delay in de-
livery of water from the hose. A significant amount of strain
QC energy is stored in the hose, even under the relatively low
v23 =
A pressure of a residential water system of approximately
50 psi. The energy is released when the tap is closed, allow-
A free body diagram of the forces acting on the piston, ing water to continue to flow out of the hose.
Figs.  5.9 and 5.10, leads to an equilibrium equation, which The energy equation for this fluid capacitor can be de-
relates the net pressure force, p1g A, of the fluid to the force rived from the energy storage equation for the spring
in the spring, FK.
( Ap )
2
The elemental equation for this fluid capacitor can be de- FK2 1g 1 A2 2
rived from the elemental equation for a spring. EK = → EK = → EC = p1g
2K 2K 2 K
dFK
= Kv23 →
(
d p1g A )=KQ C 1
dt dt A  EC = C p12g (5.25)
2
5.2  Fluid Elemental and Energy Equations 275

5.2.3.3 Energy Stored in the Compression


of the Fluid QC
Positive stress is tensile in solids. The analog of stress in flu- p1
ids is pressure where positive pressure is compressive. Liq-
uids in fluid power systems are kept under positive pressure pg
to prevent “cavitation,” or local vaporization, of the fluid.
The problem is damage to machinery from the stress created
by the collapse of the bubbles, when they are carried by the Rigid container, volume Vol
flow from a region of negative pressure into positive pres- Compressible fluid, bulk modulus β
sure. Bubbles and dissolved air in hydraulic oil are always
problems because they greatly increase the compressibility Fig. 5.12   Increasing fluid pressure within a rigid container stores strain
of the fluid. energy in a compressible fluid
Strain in a fluid is defined as volumetric strain, a fraction-
al change in volume. Express volumetric strain by its effect
of increasing the density of the fluid, as a fractional change Expressing dρ in terms of the bulk modulus, β, using
in density dρ
dp = β
ρ yields
 dρ
Volumetric Strain ≡ (5.26) ρ dp1g
ρ ρQC = Vol
β dt
In the simplest case, a linear relationship exists between the
  Vol  dp1g
applied pressure and the volumetric strain QC =  (5.28)
 β  dt
 dρ
dp = β (5.27)
ρ In this expression for fluid capacitance,

where β is the “Bulk Modulus.” This relationship has the  Vol


C= (5.29)
same form as Hooke’s Law for linear strain in a solid β

 σ = Eε (1.3)
5.2.4  Fluid System Sources
The similarity is stronger than it may appear, because
Hooke’s law can be written as Sources are power supplies, in which one of the two power
variables of the system is either known or controlled. Fluid
dL system sources are either pressure sources or volume flow
dσ = E rate sources. Pressure sources are more common than vol-
L
ume flow rate sources. Pressure sources occur naturally,
wherever there is a large reservoir at an elevated pressure,
To calculate the energy stored in the deformation of the fluid, relative to atmospheric pressure. Volume flow rate sources
assume (1) the container is rigid, and (2) more fluid can be do not occur naturally. Volume flow rate sources are created
packed into the fixed volume, Vol, of the container by com- by (1) placing a pressure source under feedback control to
pressing the fluid at increased pressure, Fig. 5.12. vary pressure as the system’s input resistance varies, or (2)
Conservation of mass requires that fluid flowing into a by placing a positive displacement mechanical pump under
container of fixed volume, Vol, must be stored in the tank, by velocity control. A positive displacement pump’s geometry
increasing the density of the fluid. displaces a certain volume of fluid every cycle. Pumps and
hydraulic motors are addressed in Chap. 6.
d ( MassStored )
Mass Flow Rate In =
dt 5.2.4.1  Pressure Sources
Fluid pressure sources exist in nature and in technology.

ρQC = Vol A natural fluid pressure source is essentially a large stand-
dt pipe-type fluid capacitance, that is, a fluid reservoir with
276 5  Fluid, Electrical, and Thermal Systems

Fig. 5.13  a An industrial station-


ary fluid power unit driven by a
a b
Three phase AC motor
three-phase AC motor. The pump
is below the motor in the tank. High pressure
Breather cap. Filtered Discharge
b A fluid power unit modeled vent to atmosphere Vent to P(t)
as a pressure source represented atmosphere Pump
schematically
High pressure
Atmospheric
Output from pump
pressure Return
Low pressure
Return to tank Fluid
Reservoir
Tank. Hydraulic
fluid reservoir

Stationary fluid power unit Fluid power unit modeled


as pressure source p(t)

sufficient height to create pressure, and sufficient volume for with hydraulic fluids for a number of purposes: to protect the
that height change negligibly when there is flow out of or metal surfaces of the pumps and actuators in the system from
into the reservoir. Earth’s atmosphere is a source of atmo- wear and corrosion; to limit foaming of the oil; to extend the
spheric pressure. An intake pipe in a lake or an ocean draws usable life of the fluid; and to change viscosity.
water from a pressure source. Biological systems have vari- Air entrainment is a significant concern. “Cavitation” is
ous means of creating pressures or vacuums. Air and blood a phenomenon, whereby a local area of low pressure with-
flow from the pumping action of muscles. Water is drawn to in a flow permits air to come out of solution, and form a
great height in trees, due largely to capillary tension. bubble. Extremely high stress results when such a bubble
Artificial pressure sources are either fluid capacitances or collapses. Cavitation erodes hardened steel. The particle of
pumps, as in liquids, and compressors, as in gases. A tank of steel introduced into the hydraulic fluid, in turn, damages the
nitrogen gas, a fluid capacitance, is a pressure source. The system. Hydraulic systems require filtration. The degree of
pressure source of a hydraulic automobile brake is a piston filtration, in terms of the size of particle removed, depends
in a cylinder. The pressure source for air-powered tools is an on the nature of the system. Filtration for servo-hydraulic
air compressor. Industrial hydraulic systems are powered by systems with actuators under feedback control has the most
pumps. These devices are closer to ideal pressure sources, stringent requirements to protect servo valves and other pre-
when the flow rate drawn from them is small, relative to the cision components.
capacity of the device. There is finite resistance to any flow. Foaming of hydraulic oil is also problematic. Foam floats
The greater the flow rate, the greater the effect of the fluid on the surface of a fluid in a reservoir and does not flow, if
resistances in the device, and the larger the pressure drop the discharge (intake to the pump) is located below the sur-
within the source. face, as one can observe in the shower.
A stationary fluid power unit, as in Fig. 5.13, provides Fluid power units are sized, in terms of the volume of
hydraulic power for industrial machinery. Water served as fluid delivered in a unit time at a specified pressure. US cus-
the original hydraulic fluid, but its corrosive properties and tomary units for fluid power are gallons per minute and psi,
lack of lubricity (lubricating ability) led to its replacement which are expressed inconsistently. Rather than convert gal-
by mineral oils and emulsions of water and mineral oil. Hy- lons to cubic inches, it is best to convert both the volume and
draulic fluids most commonly used in industrial and mobile the pressure to SI units, compute power in watts, and then
equipment are mineral oils. A phosphate ester is used for convert to the units needed to present the result.
aircraft hydraulic systems due to the flammability of the hy- The pump of a stationary fluid power unit is typically
draulic oil. Silicone-based hydraulic fluid is used for its non- powered by a three-phase AC motor. In order to maintain
flammability. An emulsion of water, mineral oil, and glycol pressure while the power unit is in stand by (not discharg-
is another variety of hydraulic fluid. ing fluid to the external system), there is a high-resistance
The properties, including density and bulk modulus, vary internal hydraulic bypass circuit from the pump to the fluid
with the type of fluid and, importantly, with the amount of reservoir, Fig. 5.14. The reservoir is vented to the atmo-
air entrainment. The density and modulus of phosphate ester sphere through a filter in the fill cap, called a “breather
and silicone-based hydraulic fluid are significantly lower cap.”The filter is intended to keep foreign matter, especially
than that of oil-based hydraulic fluid. Although we use a abrasive particles, out of the hydraulic fluid. The fluid res-
constant value for the bulk modulus, β, in calculations, it is ervoir is box shaped and sized so that the “residence time,”
actually a function of pressure. Various additives are mixed the period of time the fluid stays in the reservoir during
5.3  Linear Graphs of Fluid Systems 277

High pressure
Discharge
Vent to Q(t) C
atmosphere Pump
R internal R1
High resistance P(t)
vent to Pump
internal by-pass atmosphere
R2
Fluid Atmospheric Fluid
Reservoir pressure Return Reservoir

Fig. 5.14   Fluid power unit, modeled as a volume flow rate source. A Fig. 5.15   Fluid system consisting of a pump modeled as a pressure
high-resistance internal bypass between the pump discharge and the source, two fluid resistances, and a fluid capacitance
tank allows the pump to run, when the system does not accept flow

ment. Fluid capacitances contain one pressure node inside.


operation, is approximately two minutes. The residence The other pressure node is the “ground” pressure, usually at-
time allows the fluid to “de-gas,” i.e., have air bubbles rise mospheric pressure. The most common element to overlook
to the surface. is a fluid inertance. Don’t forget that F = ma applies to fluids.
Mobile fluid power units have the same functional compo- There must be a pressure drop, to accelerate or decelerate
nents. Power for the pump of a mobile unit is usually drawn fluid flow in a pipe. Consequently, there are pressure nodes
from the shaft of the internal combustion engine. Some elec- at either end of a fluid inertance.
trically power pumps, however, are used in smaller systems.

5.2.4.2  Volume Flow Rate Sources 5.3.1  Example Fluid System Linear Graphs
Volume flow rate sources do not exist in nature. There must
be technology involved, because the pressure to drive the 5.3.1.1  Example One
flow into the system must be varied to maintain a given flow The model of the fluid system shown in Fig. 5.15 consists
rate. Artificial volume flow rate sources accomplish this, by of a pressure source which discharges into fluid resistance
varying the pump speed or the resistance of the internal by- R1. Fluid resistance R1 connects to the junction of a fluid ac-
pass with feedback control, Sect. 9.3.7. cumulator (capacitance) C and fluid resistance R2. Fluid re-
sistance R2 discharges into the pump’s reservoir, completing
the fluid circuit.
5.3  Linear Graphs of Fluid Systems Nodes of distinct pressure are identified in the schematic
Fig. 5.16a. The ideal pipe between the discharge port of the
The reduction of fluid systems to a system equation follows pump and fluid resistance is node 1, the discharge pressure
the same procedure as for mechanical and electrical systems. of the pump. The ideal pipes connecting resistances R1 and
The first step is most important. Find the nodes of distinct R2 and the capacitance C are all node 2, the pressure of the
values of the across variable, pressure. There are always two fluid capacitance. There is a pressure difference across resis-
nodes associated with each element. Always put a node at tance R2, but the pipe connecting R2 to the fluid reservoir is
either end of pressure or flow sources, fluid resistors, and ideal. Ideal pipes have fluid properties and, therefore, have
fluid inertances. One node of a fluid capacitor must be atmo- no pressure change along them. The fluid reservoir is vented
spheric pressure (ground). The other node is the maximum to the atmosphere. Therefore, the pressure in the ideal pipe
pressure in the fluid capacitor, which may occur anywhere between the reservoir and resistance R2 is atmospheric pres-
within a fluid accumulator, or at the bottom of a water tank sure, “g.” Check for two nodes associated with each energet-
or standpipe. ic element. Create the linear graph, by first drawing and la-
Fluid system schematics contain ideal pipes, tubes, and beling the nodes of distinct pressure. Then, add the branches,
hoses, which have no energetic properties. Ideal pipes are which represent the elements between them. The branch for
analogous to ideal conductors in electric circuits, as well as a fluid capacitance is dashed after the arrowhead to indicate
massless and rigid rods or shafts in mechanical schematics. that the fluid cannot flow to ground through the capacitance.
Do not assign fluid element properties to an ideal pipe. If The linear graph is shown in Fig. 5.16b.
a pipe, tube, or hose has energetic properties, they will be
indicated on the schematic. Every fluid element must have 5.3.1.2  Example Two
two pressure nodes associated with it. Fluid resistance and The fluid system shown in Fig. 5.17 consists of a pressure
fluid inertances have pressure nodes at either end of the ele- source in the form of a fluid reservoir, a ball valve on the
278 5  Fluid, Electrical, and Thermal Systems

Fig. 5.16  a Nodes of distinct R1


pressures. b Linear graph of the a b
fluid system 1 2
C
2
1 R1
P(t)
vent to Pump
p(t) C R2
atmosphere
R2
Fluid
Reservoir g
g

C g C
Ball valve Ball valve
Fluid Closed Open
Reservoir Fluid 3
I Reservoir
R 1 R2 I

Fig. 5.17   Fluid system with a fluid reservoir acting as a pressure


Fig. 5.18   Fluid system with nodes of distinct values of the across vari-
source, and fluid resistance, inertance, and capacitance
able, pressure

discharge of the reservoir, a fluid resistance R, a fluid iner- R I


tance I, a fluid capacitance C, and second ball valve to drain 1 2 3
the system. A ball valve can be modeled as having negli-
gible resistance, when it is fully open, because the flow path
through the valve is straight, circular in cross-section, and of
the same diameter as the pipes or tubes it is connected to. In p(t) C
this system, we will model open ball valves as having zero
fluid resistance.
Identify nodes of distinct pressure on the schematic,
g g
Fig. 5.18. The pressure at the discharge port of the fluid res-
ervoir extends to the ball valve, when it is closed, and to the Fig. 5.19   Linear graph of the fluid system shown in Figs. 5.17 and 5.18
fluid resistance R when the ball valve is open. As we are in-
terested in the dynamic response when the ball valve is open,
let’s take a look at the latter case. Pressure drops across the sequently, its homogeneous, or natural, response can be os-
fluid resistance, R, in the direction of the fluid flow. Hence, cillatory. The flow can slosh back and forth. Whether water
pressure node 2 is immediately to the right of the fluid re- sloshes in a bucket depends on the relative amounts of power
sistance. The fluid inertance I, must have pressure nodes at dissipated in the fluid resistance and stored in the fluid iner-
either end. Node 2 is the pressure node for the left end of the tance and capacitance.
fluid inertance. Node 3 for the right end of the inertance is Consider the initial condition of the ball valve on the res-
in the fluid capacitance, which is connected to the fluid iner- ervoir as closed and fluid level in the capacitance below that
tance by an ideal pipe. of the fluid reservoir, shown in Fig. 5.17. If the ball valve
Create the linear graph by drawing and identifying the is opened quickly, the system experiences a step change in
nodes. Next, draw the branches between them, per Fig. 5.19. pressure at node 1. How much flow is there in the system
This system is the fluid analog of the series RLC elec- at time, t = 0+, the instant after the valve is opened? Zero.
tric circuit of Sect. 1.5.2.2. Most mechanical engineering The mass of fluid in the fluid inertance must be accelerated
students find the fluid analogy with electric circuits helpful, by the net pressure force between nodes 2 and 3. If the fluid
particularly as an aid for understanding electrical inductance, resistance is relatively low, then the flow through the iner-
the electrical analog of fluid inertance. This fluid system is tance can build up to a torrent, storing a significant amount
of second order, because the fluid inertance and the fluid of kinetic energy. On the other hand, if the resistance is rela-
capacitance are independent energy storage elements. Con- tively large, the volume flow rate in the inertance will not be
5.3  Linear Graphs of Fluid Systems 279

R1 I
C 1 2 3
R1 I
P(t)
vent to Pump
atmosphere R2
p(t) R2 C
Fluid
Reservoir

Fig. 5.20   Fluid system schematic with a fluid power unit, modeled g g
as a pressure source, two fluid resistances, inertance, and capacitance
Fig. 5.22   Linear graph of the fluid system shown in Figs. 5.20 and
5.21

C
3 node for the left end of the fluid inertance, I. The third pres-
1 R1 2 I sure node is in the fluid capacitance, C. The fluid capacitance
P(t) is connected to the right end of the fluid inertance and the
Pump
vent to
atmosphere R2 top of fluid resistance, R2, by ideal pipes, so those locations
are also part of node 3. The pipe connected to the bottom of
Fluid
g resistance R2 has no properties and, thus, no pressure change
Reservoir
along it. The left end of the pipe discharges into the fluid
reservoir, which is at atmospheric or ground pressure, “g.”
Fig. 5.21   Nodes of distinct values of the across variable pressure indi- The same pressure node extends along the pipe to the bottom
cated on the schematic of resistance, R2.
Create the linear graph by first drawing and labeling
the nodes. Then, add the branches between their associated
as great, since the resistance restricts the flow through the nodes. Define a positive direction for the through variable,
inertance, and less kinetic energy will be stored. volume flow rate. Sources are special, in that they raise the
In the case of low resistance, when strong flow through value of the across variable, pressure, in the positive direc-
the inertance fills the capacitance up to the level of the res- tion of the through variable. There is a decrease, or drop, in
ervoir, and a net back pressure is established to decelerate the across variable, pressure, across resistances and energy
the flow through the inertance, there is sufficient kinetic storage elements in the positive direction of the through vari-
energy stored in the inertance to maintain the flow into the able, Fig. 5.22.
capacitance for some period of time. The fluid level in the
capacitance overshoots the reservoir level, which is the fu- 5.3.1.4  Example Four
ture equilibrium level, storing excess energy in the capacitor. A common error drawing linear graphs from a fluid sche-
The flow then reverses back into the reservoir. The oscilla- matic occurs when dealing with a pipe, tube, or hose with
tion repeats, until the fluid resistance dissipates the energy more than one energetic property. A linear graph is an ener-
stored in the first cycle. getic model. Elements in the linear graph represent the en-
ergetic properties of the system. If one component has two
5.3.1.3  Example Three energetic properties, then it will contribute two elements in
The fluid system shown in the schematic of Fig. 5.20 is driv- the linear graph. Consider a pipe that possesses both fluid
en by a pump, modeled as pressure source. The system has inertance and fluid resistance, per the schematic shown in
two fluid resistances, R1 and R2, a fluid inertance, I, and a Fig. 5.23.
fluid capacitance, C (fluid accumulator). The pipe extending from the discharge of the pump to
Identify locations of distinct values of the across vari- the fluid capacitance is “called out” as having two energetic
able, pressure, on the schematic and label them, Fig. 5.21. properties, fluid resistance and inertance. This is a common
Properties are identified by the element parameters on the circumstance. The properties of fluid resistance and fluid
schematic. If there is no property associated with a section inertance are both distributed along the entire length of the
of pipe, then it is ideal, and there is no pressure change along pipe. They are two distinct properties in the same physical
that section. The pipe segment from discharge port of the component. The pressure drop along the pipe is due to both
pump to the fluid resistance, R1, is pressure node 1. Pressure the energetic properties, resistance and inertance. How does
node 2, on the other side of fluid resistance R1, is also the one include both properties in a linear graph of the system?
280 5  Fluid, Electrical, and Thermal Systems

Long pipe with fluid Node 2 divides pressure drop p13


between the pipe’s distributed
resistance and inertance C resistance R p and inertance I C
3
P(t) 1 I 2 Rp
vent to Pump P(t)
atmosphere R vent to Pump
atmosphere R
Fluid
Reservoir
Fluid
Reservoir g

Fig. 5.23   Fluid system with two energetic properties, fluid resistance
and inertance, in one physical component, the long pipe Fig. 5.25   A “fictitious” pressure node in the pipe has no physical loca-
tion. It divides the pressure drop, p13, into one pressure drop due to fluid
resistance and another pressure drop due to fluid inertance
First, decide whether the two elements, which represent
the properties, are in series or in parallel. The same volume
flow rate, Q, flows through two elements in series, and the hoses or “lines”, and a fluid accumulator. The properties of
pressure drop is divided, Fig. 5.24a. The same pressure the components are called out as follows. The hose from the
drops across two elements in parallel, and the volume flow high-pressure “pump” or discharge port of the fluid power
rate, Q, is divided, Fig. 5.24b. In this example, the two prop- unit to the fluid accumulator has fluid resistance R1. The
erties exist in one pipe. The energy dissipated in the fluid fluid accumulator has capacitance C. The hose from the fluid
resistance, and the energy stored in the fluid inertance stem accumulator to the low-pressure “tank” or return port has
from the same volume flow rate, Q. The energetic proper- fluid inertance I and resistance R2.
ties of inertance and resistance are represented by elements Pressure nodes are first assigned to locations on the sche-
in series. matic, which have distinct values of the across variable, pres-
The total pressure drop along the pipe is due to both of sure. These locations are the pump discharge port, the tank
the energetic properties, resistance and inertance. The total return port on the fluid power unit, and the fluid accumula-
pressure drop must be divided by the addition of a node, i.e., tor. The hydraulic line from the accumulator to the return
node 2. Node 2 has no physical location, but it does have port has the properties of inertance and resistance. Therefore,
physical meaning. Whether the resistance is assigned the a “fictitious” node is needed to divide the total pressure drop
pressure drop from node 1 to node 2, or the inertance is as- between the two contributing properties, Fig. 5.28.
signed the pressure drop from node 2 to node 3, or vice versa, The linear graph of this system is shown in Fig. 5.29.
is irrelevant. Both properties are actually distributed along
the length of the pipe. As long as both elements in the lin- 5.3.1.6  Example Six
ear graph have the same volume flow rate, Q, through them, Let’s develop a model and draw its linear graph for a familiar
and their pressure drops sum to the pressure drop along the fluid system, a garden hose. Typical residential water pres-
pipe, p13, the energetic properties are properly represented, sure is 30–40 psi. We will model the residential water pipe
Fig. 5.26. as a pressure source. Assume the following conditions. The
garden hose is attached to the sill tap (valve at the house),
5.3.1.5  Example Five which is closed. A nozzle on the end of the hose is also
Linear graphs of fluid systems presented as more realistic closed. The hose is filled with water but remains depressur-
schematics are drawn by first reading the callouts on the ized, Fig. 5.30.
schematic to understand what energetic properties are in the If the sill tap valve is opened, with the nozzle still closed,
model. Pressure nodes are identified based on the properties we would expect to hear some water flow into the hose,
attributed to the components. which would swell slightly. The water, which flowed into
The fluid system shown in Fig. 5.27 consists of a fluid the hose, would be stored in a fluid capacitance created by
power unit modeled as a pressure source, two hydraulic straining the hose circumferentially. If we next open the

Fig. 5.24  a Fluid resistance and


inertance in series. b Fluid resis-
a b
tance and inertance in parallel R I R
1 2 3 1 3
I
5.3  Linear Graphs of Fluid Systems 281

Fig. 5.26   Equivalent linear


graphs for the system shown in
a R1 I
b I R1
Figs. 5.24 and 5.25 1 2 3 1 2 3

p(t) R2 C p(t) R2 C

g g g g

Fluid Power Unit Modeled as


Pressure Source p(t)
R1 I
1 2 3
Breather Cap (vent to atmosphere)

High Pressure Output

p(t) C R2
Hydraulic Line
Resistance R1

Low Pressure Return Fluid Accumulator


Capacitance C
g g g

Hydraulic Line Fig. 5.29   Linear graph of the fluid system shown in Figs. 5.27 and
Inertance I and 5.28
Resistance R 2

Sill Tap
Fig. 5.27   A fluid system with a fluid power unit modeled as a pressure
source. Both hydraulic lines have resistance. One has resistance and Hose
Nozzle
inertance. p(t) Inertance I
Resistance R N
Resistance RH
Capacitance C
Fluid Power Unit Modeled as
Pressure Source p(t)

Breather Cap (vent to atmosphere)


Fig. 5.30   Garden hose attached to sill tap, which is connected to the
residential water system, modeled as a pressure source
High Pressure Output
1
g
Hydraulic Line
Resistance R1
The fluid resistance of the nozzle is easy to model as a
Low Pressure Return Fluid Accumulator “lumped” parameter, because it is the nozzle’s only energetic
2 Capacitance C property. The hose has all three energetic fluid attributes of
inertance, resistance, and capacitance. All three properties
Hydraulic Line 3 are distributed along the length of the hose. When we pre-
Inertance I and viously modeled a pipe with both fluid resistance and iner-
Resistance R 2
tance, we found the properties to share the same volume flow
rate and, therefore, be in series. A fictitious node is needed to
Fig. 5.28   Schematic annotated with pressure nodes. The fluid capaci- divide the pressure difference between the two ends of a pipe
tance has a distinct pressure, as do the pumps outlet and return ports.
Node 3 is a fictitious node, which divides the pressure drops due to the
into the two pressure differences, one due to fluid resistance
distributed properties, inertance I and fluid resistance R2 and the other due to fluid inertance.
How is capacitance included in the model? The capaci-
tance of the hose is distributed along the length of the hose,
nozzle, we may or may not be aware of a brief transient due due to pressure-induced circumferential strain which increas-
to the inertance of the hose. The water then flows in steady- es the cross-sectional area of the pipe. (There is also axial
state, with the volume flow rate a function of the sill tap pres- strain that increases its length, which we will neglect.) We
sure, hose resistance, and nozzle resistance. must “lump” it into a single fluid capacitance element that is
282 5  Fluid, Electrical, and Thermal Systems

 Fig. 5.31  a Pressure nodes for


lumped parameter model of a
a Hose
b RH I
garden hose. b A linear graph Inertance I 1 2 3
Resistance RH
corresponding to the nodes of Sill Tap Capacitance C
5.31a Nozzle
Resistance R N
p(t) p(t) C RN
1 2 3 g

g g g

 Fig. 5.32  a Pressure nodes re- a Hose


quire the fluid capacitance node Sill Tap Inertance I
to be the average pressure value Resistance RH Nozzle
in the hose. b A linear graph cor- Capacitance C Resistance RN
responding to the nodes of 5.32a p(t)
1 g
2 3 4 5

b RH1 I1 RH 2 I2
1 2 3 4 5

p(t) C RN

g g g

located between a node in the hose and atmospheric (ground) value of the hose resistance and inertance in half. We can
pressure. We will begin by identifying where the hose ca- then place a resistance and inertance in series, both before
pacitance cannot be located in the model. The energy storage and after the node to which the capacitance is connected. The
equation for a fluid capacitance is advantage of the additional ­elements is that the capacitance
pressure node becomes the average pressure along the length
1 of the hose, when flow is steady and the inertance, therefore,
EC = Cpxg2 does not affect pressure, Fig. 5.32.
2

where subscript x indicates the unknown location for the


lumped fluid capacitance. Capacitance is pressure-dependent 5.4 Calculating Fluid Element Parameters
energy storage. The lumped capacitance cannot be located at from Fluid Properties and Geometry
the sill tap, modeled as a pressure source. If we were to con-
nect the capacitance to the pressure source, a step change 5.4.1  Fluid Resistance
in pressure would instantly fill the capacitance, requiring an
infinite volume flow rate and an infinite fluid power flow. Similar to friction in mechanical systems, it is generally
There must be a fluid resistance or inertance between a pres- more difficult to calculate fluid resistance than either fluid
sure source and a fluid capacitance, to limit the volume flow inertance or fluid capacitance, excluding capacitance due to
rate and power flow into the capacitance to finite levels. fluid compression. The difficulty stems from having to de-
Since there must be a fictitious node between the hose resis- fine the non-linear behavior of real fluids, such that we can
tance and inertance, we will connect the capacitance there. calculate values for a resistance independent of fluid velocity
The schematic with nodes and a corresponding linear graph in the linear form, p12 = RQR.
is shown in Fig. 5.31. In the ideal case of steady flow in straight and perfectly
The linear graph of Fig. 5.31b can be improved, at a cost smooth piping, the Reynolds number, Re = ρvD /µ , where v
of increasing the number of elements. We can divide the is the representative velocity of the flow and D is the diameter
5.4  Calculating Fluid Element Parameters from Fluid Properties and Geometry 283

Table 5.1   American Petro- SAE Low-shear rate High-shear rate


leum Institute (API) Engine Oil Viscosity Minimum kinematic viscosity Maximum kinematic viscosity Maximum viscosity
Classifications 2004. (Source rating
SAE J300). W (from “Winter”) mm 2 mm 2
cSt = at 100 °C cSt = atv 100 °C 1cP = 1mPa · sec at 150 °C
indicates the low temperature sec sec
viscosity
0W 3.8
5W 3.8
10W 4.1
15W 5.6
25W 5.6
20 9.3   < 9.3 2.6
30 5.6 < 12.5 6.9
40 12.5 < 16.3 2.9
(0W-40, 5W-40, 10W-40)
40 12.5 < 16.3 3.7
(15W-40, 20W-40, 25W-40, 40)
50 16.3 < 21.9 3.7
80 21.9 < 26.1 3.7

of the pipe, tube, or hose, is used to determine if the flow is surface of the hose depends on both the surface material, an
laminar or turbulent. By definition, the transient portion of elastomer (synthetic rubber), and design of the tube’s fiber
the dynamic response of a fluid system has changing flow reinforcement. The fiber reinforcement can impart waviness
rates and hence changing Reynolds numbers. Consequently, to a surface which varies with internal pressure. One rarely
a representative Reynolds number cannot be estimated, with- sees straight hydraulic hose. Hose is typically used to allow
out preliminary dynamic modeling to provide approximate flexure at pivots, where straightness of the hose depends on
flow rates. Fluid flow is also sensitive to relatively small- the position of the machine. Hose is also employed for con-
scale, geometric details of the flow path and surface condi- venience in manufacturing. In the latter case, the hydraulic
tion (roughness) of the components. hose is run like cable, where it will fit. Consequently, it may
The typical case of a machine’s hydraulic system is char- have a circuitous run.
acterized by the hydraulic power unit (pump)’s turbulence, Although we cannot place confidence in fluid resistance
sustained by the fluid’s passage through a variety of tees, values calculated a priori, we have no choice, in the absence
valves, fittings, and actuators. Turbulent flow can be con- of test data. A reasonable strategy is to estimate fluid resis-
sidered independent of Reynolds number, since it is “slug” tance at the presumed middle of the operating range of flow
flow, Fig. 5.2a. However, the Reynolds number of the flow is rates. One estimate is for a best-case scenario. Another es-
needed to estimate the friction factor, f, which is used to cal- timate assumes unfavorable limits for the viscosity and ge-
culate the “Resistance Coefficient” K and its corresponding ometry, where “unfavorable” depends on the function of the
“Equivalent Length” of a straight pipe, which would yield system being modeled.
the same pressure drop. The Colebrook equationis used to The ISO oil standards identify oil grade by viscosity in
calculate the friction factor, f , for turbulent flow with Reyn- centistokes at 40 °C. Hydraulic oil viscosities in the range
olds greater than 4,000. of ISO grades 32–100 are typical. The viscosity ranges for
engine oils (also called “motor oil,” if the lubricating oil used
  ε  in internal combustion engines) is expressed as “weight” in
1  2.51  Society of Automotive Engineers (SAE) standards. Mul-
= −2 log  D +  where ε ≡ roughness (5.30)
f 3.7 Re f tiple viscosity oils indicate their minimum and maximum
 
weights. For example, 5W20 indicates the viscosities, when
the oil is at the operating temperature, and when it is cold.
Published design values for the roughness of steel piping, tub- Refer to Table 5.1.
ing, and hydraulic hose vary considerably. Cited values for
new steel piping or tubing have ranged from ε = 0.0006 in.
to an order of magnitude lower at ε = 0.00006 in. The span 5.4.2  Fluid Inertance
of published values for new hydraulic hose is greater, from
ε = 0.001 in. to ε = 0.0003 in. This variation reflects the fact As is the case with mechanical systems, we typically have
that hydraulic hose is a manufactured product, whereas steel greater confidence in the parameter of the kinetic energy
tubing is a processed material. The roughness of the interior
284 5  Fluid, Electrical, and Thermal Systems

storage than in the system’s other parameter values. The only fluctuations or pulsations in the system. An electric capacitor
uncertainty in calculation of fluid inertance, I is used in the same way to smooth the voltage produced by a
diode bridge in a DC power supply.
ρL Gas-charged bladder-type accumulators, Fig. 5.11, re-
I=
A semble gas cylinders. They are described by two different
volumes. The larger volume of the two is the “nominal” vol-
is the density of the hydraulic oil. Although there can be ume of the entire interior volume of the pressure vessel. The
significant variation in viscosity among hydraulic oils, their more important and smaller volume is the “effective vol-
density is much less variable. Lubricating oils range in den- ume.” It is the maximum volume of hydraulic fluid, which
sity from approximately 850 to 950 kg/m3. Hydraulic oils can be expelled by gas-charged bladder, when fluid pressure
range approximately from 860 to 900 kg/m3. Emulsions are is varied from pmax to pmin. The nominal and effective vol-
denser than mineral oils, due to their water content. The umes are essentially equal for small accumulators, but differ
above densities are uncertain due to the unknown amount of by up to 10 % for larger accumulators, because the cylinder
air entrained in hydraulic oils. The amount of entrained air volume is increased by increasing the length rather than di-
can reach 9 % by volume at atmospheric pressure. ameter. This limits the expansion of the gas bladder.
Gas-charged diaphragm-type accumulators resemble ket-
tles. The elastomeric diaphragm is attached to the interior
5.4.3 Fluid Capacitance (Hydraulic circumference midway up the side. Gas-charged diaphragm-
Accumulators) type accumulators have equal nominal and effective vol-
umes.
Fluid capacitance C is easiest to calculate using the integral Gas-charged piston-type accumulators have advantage
form of the elemental equation over gas-charged bladder or diaphragm types of greater dis-
charge when used to store energy because their charge pres-
dp4 g t2 p2 sure is higher: 100 psi below pmin, rather than 80 % of pmin. If
QC = C
dt
→ QC dt = Cdp4 g → ∫ QC dt = C ∫ dp4 g pmin were 2,000 psi, the difference in charge pressure would
t1 p1
be 300 psi between the piston-type accumulators and the
others. It is possible to find gas-charged piston-type accu-
p2 Vol mulators in more combinations of diameter and length than
Vol = C p4 g → C=
p1 p4 g2 − p4 g1 are available for gas-charged bladder or diaphragm types.
They do have frictional resistance due to the seals between
Vol the piston and cylinder, but not as much as the spring-loaded
 C= (5.31)
∆p4 g piston-type accumulator, because the pressure drop across
the piston is not as great.
The spring-loaded piston-type fluid accumulator is the
The bulk modulus, β, of hydraulic oil lies in the range of least common type of hydraulic accumulator. The frictional
250,000 to 300,000 psi. This is attributed to the unknown resistance of the piston-cylinder seals which, as a rule of
amount of entrained air and its dependence on temperature thumb is 10 % of the rated net pressure force, is the same as
and pressure. The far greater bulk modulus of hydraulic oils that of a hydraulic piston-cylinder actuator.
makes it preferable to use nitrogen as a “gas spring” in hy-
draulic accumulators.
The most commonly used type in machine design is the 5.5 Fluid Power System Hardware
nitrogen gas-charged bladder-type fluid accumulator. The and Symbols
pressure of nitrogen gas is a percentage of the design mini-
mum hydraulic fluid pressure, pmin, and the intended function Most fluid power schematic symbols are self-explanatory.
of the accumulator. The engineer would choose the highest There are three worth noting. Of these, two are self-explan-
charge pressure, typically 80 % of pmin, if the accumulator atory, when their functions are understood. The third is a
is intended to store energy, in order to allow the use of a puzzle, until its corresponding hardware is introduced.
smaller-capacity pump, or to increase the responsiveness of
a servo valve. The lowest-charge pressure, 60 % of pmin, is
chosen, if the fluid capacitance had to be added to the sys- 5.5.1  Metering or Flow Control Valves
tem, to absorb a shock loading after the sudden stoppage of
a large actuator. A pressure between those limits is selected, Metering or flow-control valves are either fixed fluid resis-
if the fluid capacitance is intended to smooth out pressure tances, per Fig. 5.33a, or manually adjustable fluid resistanc-
5.5  Fluid Power System Hardware and Symbols 285

a b a
1 3 1 3

2
2
Fig. 5.33  a Fixed resistance metering or flow control valve, b Adjust- b
able metering valve
1 3 1 3

Flow lifts ball from valve seat


Flow
2 2
c
1 3 1 3
No Flow
Flow seats ball closing valve

2
Fig. 5.34   Check valve hydraulic schematic symbol 2

Fig. 5.35   Three-position spool valve. The schematic symbol for each
es, per Fig. 5.33b. The base symbol is a flow restricting to position is shown to the left. A cross-section through the valve is shown
indicate a fluid resistance. The diagonal arrow through the on the right. a Flow from port one to port two. b All ports blocked.
adjustable valve follows the same convention used to indi- c Flow from port two to port three. In practice, all valve positions are
shown on schematic as in Fig. 5.36
cate adjustable electrical components.

because it shows the function of each position of the valve,


5.5.2  Check Valves Fig. 5.35.
The name “shuttle” valve refers to components moving,
Check valves are used to limit backflow in fluid systems. or shuttling back and forth, within the valve body. This ac-
They are the fluid analog of an electrical diode, since they tion serves to block or reveal ports and, thus, control the
permit flow in only one direction. There is also a pressure flow. A poppet-type valve sports a hollow, cylindrically
drop in the flow direction through a check valve, analogous shaped “poppet,” which is closed at one end. A taper at the
to the voltage drop with flow of current through a diode. The closed end of the cylinder, along with a spring within the
schematic symbol for a diode is a circle in an angle, which is cylinder to push the taper against the valve seat, makes it
intended to imply a ball-type check valve, in which backflow a poppet valve. The diameter of the cylinder adjusts, to ei-
pushes a ball against the valve seat, closing the valve, per ther block fluid or allow flow when aligned with ports in the
Fig. 5.34. Ball-type and poppet-type check valves are used valve body, as described.
in hydraulic systems. A poppet valve has a tapered cylinder Spool valves are so named, because the moving compo-
or cone, rather than a ball, held against the valve seat by a nent, which controls the flow, resembles a spool for thread.
spring. In low-pressure fluid systems, most check valves are The spool is a precision-machined, cylindrical part of vary-
disk type, in which a disk is held open against a spring by ing diameters. When a small diameter portion of the valve is
the forward pressure of the flow. Flap valves are also used as aligned with pressure and discharge ports in the valve body,
check valves in low-pressure systems. there is flow through the valve. When a full diameter por-
tion aligns with the flow and discharge ports, the flow is
blocked. Multiple positions of a shuttle or spool valve allow
5.5.3  Multi Position Shuttle or Spool Valves one valve to control the movement of a hydraulic piston/
cylinder.
Shuttle and spool valves are important in machine design, due When spool valves are part of an automatic system, they
to their use to control the direction of fluid power flow. These are called servo valves. The prefix, “servo,” means slave.
valves fall under the category of “cartridge” valves, since the The most precisely manufactured servo valves are known as
overall shape of the mechanism fits into a cylindrical cav- “linear,” or “proportional,” valves. They are designed and
ity in the valve body or manifold. They are multi position manufactured, so the pressure drop through the valve is lin-
valves, whose schematic symbol is a puzzle to the uninitiated, early related to the position of the spool. Linear valves are
286 5  Fluid, Electrical, and Thermal Systems

1 3 1 3 1 3 are designed to be powered by constant voltage sources.) AC


is used, whenever the circuit or electrical machine needs a
time-varying magnetic field to function. The most common
examples are electric transformers and AC electric motors.
AC circuits are the topic of Chap. 11.
2 2 2 We must defer further discussion of AC circuits, until we
have developed the theory needed for “frequency response,”
Fig. 5.36   Schematic symbol for the three-position spool valve shown which is the steady-state response of a system to a sinusoidal
in Fig. 5.36. The schematic symbol shows the flow control for each input. We shall see that the response of a circuit to a sinu-
position of the spool soidal voltage source can be extended to any voltage with
periodically varying (repeating) voltage, by means of a Fou-
rier’s series approximation. Consequently, we will broaden
used in closed-loop feedback controls of hydraulically pow- our definition of DC circuits to be all circuits except those
ered machines. excited by a periodic voltage of any shape (or waveform).

5.6  Electrical Systems 5.6.1 Analogies Between Fluid and Electrical


Systems
We will introduce elements of electrical systems, drawing
on analogies to fluid systems. We will also investigate the Fluid systems are mechanical systems, since they have
phenomenon of magnetic energy storage, which results in mass and are described by Newton’s laws. The physical
inductance. laws, which describe electrical systems, were discovered
The electric field and the interaction of electric charges after Newton’s time. Although electrical systems involve
are both familiar and mysterious. A familiar aspect is the forces, the mass of the charge carrier is generally neglected.
force exerted between socks, when they are removed from However, there are strong mathematical analogies between
the dryer. It is a substantial force, far stronger than gravity, electrical and fluid systems. Many mechanical engineering
or we wouldn’t have to peel socks apart, as they would easily students find these analogies very helpful in understanding
separate, when we pick up one of the pair. The collection of electrical systems.
electric charge, the resulting voltage, and electric field are The mathematical analogies are apparent in a comparison
created by polymeric, synthetic fibers tumbling against one of the elemental and energy equations of fluid and electrical
another in the dryer. The voltage is created by the mechani- systems, Table 5.2. Both fluid and electrical systems contain
cal separation of charges. a quantity which flows through elements and is conserved.
A mysterious aspect of the electric field is that it is appar- The flows of the quantity are driven by a “potential” drop. A
ently an aspect of space itself, and exists at a random near- pressure drop between two ends of a pipe drives fluid flow
zero level, when no charges are in the vicinity. As bizarre as through the pipe. Likewise, a voltage drop between two
it sounds, if the electric field were created by charge, then ends of an electrical resistor drives current flow through the
we would know that the field would be exactly zero, if no resistor. The energy dissipation element in both systems is
charge was in the vicinity. The Heisenberg uncertainty prin- called a resistor. Whether the element is a fluid or an electri-
ciple prevents us from knowing the electric field’s strength. cal resistor is apparent with the identification of either fluid
If the strength of the electric field must be unknown, it can- or electrical variables in the equation. Similarly, the element
not be zero. Strange, but apparently true. that stores energy as a function of the across variable, ei-
A circuit which is powered (or excited) by a constant volt- ther pressure or voltage, is called a capacitor in both types
age or current is known as a DC (direct current), even though of systems.
the voltages and currents within the circuit may oscillate dur-
ing the transient period of the response to a step change in
applied voltage or current. The steady-state response of a DC 5.6.2  Summary of Electromagnetic Phenomena
circuit is a constant voltage or current.
The steady-state response of a circuit powered by a sinu- The following summarizes the fundamental electromagnetic
soidal voltage is sinusoidal. We refer to a circuit as an AC phenomena and definitions we need for the subsequent dis-
(alternating current) circuit. It is possible to power some cir- cussion:
cuits with either a constant voltage or a sinusoidal voltage. • Electric charge is conserved.
However, in practice, AC circuits are designed to be pow- • Electric charges of opposite polarities exist with the elec-
ered by sinusoidally varying voltage sources. (DC circuits tron and proton carrying unit “elemental” negative and
5.6  Electrical Systems 287

Table 5.2   Comparison of fluid and electrical system equations • The ampere is defined by the magnetic force acting
Fluid system Electrical system between two straight conductors, specifically, the ampere
Power P� = p12Q P�= v12i is that constant current which, if maintained in two
Energy dissipation Fluid resistance Electrical resistance straight parallel conductors of infinite length, of negli-
p12 = RQR v12 = RiR gible circular cross-section, and placed one meter apart in
Across-variable
Fluid capacitance Electrical capacitance vacuum, would produce between these conductors a force
Energy storage equal to 2 × 10−7 N per meter of length.
dp1g dv12 • Modeling two collections of charge, q1 coulombs and
QC = C iC = C
dt dt q2 coulombs, as located at points, the force exerted on a
1 1 2
EC = Cp12g EC = Cv12 static or moving charge q1 by a static point charge q2 is
2 2
calculated by the inverse square law:
Through-variable
Fluid inertance Electrical inductance
Energy storage
 1 q1q2
p12 = I
dQI
v12 = L
diL Felec = rˆ12 newtons (5.36)
dt dt 4πε r12 2
1 2 1 2
EI = IQI EL = LiL
2 2 where ε is the electrical permittivity, a property of the ma-
terial between the charges. The factor, 4π, was included to
aid calculations, since an inverse square law force incorpo-
positive charges, respectively. The magnitude of the ele- rates spherical geometry. The permittivity of “free space,” a
mental charge was determined by Millikan, who quan- vacuum, is
tified the electric force acting on miniscule oil droplets
falling between charged plates.  coulomb 2
ε 0 = 8.85 × 10 −12 (5.37)
• The unit of electric charge is the coulomb, C, which N · m2
is defined in terms of the flow of an electric current in
amperes as which yields
coulomb 1 N · m2
 ampere ≡ (5.32) 
second = 9 × 109 (5.38)
4πε 0 coulomb 2

• 1 coulomb = 6.2 × 1018 “elementary” or electronic charges. • A moving electric charge, q, creates a magnetic field, B,
A coulomb is an enormous amount of charge. In practice, which is proportional to the magnitude of the charge, q,
it is difficult to isolate such a large ­charge. and the velocity of the charge, v.
• The unit difference in electrical potential or electromotive • A static electric charge, q, is unaffected by a magnetic
force is the volt, V, defined the base SI units as field, B.
• A magnetic field, B, exerts a force on a moving charge
 kg ⋅ m 2 perpendicular to both the velocity, v, and the uniform
volt ≡ (5.33)
sec3 ⋅ ampere magnetic field, B, or, in other words, normal to the plane
defined by v and B. The magnetic force is calculated as a
Rewritten in derived SI units, cross product

 Fmag = q v × B (5.39)
kg ⋅ m 2 kg ⋅ m 1
volt = → volt = ⋅m⋅
sec3 ⋅ ampere sec 2
sec⋅ ampere Note that this implies that no magnetic force acts on a charge,
1 q (or current, i), ­moving parallel to the magnetic field, B,
 volt = N · m · since
coulomb (5.34)

volt · coulomb = joule (5.35)  Fmag = q v B sin θ (5.40)



Moving a coulomb of charge against the electrical potential where θ is the angle between velocity, v, of charge, q, and the
of one volt requires 1 J of energy. Likewise, one coulomb of magnetic field, B.
charge at one volt has the potential energy of 1 J, hence the
term, “electrical potential.”
288 5  Fluid, Electrical, and Thermal Systems

5.6.3  Electrical Units Joseph Henry (1797–1878), who discovered self-inductance


and invented the electric relay. The plural of henry is either
Historically, electrical and magnetic phenomena were dis- “henrys” or “henries.”
covered by their mechanical effects. Gauss (1777–1855)
championed the cgs (centimeter-gram-second) system of  volt volt · second
henry ≡ = (5.42)
measurement, in which electrical variables were derived ampere ampere
from mechanical variables. This system was also used by second
Maxwell (1831–1879) and Thomson (Lord Kelvin) (1824–
1907). It was very convenient mathematically, but not practi- An inductance of one henry is huge. The values of electrical
cally. Unfortunately, the derived electrical units depended on inductors used in electronic circuits are typically expressed
whether the electrostatic or the magnetic phenomenon was in microhenries.
used. “Practical” electrical units were introduced in 1875, as
electrical machinery became more common. The practical
electrical units were independent of the cgs system. Modern 5.7  Electrical System Elements
electrical units are SI units. The SI unit of current, the am-
pere, was adopted in 1948. The SI unit of electric charge, the There are two energy storage elements in electrical systems.
coulomb, is derived from the ampere. One volt will drive one Electrical capacitors store energy in an electric field created
coulomb (6.24 × 1018) of charge carriers, such as electrons, by opposite electric charge, and collected on parallel conduc-
through a resistance of one ohm in one second. One coulomb tive surfaces. Inductors store energy in the magnetic field,
of charge carriers, moving across a surface or through a node established by moving electric charges. The energy dissipa-
in one second, is one ampere of current. tion mechanism in electric systems is electrical resistance.
The sign convention for electrical systems was estab-
lished by Ben Franklin, long before the electron was discov-
ered. Franklin had to guess which way the charge carriers, 5.7.1  Electrical Resistance
electrons, were flowing; from positive to negative or from
negative to positive. He guessed wrong. We still follow his Electrical conductivity is the ease with which electric charge
convention of current flowing from positive to negative, moves through a material. Conductivity and its inverse,
even though we now know that electrons—the most com- electrical resistivity, are material properties. Most materials
mon charge-carrier—are negatively charged. Consequently, have negligible electrical conductivity. Of solids, the materi-
when current is composed of electrons, as is almost always als with the highest electrical conductivity at normal tem-
the case, the flow of charged particles is from negative to perature are metals, which form crystal lattices, and share
positive, which is the opposite direction of current. The type “conduction” electrons. Silver is the most conductive, mar-
of charged particle, and, consequently, the direction of the ginally more conductive than copper, followed by gold and
flow of charged particles, is inconsequential in the macro- aluminum. Silver and gold are used to plate contacts. Silver
scopic model of electrical systems. It is of great consequence acts to resist damage due to arcing, and gold acts to prevent
on the microscopic scale. Metals have low electrical resis- oxidation. Copper and aluminum are used as conductors.
tance to the flow of electrons but high resistance to the flow Semiconductors are metals with orders of magnitude less
of protons or ions. Aqueous solutions and molten salts, on conductivity than silver and copper, hence their name. Sili-
the other hand, can transmit either type of current. con and germanium are the two, which are most widely used
The SI definitions of electrical properties are logical but in electronics manufacturing. Carbon is conductive. The first
result in awkward magnitudes. For example, the property of electric lights were arc lights with carbon electrodes. Edi-
capacitance is expressed in farads with the symbol, F. son’s first light bulb used a carbon filament. Carbon is used
in its graphite form for motor “brushes,” which are the elec-
 volt volt trical contact between the stationary frame of the motor, and
farad ≡ → farad ≡ (5.41)
coulomb ampere · second the rotating commutator contact on the motor’s shaft.
The thermal motion of mobile conduction electrons in
A one farad capacitor can store one coulomb of charge at 1 V. metals is akin to the random motion of molecules in a gas.
Logical, but a one farad capacitor is an enormous capacitor. A voltage on a conductor creates an electric field, which im-
The common unit for capacitance used in electronic circuits poses a force on the conduction electrons, thereby accelerat-
is a picofarad, pF, where pico is 10−12. Electrical inductance ing them and bending their random motions in the direction
is the voltage developed across an electrical conductor, in of the electric field. The acceleration is modest, compared to
response to the rate of change of current of one ampere per the great speed of the conduction electron’s random motion.
second. The unit of inductance is the henry, H, named for However, the average velocity of the electrons in random
5.7  Electrical System Elements 289

a 60 ft of 24 AWG
solid copper wire

b
1.5 VDC
2Ω
Fig. 5.37  a US resistor symbol, b European resistor symbol 1.5 VDC

motion is zero, since they are moving in all directions. The


acceleration imposed by the electric force is sufficient to Fig. 5.38   Resistive electric circuit
create a non-zero average velocity called the “drift” veloc-
ity. The drift velocity is the velocity of the current. Colli-
sion between the conduction electrons and the copper lattice ity is not generally reported; its inverse, electrical resistivity
of the conductor transfers energy from the electrons to the is. Copper has resistivity of:
metal lattice. This energy transfer is in equilibrium, when the
electrons are in thermal motion. However, when an exter- 1.7 × 10 −8 Ω
ρcopper =
nal electric field accelerates the conduction electrons, there m
is net energy transfer to the lattice during collisions, which
results in heating of the metal conductor. This is the mecha- The unit of electrical conductance used to be “mhos” (ohms
nism of electrical resistance. Defects in the crystal lattice in- spelled backwards) until SI was introduced. It is now sie-
crease the frequency of collisions. Electrical grade copper is mens. Inverting the resistance of copper yields a conductiv-
99.99 % pure. High purity is needed to limit disruption of the ity of:
copper crystal lattice by contaminants.
Electrical resistance is the inverse of conductivity. Increas- 1
ing electrical resistance decreases the current produced by a 5.9 × 107 7
κ copper = Ω = 5.9 × 10 siemens
voltage. Decreasing the cross-sectional area of a conductor m m
increases its electrical resistivity. Increasing the length of a
conductor increases its resistivity. This is directly analogous The resistance of a given piece of wire is
to how the fluid resistance of a hose changes with its geom-
etry. If the cross-section of a hose is reduced, there will be Length
Rwire =
less flow for a given pressure drop. If the length is increased, κ copper · Area
there will also be less flow.
Two symbols are used for electrical resistance, per AWG (American Wire Gauge) 24, single-stranded (i.e.,
Fig. 5.37. The US symbol is a zigzag, which resembles the solid) wire has a diameter of 0.020 in. The cross-sectional
coil of toaster or a strain gauge. The European symbol is a area in square meters is
rectangle. 2
π 1m 
Area =  0.020 in · = 2.03 × 10 −7 m 2
5.7.1.1  Resistive circuit calculation 4 39.37 in 
An electric circuit shown in the schematic, Fig. 5.38. Two
1.5 VDC AAA batteries in series are connected to a 2 Ω re- [Aside: The diameter of gauge of wire in AWG is not a single
sistor by 60 feet of 24 AWG (American wire gauge) solid number. The diameter of a gauge varies with the number of
copper wire. Calculate the current through the circuit and the strands in the wire.] Multiple-stranded wire has two advantag-
voltage drop across the 2 Ω resistor. es in machine design. The primary reason for the use of multi-
In this calculation, the schematic may not represent our ple-stranded wire is its long fatigue life. Even if a solid copper
model. Normally, conductors in schematics are considered conductor is not intended to flex during the operation of a ma-
ideal and without resistance. In this circuit, resistance of the chine, a conductor is subject to fatigue, if it is free to vibrate
discrete resistor is relatively small and may be on the order when excited by a machine’s motion. At best, a fatigue failure
of the wire. We need to calculate the resistance of 60 feet of leads to arcing and then an open circuit. At worst, as the crack
the 24 gauge solid copper wire in series with the 2 Ω resistor. develops, the increased resistance increases the resistive heat-
The electrical material property analogous to magnetic ing, resulting in a fire. The secondary reason is that, although
permeability is electrical conductivity. Electrical conductiv- we model electrical current as conducted uniformly over a
290 5  Fluid, Electrical, and Thermal Systems

Equivalent resistance Energetic Equations


of 60 ft of 24 AWG Continuity:  Node 1 i = iR1   Node 2 iR1 = iR2
solid copper wire Compatibility:  v13 = v12 + v23
1.2 Ω Elements:  v12 = R1iR1 v23 = R2 iR2
1 2 Energy:  No energy storage elements.
Reduction Input: v13 = 3VDC, Output: v23
i R1
i R2 v13 = 3 VDC = v12 + v23 → 3 VDC = R1iR1 + v23
3V i 2Ω
3 VDC = R1iR 2 + v23

3 v   R1 
3 VDC = R1  23  + v23 → 3 VDC =  + 1 v23
Fig. 5.39   Resistive electric circuit with the equivalent resistance of  R2   R2 
60 ft of 24 AWG solid copper wire
 R1 + R 2   R2 
3 VDC =   v23 → v23 =   3 VDC
cross-section of a conductor, the current density is higher at  R2   R1 + R 2 
the surface of a conductor. Multiple-stranded wire has greater
specific surface area, which is the ratio of the surface area to The fraction of voltage dropping across an individual resistor
the volume. Hence, more current can be conducted by less in a series of resistors is the ratio of that resistance to the total
copper. For example, a given length of AWG 24 wire com- resistance of the series. You will find this result intuitive, if
prising 41 strands for AWG 40 wire weighs 5 % less than the you make the analogy between electrical resistors and fluid
same length of single-stranded AWG 24 wire and has 5.7 % resistances, e.g., steady flow through two garden hoses in
less resistance. series.
The length in meters is Substituting in values for resistances, R1 and R2, yields the
voltage drop, v23, across the 2 Ω resistor,
 1m 
Length = 60 ft  = 18.3 m
 3.28ft   R2   2Ω 
v23 =   3 VDC =   3 VDC = 1.7 VDC
 R1 + R2   1.5 Ω + 2 Ω 
Yielding an electrical resistance of

Length 18.3 m 5.7.2 Electrical Capacitance: Energy Stored


Rwire = =
κ · Area  5.9 × 10 siemens 
7 in an Electric Field
 m
( −7
 2.03 × 10 m
2
)
Fluid systems can store energy as a function of pressure,
18.3 m by straining the container (the fluid accumulator), straining
Rwire = (compressing) the fluid, or by pushing the fluid against grav-
 1 
5.9 × 107 ity. In all the three cases, we visualize pressure pushing fluid
 Ω  2.03 × 10 −7 m 2
 m  ( ) into the capacitor, until the pressure in the capacitor equals
  the pressure acting to push fluid in, and force equilibrium
is achieved. Electrical systems store energy as a function of
voltage in an analogous way. The energy stored in the elec-
Length
Rwire = = 1.5 Ω tric field of the capacitor is analogous to the energy stored
κ · Area
in the strain of fluid in a fluid capacitor. Work is done to
create the field, by moving like charged charge carriers into
The electrical resistance of the wire is the same order of proximity, just as work must be done to compress a fluid, by
magnitude as that of the resistor. Our model for the circuit is moving the fluid molecules closer together.
shown in Fig. 5.39: Capacitance exists wherever charge exists. Charge affects
Two resistors in series are called a “voltage divider” be- the electric field, E, and is affected by the electric field cre-
cause the voltage across the pair drops across an individual ated by other charges, whether the charge is stationary, or
resistor, in proportion to its contribution to the sum of the moving as a current. In order for current to flow into an elec-
resistances. We can now calculate the voltage drop across the trical capacitor, the charge flowing in must be pushed in, to
2 Ω resistor. overcome the repulsive force of the electrical field created
by the like charge already in the capacitor, Fig. 5.40.
5.7  Electrical System Elements 291

Fig. 5.40   A parallel plate


electrical capacitor. Note that the
a b c
electric current flows through the
iC
capacitor even though no charge
passes between the plates

iC Parallel Plate Adjustable Electrolytic


Non-Polarized Non-Polarized Polarized

Fig. 5.41   Capacitor symbols. a The parallel plate symbol is used for
non-polarized capacitors. b An arrow through an electrical symbol in-
Current is shown flowing through the capacitor. In reality, dicates the device is adjustable. c An electrolytic capacitor. The curved
line is the negative terminal, the cathode
no charge-carrier crosses the dielectric (insulator) between
the parallel plates. The electrical current through the capaci-
tor consists of oppositely charged charge carriers flowing in non-analogous aspects. First, fluid capacitors must be refer-
opposite directions. The two opposite signs cancel for the enced to the environment’s “ground” pressure, usually atmo-
negative charge carriers (electrons) moving opposite the cur- spheric pressure. There is an absolute measure of zero pres-
rent direction, resulting in a positive current flowing in the sure, a perfect vacuum. It is impossible to draw fluid from a
positive direction. perfect vacuum. Conversely, there is no absolute measure of
Two styles of capacitors are in common use. Figure 5.41a voltage. Voltage is always relative. Measurements of voltage
is a schematic of a “parallel plate” capacitor. Old radios were may be made relative to “chassis ground,” which is the ground
tuned by turning a knob that swung a stack of parallel alumi- used by a power supply in an electrical or electronic device.
num plates on a spindle between a fixed stack, thereby vary- Voltage measurement may also be made relative to “earth
ing the total area between the plates and the capacitance of ground,” which is usually the ground of the building’s electri-
the variable capacitor, Fig. 5.41b. “Non-polarized” capaci- cal wiring. The reference used to calculate the energy stored
tors, i.e., capacitors which do not have positive and negative in the capacitance is the difference in voltage between the two
terminals marked on them, used in electronics today have plates.
similar designs. These capacitors have conductors separated The second difference between fluid and electrical ca-
by an insulator, a “dielectric.” One contemporary design uses pacitance concerns their through variables. The through
a thin film of plastic, which has been aluminized to make its variable of a fluid capacitance, volume flow rate, does not
surfaces conductive. The aluminized surfaces are the plates, flow through the capacitance. The fluid is either temporarily
and the film is the dielectric. A second film for insulation is stored in the capacitance, as in the case of a fluid accumula-
added. The film is then folded, or rolled up, and encapsu- tor, or the hose, tube, or pipe is strained creating a tempo-
lated. The capacitor has a large surface area with a small gap rary increase in volume. Electrical capacitors have a flow of
between the plates. positive charge in the positive direction to the positive plate,
A second type of capacitor in use is an “electrolytic” ca- and negative charge in the negative direction to the negative
pacitor. The symbol for an electrolytic capacitor has an arc, plate. The negative signs of the charge and the direction of
in place of one of the parallel lines of the plate style capaci- the charge flow cancel, leaving an apparently positive cur-
tor. The negative terminal of the capacitor is the side with the rent flowing across the gap between the plates. This current
arc, per Fig. 5.41c. Electrolytic capacitors have a single alu- is “displacement” current. Maxwell conducted experiments
minum foil plate. A piece of paper is placed on top of the foil. which demonstrated that displacement current has all of the
The stack is rolled up, placed in a small cylindrical can, filled magnetic effects associated with conventional current.
with an electrolytic solution, and then sealed. The electrolyte
acts as the second plate of the capacitor. Electrolytic capaci-
tors are used for large capacitances. Electrolytic capacitors 5.7.3 Electrical Inductance: Energy Stored
are polarized, because of the specific ions in the electrolyte. in a Magnetic field
Heed the polarity markings on the electrolytic capacitors.
If you are lucky, then you only have a non-working circuit Electrical inductance is more difficult to understand than
and leaked electrolyte to cleanup. However, capacitors are electrical capacitance for three reasons. First, we can sense
known to have exploded due to reversed polarity. the presence of electric fields, because we carry electric
Analogies between fluid and electrical systems are not charge on our bodies and clothing. Everyone has felt the
perfect. Fluid capacitance and electrical capacitance have two “static electricity” created by separating clothing that was
292 5  Fluid, Electrical, and Thermal Systems

Torrodial ferrite core vacuum, has magnetic permeability. Ferromagnetic materi-


Relative permeability µr als, iron, nickel, cobalt, and their alloys, increase the strength
Core length l c of a magnetic field, due to the alignment of the spin axes
φ Cross-sectional area A of their atoms with the applied magnetic field. Ferrite is a
φ
ferromagnetic ceramic. Its ceramic aspect makes its electri-
v1 v2
iL cal conductivity very low, which is desirable to reduce the
iL
“eddy” currents induced by time varying, or moving, mag-
netic fields. Eddy currents heat the conductors, in which they
N turns of wire in the coil are formed. Temperature of the motor conductors is gener-
Fig. 5.42   A Toroidal Inductor
ally the factor, which limits the current and, hence, power an
electric motor can produce.
This coupling is called the “flux linkage,” and is given
clinging together. (Static electricity is created by relative symbol λ (Greek for l as in linkage). The flux linkage λ be-
motion of dielectric surfaces in contact, which can result in tween a magnetic flux φ and a coil is proportional to the
charge transfer between the surfaces. The mechanical work number of turns of wire in the coil, N, the flux flows through
done to separate the surfaces produces a voltage.) We are un- or “threads” (as in threading the eye of a needle) or “links”
able to directly sense magnetic fields. Consequently, we are (as in links of a chain).
oblivious to them and unaware of their many effects.
Secondly, electric fields are created by the collection of  λ = Nφ (5.43)
like charged charge carriers, whereas magnetic fields are cre-
ated by the motion of charge carriers. Static phenomena are If the relationship between the magnetic flux φ threading a
typically easier to understand than dynamic phenomena. coil and current i through a coil is linear, then there is a linear
Thirdly, the electric forces between static charge carriers relationship between the flux linkage λ and current i, which
act in a straight line on the axis between the centers of the defines the parameter inductance L.
charge carriers, and are easy to visualize. Two vector cross-
products are needed to describe the magnetic forces acting be-  λ = Li (5.44)
tween two interacting, moving charge carriers. Consequently,
magnetic forces are difficult to visualize. Differentiating this expression with respect to time yields:
The magnetic energy storage property of electrical cir-
cuit elements is expressed as “inductance” L with the unit dλ di
=L
of henry, H, Fig. 5.43. The parameter symbol L is in honor dt dt
of Heinrich Lenz, credited with Lenz’s law. Henry was an
American scientist active in the early 1800s. A voltage v12 develops across an electrical inductor in re-
An “induced” current is created by a time varying, or sponse to a time-varying current i, or, in truth, due to the
moving, magnetic field. Lenz’s law states that an induced time-varying flux linkage, λ . The current through the coil re-
current creates a magnetic field in opposition to the magnetic sists change because energy must flow into to coil’s magnet-
field, which created the current. This is a statement of energy ic field as the current increases and, similarly, flow out of the
conservation. If the induced current fed back and increased coil’s magnetic field as the current decreases. The voltage
the magnetic field, the process would create a magnetic field acts to drive the change in the current through the coil as the
of infinite energy. current resists change. The voltage across a coil is analogous
Inductance L is a measure of how large a magnetic flux a to the force acting to accelerate a mass. The force drives the
given current i can create. Magnetic flux φ is the component change in the kinetic energy in the mass. Signs are a problem
of magnetic field B, normal to an area integrated over that for the voltage across the coil. The correct sign for any sign
area. The term, flux, is used to evoke a flow analogy, since convention is the sign which leads to energy conservation. In
the magnetic field lines form closed curves. The intensity the linear graph method, the voltage across the coil has the
of the magnetic field increases with current, the number of same sign as the change in the current through the coil.
turns of wire in the coil, and the relative permeability of the
material in the coil. The turns of wire, which form the coil,  di
v12 = L (5.45)
are said to be “linked” to the magnetic flux φ they create, dt
since the magnetic flux φ passes through the turns of the coil.
Figure 5.42. The two differential equations above reduce to a relationship
Magnetic permeability is a “material” property. Quotes between the time rate of change of the flux linkage λ and the
are used, because free space, the physics term for a complete resulting voltage, v12 .
5.7  Electrical System Elements 293

 dλ
= v12 (5.46) a b
dt L
L
1 2 1 2
Equating the definition of flux linkage λ and the relationship
between flux linkage and inductance
iL iL
 λ = N φ = Li (5.47)
Fig. 5.43  a Schematic symbol for an air core inductor. b Schematic
symbol for an iron core inductor
yields an expression for the magnetic flux φ in terms of the
inductance of the coil, the current through the coil, and the
number of turns of the coil Table 5.3   Mathematic analogy Fluid Electrical
between fluid inertance and inertance inductance
electrical inductance
 Li dQI diL
φ= (5.48) p12 = I
dt
v12 = L
dt
N
1 2 1 2
EI = IQI EL = LiL
The product of the current i passing through a coil and the 2 2
number of turns in the coil is the strength of the externally
applied magnetic field, and is known as the magnetomotive
force, abbreviated mmf. The analogy between electrical inductance and fluid iner-
tance is very helpful, Table 5.3.
 magnetomotive force ≡ mmf = Ni (5.49) The fluid inertance element represents the kinetic energy
storage, which results from the motion of fluid. A net pres-
The relationship between the strength of an externally ap- sure force is needed to accelerate or decelerate a mass of
plied magnetic field and the magnetic flux in the core of an fluid. Time is needed to change the amount of stored kinetic
inductor (Eq. 5.50) is analogous to the relationship between energy, thereby changing the energy storage variable veloc-
the voltage applied to an electrical resistance and the current ity. Electrical inductance is the magnetic energy storage,
through the resistance which results from the motion of charge carriers (current).
A voltage is needed to increase or decrease a current. Time
 Ni ≡ mmf = R φ (5.50) is needed to change the amount of magnetic energy stored,
which changes the energy storage variable current. A typical
where R is “reluctance” which is a measure of the difficulty inductor is a coil of high conductivity copper wire wrapped
of creating a magnetic field. around a ferromagnetic “core,” such as iron, electrical steel,
Magnetic reluctance varies with geometry, in the same or ferrite, which acts to enhance the magnetic field.
manner as electrical resistance varies with geometry, Inductance is counter-intuitive, unless an inductor is
Eq. 5.51. Magnetic permeability μ is analogous to electrical viewed as an energy storage element. In steady-state, a coil
conductivity. of copper wire has little resistance to the flow of current. In
the case of an ideal inductance, the coil has zero resistance
 Length
R = (5.51) to the flow of current in steady-state. There is transient or
µ ⋅ Area dynamic resistance which represents the flow of energy into
or out of the magnetic field created by the current. A volt-
Using mmf ≡ Ni = R φ yields age applied to a coil cannot instantaneously create a current
because the current through the coil is the energy storage ele-
N2 ment. A voltage applied to a coil leads to a rate of change in
R = the current through the coil, just as a force applied to a mass
L
leads to a rate of change in the velocity of the mass.
And using the definition of magnetic reluctance, Eq. 5.51,
yields
5.7.4  Electrical Systems
L · Length
µ= 2
N · Area Schematics of electric circuits are models of the circuit. All
components of every circuit possess resistance, capacitance,
This expression is useful in energy methods to calculate the and inductance. The energetic attribute which dominates a
force, or torque, created by an electric motor. circuit component is the property we assign the component.
294 5  Fluid, Electrical, and Thermal Systems

a b R
Capacitance, C
1 R 2 1 2
-
+
V1 C V2 v(t) C
Resistance R1
Voltage source
g g
Fig. 5.44  a Electrical engineering convention, b Equivalent circuit Coil around ferrite core with
with ­voltage source Induction L and Resistance R 2

Fig. 5.45   An electrical system


The dominant property is the reason the circuit designer
included the component in the system. The most obvious
electrical properties omitted from a schematic of an electric our perspective to what is often referred to as “thermal man-
circuit are the resistance and inductance of the conductors agement,” which is the removal of waste heat from machines
connecting the components. Occasionally, a designer errs and devices. Thermal management is an essential aspect of
and uses too small a wire, or “trace,” on a printed circuit machine design, because a temperature increase due to the
board (PCB). The omission inadvertently creates a resistance accumulation of energy dissipated as heat limits the operat-
which affects the performance of the circuit. ing range of a device.
Linear graphs evolve from electric circuit schematics.
There is no need to draw a linear graph of an electric circuit
schematic. There is an electrical engineering convention to 5.8.1  Thermal Power Variables
show a portion of a circuit without the source, which ren-
ders it an incomplete circuit, Fig. 5.44. The input variable is In a static system, the difference in temperature between two
identified as the voltage difference between the input node locations determines the flow of heat. The across variable of
and ground. The corresponding circuit includes the voltage a thermal system is temperature, given the symbol, θ. Tem-
source. perature is analogous to other across variables. If every loca-
Linear graphs are drawn, when working with an electrical tion in a system is at the same temperature, then there is no
system for which there is no schematic. The linear graphs heat flow in the system. As noted above, the through variable
for a coil that possesses significant inductance and resistance is heat flow rate, q.
must include a fictitious node, with no physical location, to
divide the voltage across the terminals of the coils between 5.8.1.1 Analogies and Their Limits for Thermal,
the resistance and inductance properties, Figs. 5.45 and 5.46. Fluid, and Electrical Systems
The analogies across thermal, fluid, and electrical systems
are limited but useful. Heat flows in the direction of de-
5.8  Thermal Systems creasing temperature. Electric current flows through a resis-
tor in the direction of decreasing voltage, and is a poten-
Thermal systems do not fit the set of analogies used for tial driven flow, where the potential is electrical potential
mechanical, electrical, and fluid systems. Thermal systems or voltage. The strongest analogy to a thermal system is a
have only one type of energy storage element, thermal ca- fluid system with flow velocity low enough, that its kinetic
pacitance. Also, power in a thermal system is not the product energy hence inertance is negligible. A low velocity flow
of two variables. Heat flow rate, q, is both power and the of fluid through soil or a filter is a “seepage” flow. Seepage
through variable in a thermal system. flows are pressure driven, from high pressure to low pres-
Heat is energy. Heat exists in two very different forms. In sure. Temperature and pressure both have an absolute zero
matter, heat is vibrational kinetic energy of atoms and mol- reference. Each uses a common, or human scale, reference:
ecules about their mean position. Heat is also transmitted zero Celsius and atmospheric pressure.
through matter or a vacuum as photons of electromagnetic The analogies between the fluid and charge flow and the
radiation in the infrared wave lengths. flow of heat break down, when fluid velocity or amount of
As students of thermodynamics know, a flow of entropy current is large enough to store energy in either inertance or
accompanies a flow of heat. Thermal systems are of great inductance, which has no analogy in a thermal system. There
importance in mechanical engineering due to our continued is no “inertia” in heat flow. Heat flow can start instanta-
reliance on heat engines as our prime movers. We will limit neously, because the motion involved is on the atomic level.
5.8  Thermal Systems 295

Fig. 5.46  a Electrical system


showing voltage nodes. A ficti-
a Capacitance, C b R2 I
tious node is added to the coil to g 1 2 3
divide the voltage drops due to -
resistance and inductance, which + 1
occur over the length of the coil. 3
b Linear graph of the electrical R1
Resistance R1 v(t) C
system
Voltage source
2
Coil around ferrite core with
Induction L and Resistance R 2 g g

We shall see that thermal capacitance and fluid capaci- The thermal conductivity of a single, homogeneous mate-
tance have similar restrictions due to their direct reference to rial is a “specific” property, which describes the rate of heat
their “ground.” Charge, fluid, and heat are stored in capaci- flow q through a unit area normal to the heat flow path, per
tances. However, the energy stored in an electrical or fluid unit length along the path for a unit temperature gradient,
system is calculated as the square of the voltage or pressure, -dθ/dx, in the direction of the heat flow. The conversion from
respectively. The quantity of heat stored in a thermal capaci- US customary units to SI is:
tance is proportional to the temperature, not to the tempera-
ture squared. As such, thermal capacitance is mathemati- Btu · ft W
Thermal conductivity k : 1 = 1.730
cally analogous to storing energy, by elevating mass against hr · ft 2 · o F m · oK
gravity.
(Conversion between US and SI units is complicated by a
slight difference in the definition of a Btu. The difference is
5.8.2 Modes of Heat Transfer and Their due to the variation in specific heat of water with tempera-
Corresponding Thermal Resistances ture. The above conversion uses the US definition of a Btu.)
The thermal conductivity of all materials is a function
Heat is energy. Consequently, heat flows. It does so spon- of temperature. Hence, thermal conductivity is a non-linear
taneously, in the direction of decreasing temperature. The property. It is linearized by modeling the thermal conductiv-
movement of heat is known as heat “transfer.” There are ity of a material as constant, using the thermal conductivity
three modes of heat transfer: conduction, convection, and for the expected average temperature.
radiation. A heat transfer coefficientis the rate of heat flow q through
a unit area per unit time.
5.8.2.1 Conduction
Conductive heat transfer is the movement of heat by the Btu W
Coefficient of heat transfer cp : 1 2 o
= 5.674 2
transfer of atomic or molecular kinetic energy between hr · ft · F m ·K
neighbors in a static system. Fourier’s law of heat conduc-
tion, Eq. 5.52, describes heat flow: where the subscript, p, stands for constant pressure.
In the case of conductive heat transfer, the heat transfer
 dθ coefficient, hc, is calculated, by multiplying the thermal
q = −k (5.52)
dx conductivity of the material by the length of the heat flow
path through the material. When heat passes through mul-
where θ is temperature, and k is thermal conductivity of the tiple materials in contact with one another, the “overall” heat
material. transfer coefficient is calculated, by modeling the heat path
Conductive heat transfer usually refers to heat transfer as thermal conductivities in series, or, more conveniently,
through a solid body, such as a cast iron frying pan or an as thermal resistances in series, as we shall in Sect. 5.8.31,
aluminum bar. The material need not be solid for conductive Fig. 5.47.
heat transfer to occur, but it must remain static, i.e., motion-
less, at least on a small scale. Conductive heat transfer is 5.8.2.2 Convection
the movement of heat without the movement of matter. Heat Convective heat transfer is the movement of heat with the
conduction is present in a liquid in a Lagrangian reference movement of a fluid. Convective heat transfer combines
frame, moving with the fluid. Conductive heat transfer com- conduction of heat through fluid, and the movement of heat-
pletes the distribution of heat, analogous to diffusion com- ed fluid. The division between the two mechanisms depends
pleting the distribution of matter during the mixing of fluids. on the thermal conductivity of the fluid, the heat capacity of
296 5  Fluid, Electrical, and Thermal Systems

Perfect Insulation from a solid surface to the fluid, due to the lower temperature
No heat flow
of the fluid against the solid surface. This is a familiar phe-
nomenon to those who heat water on a stove. A pan or kettle
θ1 Material 1 Material 2
which has been gently tipped back and forth to agitate the
q ∆θ θ2 q water heats faster than one left undisturbed. Increasing the
∆x temperature difference between the heated surface and the
Heat Flow ∆θ
Rate
x ∆x θ3 fluid increases the rate of heat flow.
Forced convection usually involves turbulent fluid flow,
because of the turbulence introduced into the flow by the fan
Fig. 5.47   Heat conduction through two materials with different ther- or pump, and the higher speed of the flow. Turbulent flow
mal conductivities. The thermal conductivity of Material 1 is less than enhances convective heat transfer, by mixing the fluid nor-
that of Material 2 mal to the heated surface. The pressure gradient moves the
flow parallel to the heated surface, thereby reducing the tem-
perature of the fluid on the heated surface.
the fluid, mixing of the flow, and the speed of the flow. If the Because the heat transfer from the solid to the liquid, by
flow is turbulent, as is assumed for the flow of fluids in dy- definition, occurs in fluid that is in contact with the solid
namic systems, then the mixing of heated fluid is thorough, surface, the heat transfer coefficient, hf , is called a “film”
and the temperature is uniform over a cross-section in the coefficient, and applies only in a thermal boundary layer at
pipe, tube or hose, at a distance from where the heat entered the solid-fluid interface, Fig. 5.48.
the fluid. Convective heat transfer in machine design usually oc-
Convective heat transfer is classified as either “free” or curs in a series of subsystems. For example, the power dis-
“forced” convection. Free convention occurs when the mo- sipated as heat in an automatic automobile transmission is
tion of the fluid against the higher temperature surface is due transferred from the transmission to the automobile’s radia-
to temperature dependence of the fluid’s density. Generally, tor by tubes, which connect to a coil within the tank of the ra-
fluids become less dense with increasing temperature, as the diator. The heat is then transferred to the water-based engine
increased kinetic energy of their atoms or molecules increas- coolant in the radiator. The heated water is pumped through
es the mean distance between them. The buoyancy of regions the tubes of the radiator, where it is transferred to air, that is
of higher-temperature fluid causes it to float. The convective passed over the fins of the radiator by an electrically driven
motion allows the liquid at a heated surface to be replaced by fan.
denser, colder fluid. Convective “cells” can be seen in water Heat transfer in automotive design has always been im-
being heated in a pan. Free convection is also noticeable in portant, and has become more so, as the engine compart-
the early morning, as the sun heats moist surfaces. The rising ments become smaller and more crowded, and the tempera-
mist is free convection. ture within increases.
Forced convection occurs when a pump or fan forces or
drives fluid motion against a heated surface. Early, coal-fired 5.8.2.3 Radiation
central heating systems often relied on free convective flow. The heat radiated from a blackbody, an ideal radiator, with
Modern central heating systems have forced convection of surface area A through a hemisphere, qhemisphere, is described
either heated air or water. Forced convection is clearly more by the Stefan–Boltzmann Law:
involved than free convection. The expense of the additional
hardware is offset by the increased efficiency of heat transfer  qhemisphere = Aσθ 4 (5.53)

Fig. 5.48   Heat flow from fluid Perfect Insulation


1 through solid materials 1 and No heat flow
2 into fluid 2. The heat transfer x
through the thermal bound- θ1 Solid
ary layers in fluids 1 and 2 is
described by the film coefficients θ2 Material 2
for the two fluid–solid interfaces q ∆θ θ3 q
∆x
Heat Flow ∆θ
Rate ∆x θ4
Solid
Fluid 1 Material 1 θ5
Fluid 2
Thermal Thermal
Boundary Boundary
Layer Layer
5.8  Thermal Systems 297

Fig. 5.49   Thermal system con- Perfect Insulation


sisting of a layer of stainless steel No heat flow
and a layer of aluminum
θ1 x θ2 θ3

q q
Stainless
Steel Aluminum
Heat Flow
Rate
Surface Area 0.5 in 2 in
A = 48 ft 2

where T is the absolute temperature of the surface in de- where θ is temperature, q is heat flow rate, both the through
grees kelvin, and σ is the Stefan–Boltzmann constant, variable and power, and Rthermal is the thermal resistance.
σ = 5.67 × 10−8 W/(m 2 K 4 ). Thermal resistance is calculated as the inverse of the
The Stefan–Boltzmann law describes the radiation of heat product of the heat transfer coefficient and the cross-section-
from a surface. That surface will also receive radiated heat al area of the heat flow path.
from other surfaces. The net radiative heat transfer between
two surfaces at different temperatures is the difference be-  L
Rthermal =
tween the heated radiated from the surface, and the heat radi- k ·A (5.55)
ated to the surface, which was absorbed by the surface and
not reflected. The amount radiated will, in general, differ where k is thermal conductivity, L is the length of the heat
from the ideal blackbody, and is corrected by the “emissiv- flow path, and A is the cross-sectional area of the heat flow
ity,” ε, of a surface. Likewise, the fraction of the radiation path.
striking the surface which is absorbed will differ, and is cor- Thermal resistances combine in the same way that electri-
rected by the “absorbance,” α. cal and fluid resistances combine, Sect. 5.9.1. Resistances in
series sum to yield the equivalent thermal resistance. Resis-
5.8.2.4  Phase Change tances in parallel sum as inverses to yield the inverse of the
Phase changes either store or liberate “latent” heat. Boiling equivalent resistance.
and evaporative cooling are the most common mechanisms Example: Figure 5.49 shows a thermal system, consist-
of phase change. Phase change of the working fluid in a hy- ing of a half-inch thick layer of stainless steel and a 2-in.
draulic system, such as the boiling of brake fuel, leads to cata- thick layer of aluminum. The area of the heat flow is 48 ft2.
strophic failure. The brake components must be designed to Stainless steel and aluminum have the following thermal
limit temperature rise in brake fluid. conductivity, k, at 212 °F:

Btu Btu
5.8.3  Thermal System Elements k SS = 10 o
and k Al = 119
hr · ft · F hr · ft · o F
Electric circuit calculations are performed with electrical re-
sistance, rather than its inverse, electrical conductance. Simi- Calculate the thermal resistance of each layer and the single
larly, thermal system calculations are performed with ther- equivalent thermal resistance.
mal resistance, rather than its inverse, thermal conductiv- We will first convert to SI as standard practice. Electrical
ity. The single energy storage mode is thermal capacitance, units are SI, and problems involving systems with kinetic
which must be referenced to ground temperature. energy storage are easier to work in SI. If the results must
be expressed in US Customary Units, we will convert back
5.8.3.1  Thermal Resistance at the end.
The elemental equation for thermal resistance has the same
form as electrical and fluid resistance, potential = resistance W
1.730
times flow rate: Btu m · oK W
k SS = 10 = 17.3
hr · ft · o F Btu·ft m · oK
→ θ12 = Rthermal qRthermal 1
v12 = RiR and p12 = R fluid QR fluid hr · ft 2 · o F

(5.54)
298 5  Fluid, Electrical, and Thermal Systems

W
1.730
Btu m · oK W ture, making the property non-linear. We customarily linear-
k Al = 119 = 206
o
hr · ft · F Btu · ft m · oK ize the property, by using a representative value from within
1
hr · ft 2 · o F the operating range. The heat capacity of an object is the
product of its specific heat, cp, times its mass, M:
These thermal conductivities are specific conductivities for
an area of the heat path equal to the unit length squared, i.e.,  Cp = M cp (5.57)
ft2 or m2, and the length of the heat path equal to the unit
length, i.e., ft. or m, per-unit degree temperature difference The heat capacity of a material, defined as the amount of
in either oF or oK. heat required to raise the temperature of a mass of the mate-
The heat path lengths in meters are: rial one degree, is measured at either constant pressure or
constant volume. Therefore, the amount of heat stored in the
 1 cm   1 m  mass is the product of the change in temperature and the heat
LSS = 0.5 in  = 0.0127 m
 2.54 in   100 cm  capacity of the mass, Eq. 5.58.

 1 cm   1 m   EC = C p ∆θ (5.58)
LAl = 2.0 in  = 0.0508 m
 2.54 in   100 cm 
p

Note the energy stored in a thermal capacitance is propor-


The cross-sectional area of the heat path is: tional to the change in temperature, not the change in the
temperature squared. The reason is that heat flow is power.
2 2
 144 in 2   2.54 cm   1 m  In all other types of systems, power is the product of the two
A = 48 ft 2  = 4.459 m 2
 1 ft 2   1 in   100 cm  power variables of that system. In the energy storage equa-
tions, one power variable has been eliminated by an expres-
The thermal resistances are: sion written in terms of the other, resulting in that variable
being squared.
o The heat capacity of a heterogeneous object is the sum of
0.0127 m K
RSS = = 1.65 × 10 −4 the heat capacities of the masses of materials, which com-
W W
17.3 4.459 m 2 prise the object.
m · oK

0.0508 m o
K 5.8.3.3 Heat Sources: Friction, Fluid Shear,
RAl = = 5.53 × 10 −5 and Electrical Resistance
W 2 W
206 4.459 m Mechanical energy is dissipated as heat by friction, either
m · oK
viscous friction due to the shear of fluids, or “dry” Coulomb
friction between two solid surfaces in contact. The power,
The single equivalent thermal resistance for the thermal re- which flows into a translational or rotational damper, is dis-
sistances, RSS and RAl, in series is the sum of those resistances. sipated as heat by viscous friction. It is important to clarify
o o
the phrase, “flows into,” and distinguish it from the similar
K K phrase, “flows through.” “Flows into” means “flows into
Requiv = RSS + RAl = 1.65 × 10 −4 + 5.53 × 10 −5
W W and is lost as heat.” The power, which flows into a damper
o
K and is lost as heat, is the product of the two power variables
Requiv = 2.20 × 10 −4 of the damper. It is the product of the force, or torque, act-
W
ing through the damper and the velocity difference across
the damper. The power supplied by the force source and
5.8.3.2  Thermal Capacitance the power dissipated in dampers, b1 and b2, in the system
The specific heat of a material is defined as the amount of shown in Fig. 5.50 are Psource = F (t ) v1g , Pb = Fb v12, and
1 1

heat required to raise a unit mass of the material by one de- Pb = Fb v2 g . The power, which flows through damper b1,
2 2
gree: is the product of the force acting through the damper, and
the lesser of the velocities of the nodes at either end of the
 Btu kJ damper, v2g, in this case.
cp = 1 o
= 4.184 (5.56)
lb m · F kg · o K
Psource − Pb = Pb
1 2
→ F (t ) v1g − Fb1 v12 = Fb2 v2 g
where the subscript, p, stands for constant pressure. The spe-
cific heats of solids and fluids are functions of their tempera- Conventional automobile brakes dissipate power as heat
by dry friction. Dry friction is more difficult to include in
5.8  Thermal Systems 299

b1 B
1 2 teslas
The energy dissipated
as heat each cycle equals
the area of the hysteresis loop

F(t) b2 amp.turns
H ______
meter

g
Fig. 5.51   Magnetic hysteresis dissipates magnetic energy as heat
Fig. 5.50   Mechanical system which acts as a heat source

lost in magnetic hysteresis allows more current to be passed


dynamic models, because dry friction is a non-linear phe- through the motor’s windings.
nomenon. The friction force must reverse direction, when A second phenomenon which heats an electric motor is
the relative motion reverses direction. Modeling dry friction an “eddy” current. Eddy currents are induced in the iron and
requires numerical methods. An algorithm for dry friction is steel of the motor’s rotor as it moves through the magnetic
presented in Chap. 8. field of the stator. The resistance in the rotor causes the eddy
Fluid energy is dissipated as heat, due to viscous shear. currents to dissipate, heating the rotor. The rotors of many
Although all real fluids have viscosity, the effect of viscosity electric motors are constructed of stacks of sheet steel stamp-
on the flow of fluids in a dynamic model is represented by ings which are coated or painted to insulate the sheets from
fluid resistances. All power which flows into a fluid resis- one another, in order to reduce the length of a possible eddy
tance is dissipated as heat. current path.
Electrical energy is dissipated as heat, by electrical resis-
tance. All power which flows into an electrical resistor is dis-
sipated as heat. Magnetic hysteresis is the energy lost during 5.8.4  Thermal Systems
the cycle of creating and destroying a magnetic field in a
ferromagnetic material, such as the iron of an electric motor. We will consider a system with thermal resistance and ther-
It is a significant source of heating in many motors. Magnetic mal capacitance, for the purpose of modeling heat diffusion.
hysteresis is commonly included with the resistive heating Purely resistive thermal system models are used to estimate
of motors, because both are a function of the current flow the heat flow rates and temperatures in steady-state. In order
through the motor. to model the transient period during which the system “heats
up,” thermal capacitances must be included to store heat.
5.8.3.4 Heat Sources: Magnetic and Mechanical Heat diffusion is the term which describes the transient
Hysteresis portion of heat flow. A material’s property of thermal dif-
Hysteresis occurs, when the paths followed during load- fusivity increases with its thermal conductivity, and de-
ing and unloading of a system are not the same. Hysteresis creases with its specific heat. Increased thermal conductivity
occurs in mechanical systems, when there is plastic defor- increases the heat flow rate through the material. Increased
mation of a component during loading. The energy used to specific heat requires more heat to raise the material’s tem-
create plastic deformation is dissipated as heat, and is not perature.
recovered, when the component is unloaded. Hysteresis We will model the stainless steel and aluminum thermal
also occurs when a magnetic field is created in a ferromag- system of Sect. 5.8.3.1 to create a linear graph to calculate
netic material, such as iron, nickel, cobalt, and their alloys, transient response of temperature. In order to model the
Fig.  5.51. Some of the work done rearranging the micro- thermal capacitance of the metals, we will divide the alu-
scopic magnetic domains of the material, as the magnetic minum into regions, one-half inch in thickness. The thermal
moment of the atoms are aligned with the externally applied capacitance will be calculated for the entire mass of the re-
magnetic field, is not recovered and is lost as heat. Magnetic gion, using the temperature of the in-center of the region,
hysteresis is of great practical importance, because electri- Fig. 5.52. Heat will flow through thermal resistances equal
cal machinery, i.e., electric motors, experience at least one to the thermal resistance of the thickness of the regions, a
cycle of magnetic hysteresis per revolution. The power of quarter of an inch.
an electric motor is limited by its maximum current. The The thermal resistance for half of the thickness of the stain-
maximum current, in turn, is limited by the temperature rise less steel layer is half of what we calculated in Sect. 5.8.3.1.
of the motor in operation. Reducing the amount of energy
300 5  Fluid, Electrical, and Thermal Systems

Fig. 5.52   Thermal system divid- Perfect Insulation


ed into five regions, each with a No heat flow
thermal capacitance between the
center temperature node of the
region and ground temperature RSS RSS RAl RAl RAl RAl RAl RAl RAl RAl
θ1 θ11
q in 2 3 4 5 6 7 8 9 10 qout
Heat Flow Stainless
Rate Steel Aluminum

Surface Area 0.5 in 2 in


A = 48 ft 2

3 3
Thermal resistance scales with the length of the heat flow lb m  0.454 kg   1 ft   1in  kg
ρ Al = 169 = 2, 710 3
path, all else being equal. ft 3  lb m   12 in   0.0254 m  m
 
1 0.0127 m  o
−5 K The thermal capacitance, Cp, of the half-inch thick layer of
RSS1 = RSS2 =   = 8.23 × 10
2 W 2 W stainless steel equals the product of its mass, M, and specific
 17.3 m · o K 4.459 m  heat, cp.

Similarly, the resistance of one of the eight 2-in. thick alu-  0.0254 m  kg
M SS = 0.5 in   4.459 m 2 · 7,820 3 = 443 kg
minum layers is  1 in  m

  kJ kJ
C pSS = Mc pSS = 443 kg ⋅ 0.46 → C pSS = 204
1 0.0508 m  o
−6 K kg ⋅ o K o
K
RAl =   = 6.91 × 10
8 W 2 W
 206 m · o K 4.459 m  The thermal capacitance of a half-inch thick region of the
aluminum is
The specific heat and density of stainless steel (18 Cr, 8 Ni)
are:  0.0254 m  kg
M Al = 0.5 in   4.459 m 2 · 2, 710 3 = 153 kg
 1 in  m
Btu lb
c pSS = 0.11 and ρSS = 488 m3
lb m o F ft kJ kJ
 C pAl = Mc pAl = 153 kg ⋅ 0.895 kg ⋅ o K → C pAl = 137 o K
The specific heat and density of aluminum are:
The linear graph of this model is shown in Fig. 5.53.
 Btu lb The linear graph is intimidating, but it can be easily re-
c pAl = 0.214 o and ρ Al = 169 m3
lb m F ft duced by the state-space method introduced in Chap. 7. The
state-space method can deal with multiple inputs. There are
Convert these values from US customary units to SI: two temperature sources in the linear graph to specify the
surface temperatures on ends. The reduction of this linear
 kJ  graph to a system of state equations and their solution com-
4.184
Btu  kg · o K  kJ prise Problem 7.20 of Chap. 7.
c pSS = 0.11 o   = 0.46
lb m F  Btu  kg · o K
 lb · o F 
m
5.9 Equivalent Elements in Fluid, Electrical,
3
lb m  0.454 kg   1 ft   1 in 
3 and Thermal Systems
kg
ρSS = 488 = 7,820 3
ft 3  lb m   12 in   0.0254 m  m
When two or more like elements are in series or in parallel,
we can simplify the system’s model, by combining the like
 kJ  elements into a single equivalent element. The equivalent el-
4.184
Btu  kg · o K  kJ ement will store or dissipate power at the same rate, as the
c pAl = 0.214 o   = 0.895
lb m F  Btu  kg · oK sum of the power stored or dissipated by the original ele-
 1 lb · o F 
m
5.9  Equivalent Elements in Fluid, Electrical, and Thermal Systems 301

Fig. 5.53   Linear graph of the R ss1 Rss 2 R al 1 R al2 R al 3 Ral 4


stainless steel and aluminum 1 2 3 4 5 6 7
thermal system. The linear graph
is split between nodes 6 and 7 for
this figure Css C al1 Cal 2
θ1(t)

Ral 4 R al 5 Ral 6 R al 7 Ral 8


6 7 8 9 10 11

Cal 2 Cal 3 Cal 4 θ2(t)

Fig. 5.54  a Resistances in series.


b Equivalent resistance
a b
R1 R2 R equiv
1 2 3 1 3

Fig. 5.55  a Resistances in


a b
parallel. b Equivalent resistance R1 R equiv
1 2 1 2

R2

ments. When elements in series or in parallel are combined of the equivalent element. We will derive the equivalent
and replaced by one equivalent element, the sum of either electrical resistance. The equivalent resistance for the cor-
the across variables or the through variables in the original responding fluid resistances is calculated in the same way.
elements will equal the corresponding power variable in
the equivalent element. Elements in series have the same Continuity iR1 = iR2 = iRequiv
through variable. Conversely, elements in parallel have the
same across variable. Compatibility v12 + v23 = v13
The values of the equivalent parameters are determined
by writing the continuity, compatibility, and elemental equa- Elements v12 = R1iR1 v23 = R2 iR2 v13 = Requiv iRequiv
tions for the portion of the linear graph, which contains the
similar elements to be replaced. The set of equations is re- Reduction
duced to the form of the elemental equation, written in terms
v12 + v23 = v13 → R1iR1 + R2 iR2 = Requiv iRequiv
of the through and across variables for the equivalent ele-
ment.
R1iRequiv + R2 iRequiv = Requiv iRequiv → ( R1 + R2 ) iR equiv
= Requiv iRequiv

5.9.1  Fluid, Electrical, or Thermal Resistances Requiv = R1 + R2

Resistances in series sum to yield the equivalent single re-


sistance. Resistances in parallel sum as inverses to yield the
inverse of the single equivalent resistance. 5.9.1.2 Fluid, Electrical, or Thermal Resistances In
Parallel
5.9.1.1  Fluid, Electrical, or Resistances In Series Elements in parallel have the same across variable, Fig. 5.55.
Elements in series have the same through variable, Fig. 5.54. The sum of parallel elements’ through variables equals the
The sum of their across variables equals the across variable through variable of the equivalent element.
302 5  Fluid, Electrical, and Thermal Systems

Fig. 5.56  a Fluid and thermal


capacitances must be refer-
a 1 b 1 c 1 d 1
enced to a ground pressure or
temperature. Two fluid or two
thermal capacitances in parallel.
b Single equivalent fluid or
thermal capacitance. c Electri-
cal capacitances do not need a
C1 C2 Cequiv C1 C2 Cequiv
ground reference. Two electrical
capacitances in parallel. d Single
equivalent electrical capacitance

g g g g

Continuity iR1 + iR2 = iRequiv electrical capacitor. In fact, charge carriers (electrons and
positive ions) do not pass through the dielectric (insulator)
Compatibility v12 = v12 separating the oppositely charged plates of an electrical ca-
pacitor. However, the opposite polarity charge carriers flow
Elements v12 = R1iR1 v12 = R2 iR2 v12 = Requiv iRequiv into the capacitor on opposite sides, moving in opposite di-
rections. Fluid capacitances have a single connection which
Reduction serves as both the inlet and outlet ports, depending on the
direction of the flow. Fluid cannot flow through a fluid ca-
v12 v12 pacitance without violating the physical model, by overtop-
iR1 + iR2 = iRequiv → + = iRequiv
R1 R2 ping a standpipe or rupturing a fluid accumulator.
 1 1  R R  The net result of the physical differences between fluid
iRequiv =  +  v12 → iRequiv =  2 + 1  v12 and electrical capacitances is that energetic elements in cir-
 R1 R2   R1 R2 R1 R2 
cuits and linear graphs have different constraints. Electri-
cal capacitance is more “flexible,” and can be placed any-
R1 + R2 R1 R2 where in a circuit, except in direct connection with a voltage
iRequiv = v12 → v12 = iR
R1 R2 R1 + R2 equiv source. Electrical capacitances do not need to be referenced
to ground. Consequently, electrical capacitances can be con-
R1 R2
Requiv = nected in series. Fluid capacitances must be referenced to
R1 + R2
ground. Hence, fluid capacitances cannot be connected in
series.
5.9.2  Fluid and Electrical Capacitance Thermal capacitance is similar to fluid capacitance, in that
it is referenced to ground temperature. Consequently, neither
Fluid and electrical capacitance both store energy as a func- thermal capacitances nor fluid capacitances can be connected
tion of their across variables, fluid pressure, and electrical in series. Attempting to connect a thermal capacitance in se-
voltage. However, fundamental physical differences be- ries simply results in a larger heterogeneous thermal capaci-
tween fluid pressure and electrical voltage, as well as those tance. Another difference between thermal and fluid capaci-
between volume flow rate and electric current, limit the tance is the energy equation of each. Energy stored in a fluid
analogy between them. The pressure in a fluid capacitance capacitance is proportional to the square of the pressure in
must be measured relative to a minimum “ground” pressure, the capacitor. Heat stored in a thermal capacitance is linearly
either atmospheric pressure or a complete vacuum. Sub- proportional to an ideal thermal capacitance.
stantial negative pressures can be created on a small scale
under careful laboratory conditions, by creating tension in 5.9.2.1 Fluid, Electrical, and Thermal Capacitances
liquids in capillary tubes. However, negative pressure below In Parallel
an absolute vacuum does not exist in practice, except for Elements in parallel have the same across variable, Fig. 5.56.
the “melt” strength of polymers which have some tensile The sum of their through variables equals the through vari-
strength when fluid. Low-molecular-weight fluids cavitate, able of the equivalent element.
or boil, before they reach negative pressure.
Volume flow rate and current (electric charge flow rate) Continuity iC1 + iC2 = iCequiv
are also fundamentally different from each other. While there
are both positive and negative charges, there is only posi- Compatibility v12 = v12
tive volume flow rate. Current appears to flow through an
5.9  Equivalent Elements in Fluid, Electrical, and Thermal Systems 303

Fig. 5.57  a Two electrical


capacitances in series. b Single
a b
equivalent electrical capacitance C1 C2 C equiv
1 2 3 1 3

Fig. 5.58  a Parallel fluid iner-


tances, b Equivalent single fluid
a I1 b I equiv
inertance, c Parallel electrical
inductances, d Equivalent single 1 2 1 2
electrical inductance

I2
c L1 d L equiv
1 2 1 2

L2

dv12 dv12 dv12 1 C + C2 C1C2


Elements iC1 = C1 iC2 = C2 iCequiv = Cequiv = 1 → Cequiv =
dt dt dt Cequiv C1C2 C1 + C2

Reduction

dv12 dv dv 5.9.3  Fluid Inertance or Electrical Inductance


iC1 + iC2 = iCequiv → C1 + C2 12 = Cequiv 12
dt dt dt
5.9.3.1 Fluid Inertance or Electrical Inductance
dv12 dv
(C1 + C2 ) dt
= Cequiv 12
dt
→ Cequiv = C1 + C2 In Parallel
Elements in parallel have the same across variable. The sum
of their through variables equals the through variable of the
5.9.2.2  Electrical Capacitances In Series equivalent element, Fig. 5.58.
Elements in series have the same through variable. The sum
of their across variables equals the across variable of the Continuity iL1 + iL2 = iLequiv
equivalent element, Fig. 5.57.
Compatibility v12 = v12
Continuity iC1 = iC2 = iCequiv
diL1 diL2 diLequiv
Elements v12 = L1 v12 = L2 v12 = Lequiv
Compatibility v12 + v23 = v13 dt dt dt
dv12 dv23 dv13
Elements iC1 = C1 iC2 = C2 iCequiv = Cequiv Reduction
dt dt dt
Reduction diL1 diL2 diLequiv
iL1 + iL2 = iLequiv → + =
dv12 dv23 dv13 dt dt dt
v12 + v23 = v13 → + =
dt dt dt
v12 v12 v 1 1 1
+ = 12 → + =
iC1 iC2 iCequiv iCequiv iCequiv iCequiv L1 L2 Lequiv L1 L2 Lequiv
+ = → + =
C1 C2 Cequiv C1 C2 Cequiv
1 L2 L 1 L1 + L2
= + 1 → =
1 1 1 1 C2 1 C1 1 Lequiv L1 L2 L1 L2 Lequiv L1 L2
+ = → + =
C1 C2 Cequiv C1 C2 C2 C1 Cequiv
L1 L2
Lequiv =
L1 + L2
1 C2 1 C1 1 C2 C 1
+ = → + 1 =
C1 C2 C2 C1 Cequiv C1C2 C1C2 Cequiv
304 5  Fluid, Electrical, and Thermal Systems

Fig. 5.59  a Fluid inertances in


series. b Equivalent single fluid
a b
inertance. c Electrical inductanc- I1 I2 I equiv
es in series. d Equivalent single
electrical inductance 1 2 3 1 3

c d
L1 L2 L equiv
1 2 3 1 3

5.9.4 Fluid Inertances or Electrical Inductances rarily established across a coil of low resistance wire, when
In Series there is an attempt to change the current through the coil,
except when the current is recognized as storing energy in
Elements in series have the same through variable, Fig. 5.59. a magnetic field. The storage or remove of energy requires
The sum of their across variables equals the across variable time.
of the equivalent element. Fluid, electrical, and thermal systems all have resistance
and capacitance. The mathematical analogy among the resis-
Continuity iL1 = iL2 = iLequiv tances in three types of physical systems is that the element
equation is expressed as a proportion between the power
Compatibility v12 + v23 = v13 variables of the system. The physical analogy between fluid
diL1 diL2 diLequiv and electrical resistance is strong, since the physical analo-
Elements v12 = L1 v23 = L2 v13 = Lequiv gies between pressure and voltage, as well as those between
dt dt dt volume flow rate and current, are strong. The physical anal-
Reduction ogy of dissipating heat due to friction or a friction-like phe-
nomenon in a flow restriction describes both electrical and
diL1 diL2 fluid resistance. Thermal resistance is resistance to flow. Re-
v13 = v12 + v23 → v13 = L1 + L2 ducing the cross-section of the flow and lengthening the flow
dt dt
path have the same effect as in electrical or fluid resistance.
diLequiv diLequiv diLequiv However, thermal resistance cannot dissipate energy as heat,
v13 = L1 + L2 → v13 = ( L1 + L2 )
dt dt dt because the energy in a thermal system is heat. Further, the
Lequiv = L1 + L2 power cannot flow into a thermal resistance. Power can only
flow through a thermal resistance.
An electrical capacitance does not need to be referenced
Two or more fluid inertances in series combine similarly, to ground. The energy stored in an electrical capacitance is a
I equiv = I1 + I 2 . function of the voltage across the capacitor, regardless of the
voltages of the two plates. Other than a fluid system’s capac-
itor’s reference to ground pressure, the physical analogies
Summary between the electrical and fluid capacitors are strong. One
can imagine working to cram additional fluid into a fluid
The physical and mathematical analogies between electrical accumulator, and working to cram additional charge onto
and fluid systems are very strong, but not perfect. The physi- the plates of an electrical capacitor against every increase
cal and mathematical analogies across thermal, electrical, in pressure or voltage. The mathematical analogy between
and fluid systems are quite limited. the fluid and electrical capacitor is perfect. Although one can
Fluid inertance is a helpful physical analogy for electri- imagine trying to cram additional heat into a thermal capaci-
cal inductance. The need to accelerate and decelerate a mass tance, the mathematical analogy is absent, because energy
of fluid, in order to store or remove kinetic energy, is a fact storage in a thermal capacitor is a linear function of tempera-
that we have all experienced. Electrical inductance is a new ture, not a function of temperature squared.
concept to many students. It is baffling that voltage is tempo-
Problems 305

Table 5.4   US Customary to SI Unit Conversions R1


1  in = 2.54  cm = 25.4  mm 1  MPa = 145  psi
1  gallon = 3.79  L 1  horsepower = 746  W
1 Btu = 1, 055 J
v(t) L R2

C
Fig. P5.2   RL circuit schematic
R1
P(t)
vent to
atmosphere
Pump
R2 R1 C

Fluid
Reservoir

v(t) R2
Fig. P5.1   Fluid system schematic

Fig. P5.3   RC circuit schematic


Problems

Reminders Problem 5.1  A fluid system consisting of two fluid resistanc-


1. Use nodal notation for the across variable drop, and indi- es, R1 = 100 MPa · sec/m3 and R2 = 200 MPa · sec/m3 , and a
cate the positive direction of the through variable in the fluid accumulator (capacitance), C = 6 × 10 −8 m3 /Pa, is acted
direction of the drop in the across variable. Sources are upon by a pressure source, Fig. P5.1. Resistance R2 discharg-
the exception. Sources raise the across variable in the es to atmospheric pressure in the system’s reservoir.
direction of the through variable “flow.” 5.1.a Derive the system equations for:
2. Check units in terms of the power variables of the system, i The volume flow rate from the pressure source.
not fundamental units or conventional units. ii The volume flow rate into the fluid accumulator.
3. Although system dynamics calculations can be performed iii The pressure in the fluid accumulator.
using US customary units, the set of US units can lead to and check their units
significant errors, due to the mishandling of the gravita- The capacitor is de-energized at time, t = 0 −. A step
tional constant. The units of mass moment of inertia are a change in pressure, P (t ) = 200 kPa us (t ) is applied at time,
case in point. US customary units commonly used ounce t = 0.
in2, lb in2, lb ft2, and slug ft2. It is easier to perform the 5.1.b Solve the system equations of part a.
calculations in SI Units and then convert to US customary 5.1.c Plot the responses using Mathcad or MATLAB
units to present the results, if so required.
4. A dynamic system has only one characteristic equation, Problem 5.2 An RL electric circuit where R1 = 10 Ω,
independent of our choice of the output variable. Conse- R2 = 20 Ω, and L = 3 × 10 −3 H is shown schematically in
quently, there is a time constant of the system, only one. Fig. P5.2.
All power variables in the system vary at the same rate, 5.2.a Derive the system equations for:
although their step responses will differ. Some variables i The current through the inductor.
will grow, while others will decay, but they do so in syn- ii The voltage drop across the inductor.
chronizativon. iii The current through resistor R2.
5. Remember to convert from kPa and MPa to Pa for SI and check their units.
units. The circuit is de-energized at time, t = 0 − . At t = 0, a step
6. There are 1,000 L in 1 m3. input voltage v(t ) = 48 VDC us (t ) to the circuit.
7. Engineers practicing in the United States must be able to 5.2.b Solve the system equations of part a.
convert between US customary units and SI units without 5.2.c Plot the responses using Mathcad or MATLAB.
resorting to a reference or the internet. Before you leave
school, commit to memory at least one conversion factor Problem 5.3  An electric circuit with R1 = 1 k Ω, R2 = 2 k Ω,
for each unit you will work with. The most common con- C = 1 µ F and is shown schematically in Fig. P5.3.
version needed is length, i.e., 1 in = 2.54 cm = 25.4 mm , 5.3.a Derive the system equations for:
Table 5.4.
306 5  Fluid, Electrical, and Thermal Systems

R1 R1 L
vent to
atmosphere p(t)
Pump
R2 I
Fluid
v(t) R2
Reservoir

Fig. P5.4a   Fluid circuit schematic


Fig. P5.5   RL circuit schematic

2,000 R1
p(t)
psi
1,000
v(t) C R2
0
0 1 2 3
t, sec
Fig. P5.6a   RC circuit schematic
Fig. P5.4b   Input pressure pulse

i The current from the voltage source.


i The current through the capacitor. ii The voltage across resistor R1.
ii The voltage across the capacitor. iii The current through the inductor.
iii The voltage across resistor R1. iv The voltage across the inductor.
and check their units. v The current through resistor R2.
The circuit is de-energized at time, t = 0 −. A step input of and check their units.
voltage v(t ) = 48 VDC us (t ) is applied to the circuit at time, The circuit is de-energized at time, t = 0 − . At time, t = 0,
t = 0. a step input voltage v(t ) = 48 VDC us (t ) is applied to the
5.3.b Solve the system equations of part a. circuit.
5.3.c Plot the responses using Mathcad or MATLAB. 5.5.b Solve the system equations of part a.
5.5.c Plot the responses using Mathcad or MATLAB.
Problem 5.4 The fluid system shown schematically in
Fig. 5.4a consists of a fluid power unit modeled as a pres- Problem 5.6 An electrical system consisting of a voltage
sure source, two contractions modeled as fluid resistanc- source, two resistors, R1 = 10 Ω and R2 = 20 Ω, and a capac-
es, R1 = 100 MPa · sec/m3 and R2 = 200 MPa · sec/m3 , itor, C = 1 × 10 −6 F is shown in the schematic, Fig. P5.6a.
and a long run of piping modeled as a fluid inertance, 5.6.a Derive the system equations for:
I = 40 × 106 kg/m 4 . i The current from the voltage source.
5.4.a Derive the system equations for: ii The voltage across resistor R1.
i The volume flow rate from the pressure source. iii The current through the capacitor.
ii The pressure difference across the fluid resistance, iv The voltage drop across the capacitor.
R1. v The current through resistor R2.
iii The volume flow rate through the fluid inertance. and check their units.
iv The pressure difference across the fluid inertance. The voltage applied to the system is plotted in Fig 5.6b.
and check their units. 5.6.b Solve the system equations of part a.
The system was de-energized before the pressure pulse 5.6.c Plot the responses using Mathcad or MATLAB.
shown in Fig. 5.4b was applied to the system.
5.4.b Solve the system equations of part a. Problem 5.7  An electric circuit with R1 = 1 k Ω, R2 = 2 k Ω,
5.4.c Plot the responses using Mathcad or MATLAB. C = 1 µ F is shown schematically in Fig. P5.7.
5.7.a Derive the system equations for:
Problem 5.5 An electric circuit with two resistors, i The current through the capacitor.
R1 = 10 Ω and R2 = 20 Ω, and an inductor, L = 3 × 10 −3 H , ii The voltage across the capacitor.
is shown schematically in Fig. P5.5. iii The voltage across resistor R1.
5.5.a Derive the system equations for: and check their units.
Problems 307

v(t), VDC R1
v2 = 36
R2

v1 = 12 R3
v(t) C
0 time

Fig. 5.6b   Input voltage Fig. P5.7   RC circuit schematic

R1
The circuit is de-energized at time, t = 0 −. A step input of
voltage v(t ) = 15 VDC us (t ) is applied to the circuit at time,
t = 0. R2
5.7.b Solve the system equations of part a. C1
5.7.c Plot the responses using Mathcad or MATLAB.
v(t) C2
Problem 5.8  An electric circuit with R1 = 1 k Ω, R2 = 2 k Ω,
C1 = 1 µ F and C2 = 2 µ F is shown schematically in Fig. P5.8.
5.8.a Derive the system equations for:
i The current through the capacitors. Fig. P5.8   RC circuit schematic
ii The voltage across the capacitors.
iii The voltage across the resistors.
Capacitance, C
iv The current through resistor R1.
v The current through resistor R2. -
vi The voltage across capacitor C1. +
vii The voltage across capacitor C2.
and check their units. Resistance R1
The circuit is de-energized at time, t = 0 −. A step input of
Voltage source
voltage v(t ) = 15 VDC us (t ) is applied to the circuit at time,
t = 0.
Coil around ferrite core with
5.8.b Solve the system equations of part a. Induction L and Resistance R 2
5.8.c Plot the responses using Mathcad or MATLAB.
Fig. P5.9a   Electrical system

Problem 5.9  The electrical system shown in Fig. P5.9a con-


sists of voltage source, a resistor with resistance, R1 = 100 Ω,
and a coil wound around a ferrite core. The coil has both 24
inductance, L = 3 × 10 −3 H and resistance, R2 = 20 Ω. v(t)
5.9.a Derive the system equations for: VDC
12
i The current from the voltage source.
ii The voltage across resistance R1.
iii The current through the inductor. 0
0 2 4 6
iv The voltage across the inductor. t, millisec
v The current through resistance R2.
and check their units. Fig. P5.9b   Input voltage pulse
The system was de-energized before the voltage pulse
shown in Fig. P5.9b was applied to the system. sure source, two contractions modeled as fluid resistances
5.9.b Solve the system equations of part a. R1 = 100 MPa · sec/m3 and R2 = 200 MPa · sec/m3, a long run
5.9.c Plot the responses using Mathcad or MATLAB. of piping modeled as a fluid inertance, I = 40 × 106 kg/m 4
and a fluid accumulator (capacitance), C = 6 × 10 −8 m3 /Pa .
Problem 5.10 The fluid system shown schematically in
Fig. 5.10a consists of a fluid power unit modeled as a pres-
308 5  Fluid, Electrical, and Thermal Systems

L
C
R1 I
P(t)
vent to Pump v(t) C R
atmosphere R2
Fluid
Reservoir
Fig. P5.11a   RLC circuit schematic

Fig. P5.10a   Fluid circuit schematic


v(t), VDC

2,000 v(0+) = 36
p(t)
psi
1,000 v(0-) = 12

0 time
0
0 1 2 3
t, sec Fig. 5.11b   Input voltage

Fig. P5.10b   Input pressure pulse


draws water from a reservoir at atmospheric pressure. The
5.10.a Derive the system equations for: pump is modeled as a pressure source in series with the in-
i The volume flow rate from the pressure source. ternal resistance of the pump. The pressure of the source is
ii The pressure in the fluid accumulator. 150 psi. The maximum steady-state flow rate of the pump is
iii The volume flow rate through the fluid inertance. 1250 gallons per minute when discharging directly to the at-
iv The pressure difference across the fluid inertance. mosphere. The maximum steady-state flow rate of the pump
and check their units. discharging through the pipe and nozzle is 800 gallons per
The system was de-energized before the pressure pulse minute.
plotted in Fig. P5.10b was applied to the system. 5.12.a What is the internal resistance of the pump?
5.10.b Solve the system equations of part a. 5.12.b What is the internal diameter of the pipe in inches, if
5.10.c Plot the responses using Mathcad or MATLAB. the system reaches 70 % of its steady-state flow in 2
sec after the nozzle valve is opened?
Problem 5.11  An electrical system consisting of a voltage
source a resistor, R1 = 1 Ω, a capacitor C = 1 × 10 −6 F , and an Problem 5.13 An electric circuit consisting of a voltage
inductor, L = 3 × 10 −3 H , is shown in the schematic P5.11a. source, two resistors, R1 = 10 Ω and R2 = 2 Ω, a capacitor
5.11.a Derive the system equations for: C = 2 × 10 −6 F, and an inductor, L = 0.015 H, is shown as the
i The current through the inductor. schematic Fig. P5.13a.
ii The voltage across the inductor. 5.13.a Derive the system equations for:
iii The current through the capacitor. i The voltage across the capacitor.
iv The voltage across the capacitor. ii The current through resistor R1.
and check their units. iii The current through the inductor.
The system was in steady-state under the previously ap- iv The voltage drop across the inductor.
plied step input of 12 VDC when a step change was made at v The voltage across resistor R2.
time, t = 0, increasing the voltage to 36 VDC, as shown in and check their units.
Fig. 5.11b. The system is initially de-energized before the voltage
5.11.b Solve the system equations of part a. pulse shown in Fig. P5.13b was applied.
5.11.c Plot the responses using Mathcad or MATLAB 5.13.b Solve the system equations.
5.13.c Plot the responses using Mathcad or MATLAB.
Problem 5.12  A firefighting system consists of a pump and
a 135 ft. long, horizontal, circular cross-sectioned steel pipe, Problem 5.14  The electric circuit shown in Fig. P5.14 con-
terminated with a nozzle, as shown in Fig. P5.12. The pump sists of a voltage source, a resistor, R = 1 Ω, a capacitor,
Problems 309

Internal Resistance
Fig. P5.12   Firefighting system of the Pump, R Internal

135 ft
p(t)
Pump Pipe, Inertance I Nozzle
vent to
Resistance R Nozzle
atmosphere

Fluid
Reservoir

L C R

v(t) C
v(t) R1 R2

Fig. P5.13a   RLC circuit schematic Fig. P5.14   RLC circuit

10 L1

5
L2
v(t) C1
0
VDC 1 2 3 4
t, millisec
-5 v(t) R C2

-10

Fig. P5.15a   RLC circuit schematic


Fig. P5.13b   Input voltage

R = 10 Ω, two capacitors, C1 = C2 = 2 × 10 −6 F , and two in-


−6
C = 2 × 10 F, and coil with 1.4 millihenrys of ­inductance ductors, L1 = L 2 = 0.015 H ,
and 2 Ω of resistance. 5.15.a Derive the system equations for:
5.14.a Derive the system equations for: i The voltage across capacitor C2.
i The voltage across the capacitor. ii The current through resistor R.
ii The current through resistor R. iii The current through inductor L1.
iii The current through the inductor. iv The voltage drop across the inductors.
iv The voltage drop across the inductor. v The voltage across resistor R.
and check their units. and check their units.
The system is initially de-energized before a step input of The system is initially de-energized before the voltage
24 VDC was applied. pulse shown in Fig. P5.15b was applied.
5.14.b Solve the system equations. 5.15.b Solve the system equations.
5.14.c Plot the responses using Mathcad or MATLAB. 5.15.c Plot the responses using Mathcad or MATLAB.

Problem 5.15 The electric circuit shown in the sche- Problem 5.16 A hydraulic system consists of a pump
matic, Fig. P5.15a, consists of a voltage source, a resistor, modeled as a pressure source which draws fluid from a
reservoir vented to the atmosphere, two fluid resistances,
310 5  Fluid, Electrical, and Thermal Systems

100
C1 C2
50
R1 R2
v(t)
p(t)
Pump
0
VDC 1 2 3 4 vent to
atmosphere R3
t, millisec
-50 Fluid
Reservoir

-100
Fig. P5.17a   Fluid system schematic
Fig. P5.15b   Input voltage

Fluid Power Unit Modeled as 3,000


Pressure Source p(t)

Breather Cap (vent to atmosphere)


2,000
High Pressure Output p(t)
psi
Hydraulic Line 1,000
Resistance R1

Low Pressure Return Fluid Accumulator


Capacitance C
0
0 1 2 3
t, sec
Hydraulic Line
Inertance I and
Resistance R 2
Fig. P5.17b   Input pressure pulse

Fig. P5.16   Fluid system ii Design a cubical steel box for the reservoir, such
that there is 2 in. of air above the fluid. Round up
the dimensions to the nearest inch.
R1 = 100 MPa · sec/m3 and R2 = 200 MPa · sec/m3, and a
fluid accumulator (capacitor), C = 0.0040 m3 /MPa . The re- Problem 5.17 The hydraulic system schematic shown
turn line is 3/4 in. I.D. and 40 feet long. The density of the in Fig. 5.17a consists of a pump modeled, as a pressure
hydraulic oil is approximately 875 kg/m3. source, which draws fluid from a reservoir vented to the
5.16.a Derive the system equations for: atmosphere, three fluid resistances, R1 = 100 MPa · sec/m3 ,
i The pressure in the fluid accumulator. R2 = 200 MPa · sec/m3, and R3 = 300 MPa · sec/m3, and two
ii The volume flow rate into the fluid capacitor. fluid accumulators (capacitors), C1 = 0.04 m3 /MPa and
iii The volume flow rate through fluid resistance R2. C2 = 0.05 m3 /MPa.
iv The volume flow rate from the pump. 5.17.a Derive the system equations for:
and check their units. i The pressure in the fluid accumulator, C1.
5.16.b Calculate the fluid inertance of the return line ii The volume flow rate into the fluid capacitor, C1.
The system was running in steady-state with the pump iii The pressure in the fluid accumulator, C2.
pressure of 1,800 psi when, at time, t = 0, the pressure was iv The volume flow rate into the fluid capacitor, C2.
given a step increase to 3,000 psi. v The volume flow rate through fluid resistance R2.
5.16.c Solve the system equations of part a and plot the vi The volume flow rate from the pump.
responses using Mathcad or MATLAB. and check their units.
5.16.d The fluid reservoir of a fluid power unit is sized so The system was de-energized, when it was given the pres-
that the residence time, the average time that fluid sure pulse plotted in Fig. P5.17b.
remains in the reservoir during operation, is two min- 5.17.b Solve the system equations of part a and plot the
utes, in order to degas the hydraulic fluid. responses using Mathcad or MATLAB.
i Determine the capacity of the pump needed to
provide 125  % maximum flow calculated in Problem 5.18  The fluid system shown in the Fig. P5.18a
part c. consists of a pump, modeled as the pressure source, P( t),
Problems 311

Rcoil L coil
C
R1
P(t) v(t)
vent to Pump 1 µF
atmosphere
R2
Fluid
Reservoir

Fig. P5.19a   RLC circuit schematic

Fig. P5.18a   Fluid system schematic


Determine the unknown fluid resistance, R2 and capaci-
tance, C. Report your results in SI units.
250

Problem 5.19 The RLC circuit shown schematically in


200
Fig. P5.19a consists of a capacitor in series with 2 m of 22
Q pump 150
AWG solid copper wire wound around a ferromagnetic core.
gal The resistance 22 AWG solid copper wire is 16.8 ohms per
___
min 100 1,000 ft. Note the coil has two energetic properties, resis-
tance and induction. Induction enhanced by the presence of
50 a ferromagnetic material leads to energy loss, due hysteresis
of the magnetic field in the ferromagnetic material. Induc-
0 tion enhanced by ferromagnetic material is also non-linear.
0 1 2 3 4 5 6 7 8
The circuit was charged to 15 VDC. Then the connection
time, seconds
between the voltage source and the circuit and the end of the
Fig. P5.18b   Pump test data coil was grounded. The response of the voltage across the
capacitor was captured on an oscilloscope. A portion of the
response is shown in Fig. P5.19b.
two fluid resistances, R1 and R2, and fluid capacitance, C. 5.19.a Use the value of the resistance of the copper wire
The pump’s reservoir is vented to the atmosphere. The value above to determine the capacitance and inductance
of the fluid resistance, R1 is known, R1 = 300 MPa · sec/m3. of the circuit.
The values of the parameters R2 and C are unknown and to be 5.19.b Calculate the value of the apparent resistance, Rcoil,
determined from a dynamic test. The pump was running in from the observed response of the circuit. Determine
steady-state with a pressure of 500 psi when, at time t = 2.0 the capacitance and inductance of the circuit.
sec, the pressure was suddenly increased to 1,000 psi. 5.19.c How different are the values calculated in parts a and
The pump’s pressure and volume flow rate are plotted, b? If the difference is significant, what may account
Fig.  P5.18b, and tabulated, Table P5.18, in US customary for the discrepancy?
units.

Table P5.18   Pump Test Data Time (sec) Pump pressure (psi) QpumpGPM Time (sec) Pump pressure (psi) QpumpGPM
0  500  45.5 3.1 1000 123
1.0  500  45.5 3.2 1000 119
2.0 1000 228 3.3 1000 115
2.1 1000 211 3.4 1000 112
2.2 1000 196 3.5 1000 110
2.3 1000 183 3.6 1000 107
2.4 1000 171 3.7 1000 105
2.5 1000 161 3.8 1000 103
2.6 1000 152 3.9 1000 102
2.7 1000 145 4.0 1000 101
2.8 1000 138 5.0 1000  94
2.9 1000 132 6.0 1000  92
3.0 1000 127 7.0 1000  91
312 5  Fluid, Electrical, and Thermal Systems

Fig. P5.19b   The voltage across 1.5


the capacitor during the latter
portion of the discharge of the
RLC circuit
1.0

0.5

vcap ,VDC 0

-0.5

-1.0
0.0 0.1 0.2 0.3 0.4 0.5
time, milliseconds

Chapter 5 Appendix ning on their own axis. Magnetic fields in electrical ma-
chines are created by electric currents flowing along con-
Engineering Electromagnetics ductors.
We will start with the fundamental phenomenon of force,
A basic understanding of magnetic phenomena is essential F, induced on a positive charge, q, in a conductive rod, by
for mechanical engineers, given the importance of electric moving the conductor with a constant velocity, v, at right
motors in machine design. We will begin with the basics of angles to a uniform magnetic field, B, Fig. A5.1. This force
electromagnetics. With all physical phenomena, a “realis- is referred to as “induced” by the motion of the conductor
tic” engineering model is a matter of degree. We will keep in the magnetic field. The force causes an accumulation of
our model of electromagnetics simple and ideal, so that it is positive charge at the top of the rod, as shown in Fig. A5.1.
generally representative and applicable, but does not require Negative charges are driven in the opposite direction, and
(much) vector calculus. In any case, the non-linearity and accumulate at the bottom of the rod. The strength of the
uncertainty of the material properties, magnetic permeability “electromotive force” (EMF) is expressed in volts. A volt is
in this case, necessitates the use of a simple model, unless the defined as the strength of electromotive force against which
engineer has the need and the means to fully characterize the a joule of the work is done, to move a coulomb of positive
materials he or she is working with. charge. A coulomb of charge is defined as charge delivered
We will introduce an engineering model created to ex- by a one ampere of current in one second.
ploit the similarity between a resistive electric circuit, and We are accustomed to forces which are collinear with
the steady-state behavior of an electromagnet. We then in- motion, such as a force acting to accelerate a mass, m, per
vestigate the non-linear magnetic properties of the ferromag- FM = Ma , and the force on a charge, q, due to an electric
netic materials we use to enhance the strength of the mag- field E, per FE = q E. A force sideways to the direction of
netic fields we create for electric motors. We conclude this motion seems odd at first, but we also deal with non-collin-
chapter with the topic of flux linkage, which we will use in ear forces in mechanics. Centrifugal and centripetal forces
energy method analyses of electric motors. act at right angles to the direction of rotational motion. A
challenge in developing a working understanding of electro-
magnetic phenomena lies in applying analyses previously
Electromagnetic Force used in Dynamics and Fluid Mechanics to a new subject.
It is helpful for mechanical engineering students to keep
All magnetism is electromagnetism. The charge can be an in mind that electromagnetic phenomena were discovered
electron orbiting an atomic nucleus or an electric current. and characterized by the forces associated with them. Our
The magnetic fields of permanent magnets are created by primary interest in these phenomena is that we need those
the alignment of the individual magnetic fields of electrons, forces, and the torques which can be created, to drive our
which are in motion orbiting their atomic nuclei and spin- machines. A secondary benefit in studying these phenomena
Chapter 5 Appendix 313

Force on Positive Charge q magnetic field B


F=qv B thumB
Uniform magnetic
field B into the Speed (velocity) of charge
plane of the paper
Second finger

+ Positve Charge
+
+
induced Force
+
Conductor
First finger
Velocity v
Fig. A5.2   Right-hand rule mnemonic for the cross ­product F = q v × B

-
-
-
- Negative Charge
The small step from the fundamental physical phenomena,
illustrated in Fig. A5.1, to an electrical machine, a genera-
tor, or, the inverse, an electric motor, is to provide a path to
Fig. A5.1   Creation of electromotive force by the motion of a conductor
across a uniform magnetic field complete a circuit from the top to the bottom of the conduc-
tive rod moving in the magnetic field, Fig. A5.3.
Charge no longer accumulates at the ends of the mov-
is the strengthening of important analytical skills which have ing conductor since the charge is now free to flow along the
wide application. stationary conductor. The EMF creates a current, i, which
The cross product, F = q v × B, defines the relative ori- flows through the stationary conductor from the positive end
entations of the force, the velocity of the charge, and the of the moving conductor to the negative end. The strength,
magnetic field. A mnemonic is often helpful to visualize the or voltage v, of the EMF is proportional to both the velocity,
force induced on a charge during its motion through a mag- v, of the moving conductor and the strength of the magnetic
netic field. The mnemonic relates the thumb and first (index) field, B.
and second (middle) fingers of the right hand, held such Power is converted from mechanical power to electrical
that they form three orthogonal axes, to the three vectors in power
F = q v × B, Fig. A5.2. (This mnemonic allows engineering
students to make rude gestures and claim they were only P�
= Fext · v = iv
finding the direction of the induced electromotive force.) We
need to use “speed” as the synonym of “velocity.” omitting the negative sign, which would be included using
linear graph sign convention.
thumB = B First finger = Force  If the externally applied force is the input to this system,
Second finger = Speed = velocity then this device is a generator converting mechanical power

Fig. A5.3   A Simple Generator Uniform magnetic field B into the plane of the paper
or Electric Motor
positive
+
Moving conductor
creates EMF and
current i

Externally applied force Fext Velocity v Current i

-
negative

Stationary conductor creates closed circuit


314 5  Fluid, Electrical, and Thermal Systems

i1 dL 1 i1
cross product yielding
permeability the magnetic field
L1 of free space acting on i 2 dL 2

^
r12
dFmagnetic = i 2 dL 2 ×
µ0
_____
4π r122
(i dL × r )
1 1 12

i2
Point A L2 cross product unit vector from
yielding the force 2 is the area dL1 to dL2
4πr12
acting on i 2dL 2 of a sphere with
Fig. A5.4   Two current carrying conductors. Current element i1dL1
radius r12

to electrical power. If current is supplied by an external Fig. A5.5 Annotation of the equation for the differential magnetic
source, then the device is a linear electric motor. force acting between two current elements

Electromagnetic Force between Two Current of a vacuum. The denominator term, 4π r122 , is the surface
Elements area of a sphere with radius, r12 .
Although it is conventional to place scalar terms in front
The simple motor of Fig. A5.3 assumes the existence of a of the cross products, the equation is easier to interpret when
uniform magnetic field. ­Although some electrical machines it is rearranged as:
contain permanent magnets, these are limited in size. The
motion of all large electrical machines is created by the elec- µ0
tromagnetic force between two current carrying conductors, d Fmagnetic = i2 d L 2 × (i1d L1 × rˆ 12 )
4π r122
either copper wires or bars. Both conductors create magnetic
fields. Although we will sometimes find it convenient to
µ0
perform calculations in terms of the interaction of the two  d Fmagnetic = i 2 d L 2 ×
4π r122
(i d L
1 1 × rˆ12 ) (5.60)
magnetic fields, we will begin with the more fundamental
calculation of the electromagnetic force, expressed in terms
of the two currents. The equations are formidable, because and annotated as in Fig. A5.5.
two cross products are required. What makes this equation difficult to visualize are the two
The electromagnetic force acting between two current-car- cross-products, Fig. A5.6. The cross-product, i1d L1 × rˆ12 , is
rying conductors separated by air or in a vacuum, Fig. A5.4, is normal to the plane of i1d L1 and r12, when it is scaled by
expressed in the following differential vector equation. µ0
the ratio, , it is the magnetic field vector at the end of
4π r122
 µ0 the position vector, r12. The “density” of the magnetic field
d Fmagnetic = i2 d L 2 × (i1d L1 × rˆ 12 ) (5.59)
4π r122 created by i1d L1 decreases as the inverse square of distance
from i1d L1. Consequently, the density of the magnetic field
where rˆ 12 is the unit vector in the direction from element d L1 created by i1d L1 is constant on any sphere centered at i1d L1.
to d L 2. µ0
Equation 5.59 is a differential equation, to allow us to ad-
The second cross-product, i2 d L 2 × (i1d L1 × rˆ12 ), is
4π r122
dress the general case, in which either wire 1 or wire 2 or the magnetic force acting on i2 d L 2. The vector, d Fmagnetic, is
both are not straight. Infinitesimal wire segments, d L1 and
normal to the plane of i2 d L 2 and i1d L1 × rˆ12.
d L 2, allow us to express geometric relationships needed
between quantities with lengths and directions, which are,
therefore, vectors. However, the physical interaction occurs
between moving charges, not infinitesimal lengths of wire. The Magnetic Field B
Current is the flow rate charge. The product of current and
the length of a conductor are used, in place of charge mea- The cross-product
sured in coulombs. The constant, µ0, is the “permeability of
free space,”where “free space” is a physics term meaning a  dB =
µ0
4π r122
(
i1d L1 × 
r12 ) (5.61)
volume free of material, i.e, a complete vacuum. The mag-
netic permeability of air, µ air , is approximately equal to that
Chapter 5 Appendix 315

sphere with
i1 dL 1 i1
radius r12
centered at i1dL 1 s=L
s=0 ^
i1dL1 r 12

i2
r12 Point A L2

i1dL1 × r 12 Fig. A5.8  Electromagnetic force acting at point A on current carrying


i2dL 2 conductor L2 from infinitesimal current element, i1dL1

Fig. A5.6  The cross product i2 d L 2 × (i1d L1 × rˆ12 )


L0
__
2
i1dL1 s idL
plane of cross-product (
i 2 dL 2 × i1dL1 × r 12 ) 0
θ
L0
i1dL1 × r 12
r12 r x
i1dL1 × r 12
Point A
i2dL 2
plane of cross-product Fig. A5.9   Magnetic field density at a distance x from a straight-wire-

(
i 2 dL 2 × i1dL1 × r 12 ) carrying current i

Be careful with the vector notation. The difference between


Fig. A5.7  The cross product i2 d L 2 × (i1d L1 × rˆ12 )
the last two expressions is easily overlooked. The vector rˆ12
is the unit vector in the direction of r12, whereas the vector r12
is the infinitesimal portion of the magnetic field, dB, at the r12, is a position vector with magnitude, r12 = r12, Fig. A5.8.
created by the i1d L1, and scaled by the permeability of free
r12
space, µ0 , and divided by the geometric factor, 4π r122 , equal  rˆ12 = (5.63)
r12
to the surface area of a sphere with radius, r12 . The integral
form of this equation is sometimes credited to Ampere but
more often Biot, or both Biot and Savart, as the Biot–Savart As one can easily imagine, integrals of this type range from
law. It dates from 1820. It is a “magnetostatic” approximation straightforward to very challenging, depending on the geom-
requiring steady currents. Time-varying currents, such as etry of wire L1. We will deal with two simple geometries that
AC, accelerate and decelerate charges, causing them to radi- allow us to approximate all situations of interest to us. Wire
ate electromagnetic energy. The magnetostatic approxima- L1 will be either straight or bent into a circle.
tion also neglects the dimensions and characteristics of the
conductors. Magnetic Field Density Calculations
If we express r12 and r122 as functions of distance, s, along
wire, L1, then an evaluation of the resulting line integral over Magnetic Field Density at a Distance x from a Straight-
the length of wire, L1, yields the magnetic field density, B, at Wire-Carrying Current i
point A on wire L2, The case of a field surrounding a straight wire of length L0 in
air is shown in Fig. A5.9. Let us consider a location at a dis-
tance x from the wire in the normal plane passing through the
µ0 µ i d L1 × rˆ12
2 (1
B=∫ i d L1 × rˆ12 ) = 0 1 ∫ wire’s midpoint, at s = L0 /2.
4π r12 4π r122
L0 L0
µ0 µ i dL × rˆ
 µ0 i1 d L1 × r12 B= ∫ (idL × rˆ ) = 0 ∫
4π ∫
B= (5.62) 0
4π r 2 4π 0
r2
r123
d L × rˆ = d L rˆ sin (θ ) = d L sin (θ )
316 5  Fluid, Electrical, and Thermal Systems

1.0 L0
µ i  2α − L0 2
B = 0 2
4π  x 4 x + L0 − 4 L0 s + 4 s 
0.8
L0
________ 2 2 2
 0
4x 2 + L 02 0.6

x=1 0.4  L0  
µ i  2 − L0
2   − L0 
x=5 B = 0 2  2 

 2 

4π  2
0.2
 x 4 x 2 + L2 − 4 L  L0  + 4  L0    x 4 x + L0  
0


0
 2    
2 


0
0 5 10 15 20 25
L0
µ0 i L0
Fig. A5.10   Effect of the finite ends of a straight conductor on the B=
magnetic field density, B, on a plane normal to the midpoint of the 2π x 4 x 2 + L20
conductor. L0 is the length of the conductor. x is the distance from the
conductor
The effect of the ends of the wire on the density of the mag-
netic field at the center decreases, as the wire becomes lon-
since rˆ = 1. ger. However, “end effect” is a function of both the length
of wire L0 and the distance x of point A from wire. A plot
dL rˆ sin (θ ) dL sin (θ )
L0 L0 L0
µ0 i dL × rˆ µ0 i µ0 i L0
B=
4π ∫
0
r2
=
4π ∫
0
r2
=
4π ∫
0
r2
of the function,
4 x 2 + L20
vs. L0, for values of x = 1 and

x = 5 shows how rapidly it approaches a final value of unity,


We will exploit the symmetry of the configuration and dou- Fig. A5.10. Whether one judges the end effect to be negli-
L gible depends on the precision needed in the analysis. A con-
ble the integral we ­evaluate, from s = 0 to s = 0 . We must L0
2 venient value for our purposes is ≥ 5, which corresponds
now express r 2 and sin (θ ) in terms of x and s. The Py- x
thagorean Theorem yields the length of the position vector to an error of less than 2 %.
r squared
2 µ0 i L0 L0
2 L  B straight = for <5
r = x +  0 − s 2
wire 2π x 4 x 2 + L20 x
 2 
and
From the definition of sine and the Pythagorean Theorem,
µ0 i L0
B straight = for ≥5
x x wire 2π x x
sin (θ ) = =
r  L0 
2
L0
x2 +  − s When ≥ 5 , the magnetic field density around a straight
 2  x
current-carrying wire is inversely proportional to the dis-
L0 tance, x, from the wire. Lines of equal strength of the field
µ0 i d L sin (θ ) are circles centered on the wire, in planes normal to the axis
B=
4π ∫
0
r122 of the wire. The sense of direction of the magnetic field is
established by the cross-product, id L × r. ˆ In practice, the
L0
right-hand rule is used. One imagines grasping the wire with
2
µ0 i 1 x one’s right hand with one’s thumb pointing in the direction of
B=

2 ∫
0 2  L0 
2
 L0 
2
ds
the current. The fingers of the right hand then curl around the
x + − s x2 +  − s wire in the direction of the magnetic field lines, Fig. A5.11.
 2   2  We speak of a magnetic field as “flowing,” so that we
L0 can use familiar physical phenomena to help us visualize the
µ0 i 2 x physics of electromagnetism. The magnetic “flux,” φ (Greek
4π ∫0 
B= 2 1.5
ds for f, as in “flow”), is the total magnetic field, which passes
 
2
 L0 through either half of a plane containing the wire. The SI unit
 x 2
+  − α  
 2  for magnetic flux φ is Wb, webers. The magnetic flux φ is a
Chapter 5 Appendix 317

Magnetic field line


idL
B
Current i θ
r
i Pt. A
2R
i +α
Fig. A5.11   Circular magnetic field line B around a straight wire carrying
current i

Fig. A5.13   Point A at the center of a loop wire carrying current i


Magnetic field lines
B
Current
tributed by an infinitesimal element of the loop of wire, lies
i
on the plane normal to the infinitesimal length of wire. The
magnetic fields of adjacent ­elements of the wire overlap (or
superpose) within the loop. If the wire loop is circular, then
the planes normal to the axis of wire intersect at the center
of the wire loop.
Fig. A5.12   Magnetic field lines B, around a single loop of wire carry- We will begin with the equation for magnetic field den-
ing current i sity B

d L ×
L0 L0
µ0 µi
scalar, and is the same across any surface that cuts across the
B= ∫
0
4π r 2 (
id L × 
r = 0

) ∫0
r 2
r

magnetic field, from the wire to infinity.


The amount of magnetic flux φ passing through a unit The distance to point A is constant and equal to the radius R,
area is the “magnetic flux density,” B, B = φ / area. The SI since point A is at the center of the loop
unit for magnetic flux density B is T, teslas. Magnetic flux
density B is a vector. The variation of the magnitude and
µ0 i d L × 
L0 L0
r µ i
direction magnetic flux density B with position (or in case of B=
4π ∫0 r 2
= 0 2
4π R ∫ d L × r
a straight wire, with distance from the wire) is a vector field. 0

Rather than draw a multitude of tiny vectors to represent the


magnetic field, it is easier and more common to visualize The cross-product is also constant, since the lengths of d L
magnetic fields using “field lines”, also known by the older and 
r are constant, and the angle θ between them is constant
term, “lines of induction.” The spacing of field lines is used π
and equal to = 90o .
to convey the magnetic flux density. More tightly spaced 2
lines denote higher magnetic flux density. An important dif-
ference between electric fields and magnetic fields is that d L ×
r = dL 
r sin 90o = d L ( )
magnetic field lines always form closed loops, which is the
basis for the “flow” analogy.
π
since 
r = 1 and sin   = 1. This leaves
 2
Magnetic Flux Density B at a Distance x within a Loop
of Wire Carrying Current i L0 L0
There are two ways to strengthen the magnetic field created
µ0 i d L × r µ i
by a given current. First, we can shape the wire. For a given
B=
4π ∫
0
r2
= 0 2
4π R ∫ dL
0

current, the magnetic field inside a loop of wire is substan-


tially stronger than the magnetic field next to a straight wire, We will exploit the symmetry of this problem to simplify the
Fig. A5.12. calculation, and express the integration in terms of the angle
We will consider two cases. First, we will calculate the α, defined graphically above.
magnetic field density B at the center of the loop. Next, we
L0 2π
will formulate the general case to calculate the magnetic µ0 i µ0 i
field density B at any point within the loop. B center =
of loop 4π R 2 ∫
0
dL=
4π R 2 ∫ R dα
0
Magnetic Flux Density B at the Center of the Loop µ0 i µ iR 2π µ i

2[
The radius of the wire loop is R, Fig. A5.13. The circum- B center = Rα ] 0 = 0 2 = 0
of loop 4π R 4π R 2R
ference of the loop is L0 = 2π R. The magnetic field, con-
318 5  Fluid, Electrical, and Thermal Systems

North In electromagnetics, field lines exit the coil at the North pole
and enter the coil at the South pole. Within a coil, the mag-
netic field lines are oriented from “South” to “North.” In the
space around the coil, the field lines are curves oriented from
North to South.
Current i
Magnetic
Current i Field Lines B  agnetic Moment and Engineering
M
Approximations of Magnetic Field Density

The “solenoidal current density,” j s , is a useful quantity in


many calculations.
South
 N i mmf
js ≡ = (5.64)
Fig. A5.14   Cross-section of the magnetic field within and around a l l
coil
The term, density, is odd in this context, since the denomina-
tor is length, not area. Its use derives from the model of an
where the direction of the magnetic field is determined by ideal cylindrical conductor with zero wall thickness replac-
the right-hand rule and is out of the plane and normal to the ing the physical coil. The current Ni is distributed uniformly
plane of the paper in this example. along the length of the cylinder, and flows circumferentially
around the ideal conductor. The units of the solenoidal cur-
rent density reflect recurring confusion due to angular dis-
Coils or “Solenoids” placement and rotation lacking units. The units are

A logical extension to a single loop of wire is to “stack Ni  amp   amp·turns 


 j  =   = 
s
=
up” loops of wire, by creating a helical coil, Fig. A5.14.  l   m   m 
The coil was invented by Ampere, who coined the term,
“solenoid,”Greek for “channel,” to describe it. Today, “sole- We will use the units, [ amp · turns/m ], since they help re-
noid” commonly means a linear electromechanical actuator. mind us that the number of turns N is a factor.
The effect of strengthening the field by shaping the wire is A second derived quantity useful in describing permanent
taken to its limit with toroidal coil, since the magnetic field magnets is the “magnetic moment” of a current-carrying
lines never leave the coil. A coil-wrapped toroidal core of loop with radius R and cross-sectional area A
ferromagnetic material, discussed below, is known as a Hay-
ward ring, after the American physicist from the late 1800s. magnetic moment ≡ i π R 2 = i A

Magnetic Sign Convention Magnetic moment yields the torque exerted about the axis
Historically, magnetic phenomena were first observed with of a loop with the current i when multiplied by the magnetic
permanent magnets, called “lode stones” of the ore magnetite. field density B. If B is not normal to the plane of the current
It was first recorded in China that permanent magnets could loop, then the dot product between B and a normal to the
be created by magnetizing iron needles with a lode stone. The plane of the loop is used.
Chinese were also the first to use compasses. Well before elec- Although it is relatively easy to program magnetic field
tromagnets were invented, terminology from cartography had calculations in MATLAB, or other languages, approxima-
been applied to the orientation of a magnet field of a com- tions based on the closed-form results that were available
pass needle within the magnetic field of the earth. (The earth’s prior to machine computation generally have sufficient ac-
magnetic field is created by the motion of charge with the flow curacy for engineering. The discrepancy between the more
of liquid iron in the earth’s core.) precise numerical results and the closed-form approxima-
Unfortunately, and oddly, the sign conventions for geo- tions will be far less significant, than the uncertainty and
magnetism and electromagnetics are opposite to one anoth- error introduced by including ferromagnetic materials in the
er. In the geomagnetic sign convention, field lines enter the flux path to increase the magnetic flux density.
earth at the North pole and exit the earth at the South pole.
Chapter 5 Appendix 319

2.5
The following approximations assume that the magnetic
flux density is uniform across the radius of a loop or a coil.
2.0
They are based on the closed-form results using the Biot– saturation
Sarvart model with infinite current density for a loop or coil Magnetic
in air. Field 1.5
Density
B preferred
1.0
Single-Loop Approximations teslas operating
Magnetic flux density within a loop of radius R range
0.5

µ0 i
B= 0
2R 1 10 100 1,000 10,000 100,000
amp .turns
_____
Applied Magnetic Field Intensity, H
Magnetic flux density outside a loop of radius R at a point meter

of distance r from the center of the loop, the “far field” ap-
Fig. A5.15   DC magnetization or B–H curve
proximation

 webers   henrys 
µ 2iπ R 2 µ0 2iA µ0 = 4π ·10−7  = 4π ·10 −7 
B= 0 =

 ampere − turn − meter   meter 
4π r 3 4π r 3
Although permeability is usually referred to as a “material”
where A is cross-sectional area of the loop, and the product property, this is not strictly correct, since a magnetic field
iA is the “magnetic ­moment” of the loop. The magnetic mo- can permeate a vacuum in which there is no material. The
ment is, indeed, the moment which would act on the loop to term, 4π, is introduced for convenience when performing
align the loop with the magnetic field. calculations, since the magnetic density created by an infini-
tesimal current element obeys the inverse square law.
Coil Approximations Ferromagnetic materials, iron, cobalt, nickel, and their al-
Magnetic flux density at the middle of a solenoid of suffi- loys, have permeabilities significantly greater than that of
cient length such that the end effects can be neglected free space. We increase the strength of a magnetic field in
an electrical machine by creating a “flux path” for the mag-
µ0 Ni netic field through ferromagnetic materials. Materials have
B= = µ0 j s
l magnetic properties on the atomic level due to the motion of
their electrons. If the vectors which represent the axes mag-
Magnetic flux density at any radial location at the middle of netic dipoles created by the orbiting electron are randomly
a solenoid of sufficient length such that the end effects can oriented then there is no net magnetic field on the macro-
be neglected scopic scale. Ferromagnetic materials significantly enhance
the strength of an applied magnetic field because the mag-
µ0 Ni µ0 j s netic dipoles created by orbiting electrons in neighboring
B= =
2l 2 atoms align into “magnetic domains.” The magnetic dipoles
of the magnetic domains within the ferromagnetic material
Note that the magnetic flux density at the ends of a long coil orient in response to an externally applied field. However,
is half that at the middle. the relationship between strength of the local magnetic field
and the degree of orientation of the magnetic domains within
a ferromagnetic material is a non-linear relationship.
Magnetic Permeability μ and Ferromagnetic The non-linearity of the relationship between the applied
Materials external magnetic field strength and the resulting magnetic
field density can be seen in two plots used to characterize the
The ease with which a magnetic field can be created in a magnetic properties of materials. The first is a DC magneti-
volume or material is called the magnetic permeability, and zation, or B–H curve, which is a plot of the magnetic field
is given the symbol, μ (Greek for m). The field calculations density which would be created in a ferromagnetic material
above assumed that the volume surrounding the current- on its first exposure to an applied external magnetic field, Fig.
carrying wire was filled with air. The permeability of air is A5.15. The second non-linearity is the ferromagnetic mate-
approximately equal to that of a free space (a vacuum), µ0. rial’s “memory,” which is why a ferromagnetic coating is used
320 5  Fluid, Electrical, and Thermal Systems

for computer hard disks. The response of a material to its first 3.5

magnetization is a reproducible standard test, but doesn’t re- 3.0

flect the behavior of the material in actual use, when it may 2.5
experience thousands or millions of cycles of magnetization. Magnetic
There are three regions of interest in this “S” shape. Field 2.0
Density
Note that it is a semi logarithmic plot. The initial region B
1.5

has a low ∆B / ∆ H slope, indicating that low applied mag- teslas 1.0
netic field strength accomplishes disproportionately little
0.5
orientation of magnetic domains. Increasing H from 1 to 5
amp · turns/meter increases B from approximately 0 to 0.25 0.0

teslas. The slope ∆B / ∆ H then increases sharply, and it is 10 100 1,000 10,000 100,000 1,000,000

this region which is the preferred operating range, where we amp .turns
_____
Applied Magnetic Field Intensity, H
meter
can create a substantially more powerful magnetic field den-
sity with the least amount of current. Increasing H from 6 to Fig. A5.16   B–H curve of annealed iron, data from Sears (1951)
10 amp · turns/meter increases B from approximately 0.3 to
1 teslas. As the slope of the curve lessens with increasing ap- of a ferromagnetic material. When the applied magnetic
plied magnetic field strength, there is diminishing return for field strength that has produced a state of magnetization is
the investment in electric current. Eventually, the maximum reduced, magnetic domains lose orientation at a lower ap-
possible orientation of the magnetic domains is achieved. plied magnetic field strength than that at which they gained
The magnetic field density will continue to increase with in- orientation.
creasing applied magnetic field strength, but only at the rate The area within the hysteresis loop is energy in joules per
it would for air or a vacuum, since the ferromagnetic mate- cubic meter lost as heat to the ferromagnetic core. This can
rial can no longer “amplify” the magnetic field. be seen from the units of H and B
The magnetic properties of the ferromagnetic materials
A Wb A V ⋅ sec A ⋅ V ⋅ sec   J 
vary widely between the different materials, between two [H ][ B ] =    2  =    2  =  =
specimens of the same material processed differently, and  m   m   m   m   m3   m3 
even in a specimen, as its magnetic history changes. Anneal-
ing an alloy “resets” the magnetic domains, by providing the (5.65)
thermal energy necessary for them to return to their random-
ly oriented state. Hysteresis is important when the electrical power driving the
The B–H plot and plots of permeability and relative per- system is AC, since the energy loss due to hysteresis occurs
meability for annealed iron are shown in Figs. A5.16 and each cycle. The magnitude of the hysteresis energy loss re-
A5.17. duces the energy efficiency of the device. The factor which
The B–H curves for commercially available silicon steel constrains the amount of electrical power that can be applied
used in electrical machines often has two sets of units for to electrical machines is the temperature rise of the core and
both axes. The units kilogauss and oersteds are cgs. The units the winding or coil. Energy lost as heat due to hysteresis of
amp · turns/meter and teslas are mks. the ferromagnetic core and electrical resistance of the wind-
The magnetic properties of engineering ferromagnetic ing both contribute to the temperature rise.
materials alloys are dependent on their temperature history, Because the electrical power which can be applied to the
heat treating, or annealing. Just as one cannot have confi- machine is dependent on the internal temperature, many
dence in published strength values for steels, neither can one mechanical engineers are engaged in heat transfer analyses
have confidence in published magnetic properties, unless the of electrical equipment. An example is the computer indus-
heat treatment of a specific part is known and identical to try, where increasing density of integrated circuit chips has
that of the specimens the published data is based on. In fact, led to more elaborate cooling schemes, including aluminum
magneto-inductive non-destructive testing is used to mea- heat sinks and fans mounted directly on central processing
sure strength, surface hardness, and the depth of case harden- units. Google, the internet search corporation, chose a site
ing, since variation in these properties can be correlated with directly adjacent to a river for its main computing facility,
variation in electrical conductivity and magnetic permeabil- because of the need to cool its machines.
ity. Conversely, manufacturing processes which affect these Some magnetic domains will retain orientation at zero ap-
mechanical properties also affect electrical conductivity and plied magnetic field strength, creating a “permanent” mag-
magnetic permeability. net, Fig. A5.19. A permanent magnet can be demagnetized
The second important plot of magnetic behavior is a hys- by reversing the externally applied magnetic field. The per-
teresis plot, Fig. A5.18. Hysteresis arises from the “memory” manent magnets used in electrical machines were magne-
Chapter 5 Appendix 321

Fig. A5.17   Magnetic perme- 0.008 6,000


ability and relative permeability
of annealed iron, data from Sears 5,000
(1951) 0.006
Permeability 4,000
µ Relative
Permeability
weber
______• amp 0.004 3,000
meter µr
2,000
0.002
1,000

0.000 0
10 100 1,000 10,000 100,000 1,000,000
amp • turns
_____
Applied Magnetic Field Intensity, H
meter

B B
teslas
teslas

“residual inductance”
amp.turns
______
“remanence” or Br
H meter “retentivity”
amp.turns
H ______
meter

“coercive force” Hc
Fig. A5.18   Magnetic hysteresis loop. H is the intensity of the applied
magnetic field. B is the magnetic flux density Fig. A5.19   Portion of hysteresis loop showing the residual magnetic
flux density, Br, at zero applied magnetic field intensity, Hc

tized carbon steel or cobalt steel, until the second half of the
twentieth century, when the “harder,” or more-difficult-to- Table A5.1   Relative Permeability µr
demagnetize ferromagnetic alloys “Alnico” (for aluminum, Iron 150 to 5,000
nickel, and cobalt), were developed. Magnetic ceramic ma- Nickel 110 to 600
terials, which include rare-earth elements, were first intro- Cobalt 70 to 250
duced in the 1960s. Although they are mechanically brittle, Carbon Steels 50 to 100
they have superior magnetic properties. Keep in mind even Cast Irons 100 to 750
Stainless Steels 1 to 3,000
the best “permanent” magnet can be demagnetized, by ac-
Electrical (silicon alloy) Steels 2,000 to 6,000
cidentally reversing the polarity of a coil.
The scale of magnetic permeability makes its numerical
values somewhat difficult to work with. Calculations are
simplified, and errors are reduced, by the “relative perme- Low-cost barium-based ferrites have relative permeabilities
ability” of a ferromagnetic material, where “relative” is with in the range of cast iron. Their advantage over cast iron is
regard to the permeability of free space. The “relative perme- their substantially lower electrical conductivity and, hence,
ability” of a material, µr, is defined as less heating due to eddy currents. Ferrites formulated to en-
hance magnetic fields under low frequency AC or DC ex-
 µ citation typically have relative permeabilities between 750
µr ≡ (5.66)
µ0 and 10,000, but some range up to 100,000. At the extremes,
magnetic glasses can be manufactured with relative perme-
Handbook and Web values of the relative permeability for abilities of 1,000,000. Ferrites can have both high magnetic
common engineering materials are reported as ranges, due to permeability and high electrical resistivity, as opposed to
the effects of processing on crystal structure. electrical steel which has relatively low electrical resistivity.
“Ferrites”are ceramics composed of barium, manganese, Unwanted currents induced in electrical machinery by time-
and/or zinc and iron oxide, with other materials as “binder.” varying magnetic fields are called “eddy currents.” Eddy
The relative permeability of ferrites varies widely, since currents are an energy loss and can be a significant prob-
some are custom-formulated for different applications. lem in electrical machines because of the resulting resistive
322 5  Fluid, Electrical, and Thermal Systems

heating. The high electrical resistivity of ferrites reduces  φ


B = (5.67)
eddy currents. A

where A is the cross-sectional area of the magnetic flux path.


 agnetic Flux Density B, Flux φ, and Applied
M In order to calculate the magnetic flux density when the
Magnetic Field Intensity H flux path comprises materials with different permeabilities,
we must introduce the vector field, H, called the “applied
The relative permeability of the ferromagnetic materi- magnetic field intensity” or “magnetic intensity.” The mag-
als in the magnetic flux path within an electrical machine netic flux density, B, is proportional to the vector field, H.
(ferrite, electrical steel, carbon steel, and cast iron) is large The proportionality constant is the magnetic permeability, μ.
enough that the portion of the magnetic flux outside of the
ferromagnetic material can be neglected. Ferromagnetic ma-  B = µH (5.68)
terials can be thought of as not only intensifying the mag-
netic field but also channeling it, since the magnetic field Why work with B and H rather than just B? Why add the
outside of the ferromagnetic material is generally negligibly complication of another vector field, if B and H are just
small, compared to the field inside the ferromagnetic ma- scalar multiples of each other? There are two reasons. First,
terial. Hence, we will model the magnetic flux as confined although we will make the engineering approximation that
to, or channeled by, the ferromagnetic materials of the ma- there is a single value of permeability for a given material, as
chine, with one important exception. Electric motors must shown above, the relationship between the applied magnetic
have gaps of some finite width between their stationary and field intensity, H, and the resulting magnetic flux density,
moving elements. The effect of gaps which are in the flux B, is non-linear and multi-valued. The contribution of fer-
path must be included in our analyses, as they are always romagnetic materials to the magnetic flux density, B, eventu-
significant. ally is also limited. The limit, called “saturation,”is reached
Magnetic flux density B is a vector field. Perhaps “mag- when all domains are oriented to the applied magnetic field.
netic field” would be a more straightforward, modern name, Any further increase in the applied magnetic field intensity,
but, as noted above, the “lines of induction,” or magnetic H, does increase the magnetic flux density, B, but the incre-
field lines, form closed curves. As a result, our engineering mental increase is only at the permeability of free space, µ0,
models are based on a flow analogy. The term, flux density, as if the ferromagnetic material were not present.
reinforces the flow analogy. Second, some electrical machines of interest, specifical-
As we calculated, the magnetic flux density is non-uni- ly some stepping motors and the type of AC motor used in
form across the radius of a coil with an air core. The mag- some hybrid automobiles, use rotors which are “permanent”
netic flux density cannot be expected to be more uniform, magnets. The magnetic flux of rotating motors flows through
when the core is a ferromagnetic material with a non-linear the rotor. The permanently magnetized rotor and the exter-
permeability. It is, in fact, generally a great deal less uni- nally applied magnetic field together produce the magnetic
form. We will, nonetheless, make the additional engineer- flux through the motor. The field, H, is used to account for
ing approximation that the magnetic flux density is uniform both effects.
across any surface through the flux path. A realistic model Our analysis will exploit the similarities between the flow
of the magnetic field within an electrical machine requires a of magnetic flux through an electrical machine, and the flow
finite-element model based on a well-characterized ferro- of electric current in an electrical circuit comprising a volt-
magnetic material. However, as is the case in stress analysis, age source and resistors. The analogy, called a “magnetic
premature use of finite-element models electromagnetics is circuit,” draws the parallel between (1) voltage (electromo-
poor engineering, as they tend to lead to well analyzed, bad tive force) in an electric circuit and magnetomotive force in
designs. Far simpler models are always used for the initial a magnetic circuit; and (2) the resistance-to-current flow in
design. They greatly simplify calculations, while maintain- an electric circuit and amount of magnetomotive force need-
ing an acceptable accuracy to parse through design alterna- ed to establish a magnetic flux φ in a magnetic circuit. The
tives. Once a candidate design is identified, then a finite- drop in voltage in the direction of current flow in an elec-
element model may be appropriate. For our purposes, the tric circuit is proportional to the resistance of the elements
simple models will suffice. in the circuit. Similarly, in a magnetic circuit, the drop in
By confining the magnetic flux path to the machine’s the magnetomotive force in the direction of the “flow” of
“iron” (ferromagnetic materials) and the gaps between the flux φ is proportional to the “reluctance,” R, of that portion
stationary and moving elements, and assuming the flux den- of the magnetic circuit. Reluctance R  is a nineteenth-century
sity is uniform across the path, we can relate magnetic flux φ term, coined to draw the analogy between a purely resistive
and flux density B as:
Chapter 5 Appendix 323

φ φ

Coil1 Coil 2 Coil1 Coil 2


N1 turns N2 turns N1 turns N2 turns

i i i i

Magnetic flux φ “path” Surface defined by the


assumed on centerline magnetic flux φ path
Path length L c “cutting” the conductors
Ferromagnetic core
Cross-sectional area A Fig. A5.21   Ferromagnetic core with two coils cut by surface across
flux path
Fig. A5.20   One wire wrapped in two coils in series on a ferromagnetic
core
(counter-clockwise when looking down on a horizontal
cross-section). Consequently, current i creates mmf in the
electric circuit and the magnetic circuit model. The magnetic same direction through both coils, Fig. A5.21. Hence, the
circuit model is developed below. mmf creating the magnetic field is

mmf = mmf1 + mmf 2 = N1i1 + N 2 i2 = ( N1 + N 2 ) i


Magnetic Circuit Model

Magnetomotive Force, MMF We increase the current density without increasing the
We will now simplify our calculation of the magnetic flux amount of current, by increasing the number of turns in the
density produced by coils, by creating a lumped parameter coil.
model. We will create a pseudo-circuit and, to make it useful, We will begin with the case, in which the mmf in the cir-
we will simplify the geometry of the flux path and linearize cuit is created by coils, and the flux path is through a ferro-
the magnetic phenomenon. magnetic core with constant cross-section. Our second case
Coils are used in all motors to intensify the magnetic field will include a gap in the flux path. The third case will include
created by a current. The number of turns of wire in the coil, a permanent magnet.
N, times the current flowing through the wire, i, is known H is a vector field in the direction of the magnetic flux
by the nineteenth-century term “magnetomotive force,” ab- density, B, but of a different, and much smaller, magnitude.
breviated as “mmf.” Excluding the presence of a permanent magnet in the flux
path, the magnitude of magnetic field intensity, H, is the
 mmf = N i (5.69) drop in the magnetomotive force, Δmmf, in the direction of
the flow of flux φ, divided by the length of the segment flux
We will model the magnetic field through motors using what path, ∆l , where ∆l is a vector pointing in the direction of the
is known as a “magnetic circuit,”in which the mmf drives the flux φ.
“flow” of the magnetic field in analogy to a voltage source,
also known as electromotive force, driving the flow of elec-  ∆ mmf
H= (5.70)
tric current. ∆l
Although we will see that the ferromagnetic materials
used in electrical machines have non-linear magnetic proper- Magnetic field intensity H is the intensity of the “effort” re-
ties, we will use the engineering approximation of superposi- quired at a location along a flux path to maintain a magnetic
tion, and add the individual mmf of multiple current-carrying flux φ. For a given drop in magnetomotive force Δmmf along
coils to find the net mmf acting to “drive” the magnetic field. path of length ∆l , the magnetic flux φ increases with increas-
A ferromagnetic “core” with one wire wrapped in two ing magnetic permeability μ and increasing cross-sectional
coils in series is shown in Fig. A5.20. The direction of the area A of the flux path. The dependence of the magnetic flux
mmf produced by each core is determined by the right-hand φ on a “material” property μ and the cross-section area A
rule. (Imagine grasping the wire with one’s thumb pointing allow us to build analogies between the creation of a mag-
in the direction of the current. The magnetic field circles the netic flux φ and potential driven flows of fluid, heat, or elec-
wire in the direction of one’s fingers.) The same current, i, tricity.
passes both coils, which are wrapped in the same “sense”
324 5  Fluid, Electrical, and Thermal Systems

Table A5.2   Basis of magnetic Fluid Electrical Thermal


circuit analogy. Δl is the length of
Flow variable Volume flow rate, Q Current, i Heat flow rate, q
the flow or flux path and A is its
cross-sectional area. Driving potential Pressure, p Voltage, v ( emf) Temperature, T
Flow-related material Fluid permeability, k Electrical conductivity, κ Thermal conductivity, k
property
Flow resistance Fluid resistance, R f Electrical resistance, R Thermal resistance, RT
Elemental parameter

Flow equation ∆ p = Rf Q ∆ emf ≡ ∆v = iR ∆T


q=
RT
Flow resistance param- ∆l ∆l ∆l
eter relationship Rf = R= RT =
kA κA kA

The corresponding flow equation which relates the mag- it is negative. In this magnetic circuit, Coils 1 and 2 both
netomotive force to the resulting magnetic flux φ uses the contribute positive mmf. The total mmf is
resistance-like magnetic reluctance, R
mmf = N1i1 + N 2 i2
 mmf = R φ (5.71)
Since the same current flows through both Coil 1 and Coil 2,
where
Magnetic Reluctance i1 = i2 = i
Flux PathLength l
R = =
Permeability ⋅ Cross − Sectional Area µ A Hence,

Case I Magnetic Circuit with a Uniform Ferromagnetic mmf = N1i + N 2 i = ( N1 + N 2 ) i


Core
A ferromagnetic core has two coils connected in series, as We now use the definition of magnetic reluctance R to es-
shown in Fig. A5.20. Express the magnetic flux φ as a func- tablish its value.
tion of current i. We will use the magnetic circuit model:
lc
R =
mmf = R φ µ Ac

We will first establish the relationship between current i Substituting into the magnetic circuit equation
and the mmf in the magnetic circuit. How do we handle
two coils on the same flux path? We use superposition. The l
total, or net, mmf acting to establish (or drive, if you prefer) ( N1 + N 2 ) i = µ cA φ
the magnetic flux in the circuit is the sum of the mmf of the c

two coils and rearranging

mmf = mmf1 + mmf 2  µ Ac 


( N1 + N 2 )  i=φ
 lc 
Be careful with the direction of the mmf. The assigned posi-
tive direction for the magnetic flux φ is shown on the figure Example Magnetic Circuit
as clockwise. Magnetomotive force which acts in that direc- Consider the toroidal ferrite core and coil shown in
tion is positive. Magnetomotive force which acts in the op- Fig. A5.22. A coil of N = 200 turns of wire, carrying a cur-
posite direction is negative. Use the “right-thumb rule” to rent, i = 3A, creates an electromagnet with a magnetomotive
establish the direction of the mmf. Imagine grasping a turn force, mmf = N i . The length of the flux path, called the core
in Coil 1 with the thumb of your right hand pointing in the length, lc , is approximated as the circumference of the circle
direction of the current. If, on the side of the wire adjacent to with the mean radius of the torus, r = 0.5 in . There is a ra-
the ferromagnetic core, your fingers curl around the wire in dial gap through the torus with gap length, g = 0.020 in. The
the clockwise direction, then the mmf is positive, otherwise torus has cross-sectional area Ac = 0.0625 in 2. The relative
Chapter 5 Appendix 325

gap length g = 0.020 in Core Reluctance


Rc
1 2
Torrodial ferrite core
Relative permeability µr
Core length l c φR
φ c
φR Gap
Cross-sectional area A
φ
mmf φ g Reluctance
Rg
i
i
3
200 turns of wire in the coil Fig. A5.23    Magnetic circuit model. The magnetomotive force,
mmf = Ni = 200 turns times 3  amps = 600 A-turns
Fig. A5.22   Ferrite torus with a gap of length g = 0.020 in . and a 200-
turn coil

The magnetic reluctances of the core, R c, and the gap, R g,


permeability of the ferrite is µ r = 15, 000. Calculate the mag- are in series. They sum to create the reluctance of the mag-
netic flux φ and the drop in mmf across the gap. netic circuit, R .
We will model this system as a magnetic circuit Fig. A5.23.
We assume that all of the magnetic flux φ is confined to the R = R c +R g
ferrite core, except where the flux path crosses the air gap.
The mmf force is analogous to a voltage source. The magnet- Electrical and magnetic calculations are performed using SI
ic reluctances of the ferrite torus and the air gap are in series (or, sometimes cgs) units. We need to convert the dimensions
and are analogous to series electrical resistances. given in US customary unit into SI.
The mean radius of the torus is 0.5 in. The circumference
Continuity Node 1  φ = φR c   Node 2  φR c = φR g of the flux path, including the air gap, is
Circumference = π D = 2π r
Compatibility mmf13 = mmf12 + mmf 23
 1m 
Elements mmf12 = R c φR c   mmf 23 = R g φR g Circumference = 2π  0.5 in = 0.080 m
 39.37 in 
Energy No energy storage elements.
The length of the air gap in meters is
Before we can perform a reduction, we must first calculate
the magnetic reluctances of the ferrite torus and the air gap.  1m 
lg = 0.020 in  = 0.00051 m
The calculation of the magnetic reluctances is analogous to  39.37 in 
the above calculation of the electrical resistance of a solid
copper wire. The “material property,” magnetic permeabil- The core length is the length of the flux path minus the gap
ity, takes the place of electrical conductivity. The geometric length
quantities of cross-sectional area and flux-path length have
the same influence on the magnitude of the magnetic flux, as lc = 0.080 m − 0.00051 m = 0.07949 m
they do on the magnitude of the current in the analogous re-
sistive circuit. Increasing the cross-sectional area of the flux The cross-sectional area of the core is assumed to equal the
path increases the flux “pushed” by a given mmf. Increasing cross-sectional area of the flux path through the gap. The
the length of the flux path decreases the flux. Calculation of cross-sectional area in square meters is
the magnetic reluctance, R , and the electrical resistance, R,
use relationships of the same functional form.  6.452 × 10 −4 m 2  −4
Ac = 0.0625 in 2   = 4.03 × 10 m
2

 1 in 2
Length
Electrical Resistance R =
κ · Area The relative permeability of the ferrite is µ r = 15, 000. Re-
and call that relative permeability is defined as:
Length µ
Magnetic Reluctance R = µr ≡
µ · Area µ0
326 5  Fluid, Electrical, and Thermal Systems

where the permeability of free space is µ0 = 4π × 10 −7 henry/m.


 Rg 
The symbol for henry is L. Take care that the unit for mmf 23 =   Ni
inductance does not accidently morph into a length. Like-  Rc + R g 
wise, take care that the unit for current, A, does not morph
into an area A. This result has the same form as the calculation of the volt-
age across a resistor in a voltage divider.
We now calculate both reluctances using the relationship,
Length
R = .  A 
µ · Area 1.00 × 107
 weber 
mmf 23 =  600 A ⋅ turns
R g calculation. Recall that the magnetic permeability of A A 
 1.01 × 105 + 1.00 × 107 
air is approximately equal to that of free space.  weber weber 

Length Length  A 
Rc = = 1.00 × 107
µ · Area µ0 · µ r · Area  weber  600 A · turns = 594 A · turns
mmf 23 = 
A 
 1.01 × 107 
0.07949 m  weber 
Rc =
 4π × 10 −7 henries 
 m
(
 (15, 000) 4.03 × 10 m
−4 2
) It is important to recognize that 99 % of the applied mmf
is expended in creating the magnetic field across the gap.
A Reducing the length of gaps in electrical machines greatly
R c = 1.05 × 105
weber increases the amount of electrical power they can convert to
mechanical power.
Note that the magnetic reluctance of the gap is 100 times
larger than the magnetic reluctance of the core. Calculation of the Magnetic Field Created
Reduction 1 Input  mmf13 = 600 amp − turns, Output φ by the Poles of Permanent Magnet
Permanent magnets are used in stepper motors, DC servo-
mmf13 = mmf12 + mmf 23 → mmf13 = R c φc + R g φ g motors, and, more recently, in AC motors used by hybrid
automobiles. A permanent magnet in the flux path increases
mmf13 Ni
(
mmf13 = R c + R g φ ) → φ= =
Rc + R g Rc + R g
the mmf in the magnetic circuit. A straight forward method
for including the mmf, contributed by a permanent magnet in
a magnet circuit, is to represent the permanent magnet by the
600 A ⋅ turns equivalent “surface current,” which is a fictitious coil with
φ= the current that would create the same magnetization in the
A
5 A
1.05 × 10 + 1.00 × 107 volume, occupied by the permanent magnet.
weber weber
We will assume that the strength of a permanent magnet
600 A ⋅ turns is known. However, in practice, the strength of a permanent
φ= = 5.9 × 10 −5 webers
7 A magnet is rarely known and must be measured. Briefly, the
1.01 × 10
weber standard technique for measuring the strength of a magnetic
field is to observe the dynamic response of a test coil first
“Turns,” the number of turns of the coil, is dimensionless. placed in and then quickly removed from the magnetic field.
It is common for coils in motors, solenoids, and transforms The test coil used is known as “search,” “flip,” or “snatch”
to have many hundred turns of fine wire, so as to create a coil. The terms “flip” and “snatch” coils are descriptive of
substantial mmf with a small current. the test process. In the case of a “flip” coil, a dynamic re-
Reduction 2 Input  mmf13 = 600 amp − turns, Output sponse of the test coil is created, by “flipping” or rotating the
mmf 23 orientation of the test coil from normal to the field lines to
parallel to the field lines, thereby reducing the flux linkage
mmf 23 = R g φR g → mmf 23 = R g φ quickly to zero. Similarly, a snatch coil is quickly moved,
or “snatched,” from the unknown field to a location where
 mmf13   Rg  the field is known. An alternate technique is to measure the
mmf 23 = R g   → mmf 23 =   mmf13 force exerted by the permanent magnet, which, as we will
 Rc + R g   Rc + R g 
Chapter 5 Appendix 327

see when we discuss electric motors, is directly related to the North


__
magnetization. m rN
There are three equivalent ways of expressing the strength Point A
of a permanent magnet. As noted above, a permanent magnet
can be expressed in terms of the equivalent surface current, __
m rN
________
which would replicate the magnetic field density, B. Alterna- HN = __ 3
__
tively, a permanent magnet can be described using the quan- __
- m rS 4πµ0 |rN |
tity, “magnetization M,” which is the magnetic moment per
rS H S = ________
__ 3
4πµ0 |rS |
unit volume. Lastly, the strength of a permanent magnet can
be described in terms of the strength of the magnetic “poles.” m
The magnetization M of a cylindrical permanent magnet South
with length l and cross-sectional area A is defined in terms of
the equivalent surface current as Fig. A5.24   Magnetic field intensity components at Point A due to
point magnetic poles
 ( Ni )s
Magnetization ≡ M = µ0 (5.72)
l
of its size relative to the distance to a location of interest, is
where the subscript, “s,” denotes the equivalent surface cur- calculated as
rent (or fictitious coil).
Note the equivalent surface current is assumed to mag-  j
1 m ri
netize a volume with the permeability of free space, µ0. It H=
4πµ0 ∑r 3
(5.75)
is instructive to compare this expression for magnetization i=0 i
with the expression for the magnetic flux density, B, in a core
of length lc with permeability, μ. where the position vector, r, is directed from the pole to the
point of interest, when m is a North pole and directed from
 Ni the point of interest to the pole, when m is a South pole. Note
B= µ (5.73) that pole strength m scales the position vector r, i.e.,
lC
 1 m ri
Hence, magnetization M is the magnetic flux density of the Hi = (5.76)
4πµ0 ri 3
permanent magnet.
The third way to describe the strength of a permanent
magnet is in terms of magnetic poles, Fig. A5.24. The North and ri point in the same direction. There is no vector product
and South magnetic poles are where the field lines (or lines in this calculation.
of induction) exit and enter the surface of a body, respec- Magnetic intensity H is a linear vector quantity. If the cur-
tively. The spacing of field lines is a method of graphically rent flowing through a coil is doubled, then the magnitude of
conveying flux density. The “pole strength,” m, is indicated the vector field representing the magnetic intensity is dou-
by the number of field lines. Conversely, the pole strength bled. It is important to note the magnetic flux density vector
per unit area is field is non-linear, in general, due to the non-linear magnetic
permeability of ferromagnetic materials.
 Pole Strength m
≡ = M ≡ Magnetization (5.74)
Area A Magnetic Circuit Calculation Including a Permanent
Magnet in the Flux Path
Magnetization M is a vector quantity. Its direction is deter- “Salient poles” on a motor’s stator (the stationary electro-
mined by the normal to the area, if the cross-sectional area magnetic components fixed to a motor’s frame) are coils
is a plane, and whether the magnetic pole is North or South. with ferromagnetic cores, which extend toward the rotor, so
If the cross-sectional area is not planar, then trigonometry, as to minimize the gap width and concentrate the magnetic
geometry, or vector Calculus would be needed to determine flux, Fig A5.25. Salient poles are found in some convention-
the direction of the magnetization vector, M. We will ap- al electric motors and all “stepper,” or “stepping,” motors. A
proximate the surface of the pole as a plane to simplify our stepper motor rotates from one equilibrium position to the
calculations. next, as the sets of coils on its salient poles are energized
The magnetic intensity, H, in a region surrounding a mag- with currents by a logic circuit, following an algorithm based
netic pole which can be considered a “point pole,” because on the type of stepper motor, permanent magnet, hybrid, or
328 5  Fluid, Electrical, and Thermal Systems

Stator, ferromagnetic, permeability µ using these values, the fictitious solenoid current, and the
Coil, N turns
input current.
Step 1 Determine the mmf of the permanent magnet. In the
i
l1 abstract, the magnet’s strength is expressed in terms of pole
Rotor g strength, m
N
permanent magnet, l2 I.D. O.D.
pole strength = m S
g Pole Strength m
i l1 ≡ = M ≡ Magnetization
Area A
Coil, N turns
w
We estimate the pole area, neglecting the curvature of the
Cross-section depth d into plane of paper rotor, as

Fig. A5.25   An electromagnetic model of one pair of poles of a perma- Apole = wd


nent magnet or hybrid stepper motor

Hence, the magnetization, M, is


variable (switched) reluctance, and its use, high torque, high
m m
speed, or fine position control. Briefly, stepper motors differ M= =
A wd
by
• the number of sets of coils,
• whether there are two coils on one salient pole, and it is directed out of the North pole of the permanent mag-
• how many poles are spaced around the circumference of net.
the stator, In practice, we cannot easily measure the pole strength
• the number of “stacks” of poles in the stator’s axial direc- of a permanent magnet stepper motor directly. Instead, we
tion, work backward from the fact that the magnetization, M, of
• whether the rotor is a permanent magnet, and, if it is, the a permanent magnet is equal to its flux density, B, which we
arrangement of the magnetic domains, and calculate from the magnetic flux φ:
• whether the rotor is “castellated” (grooved to form small
salients).
φ
All stepper motors rotate to reduce the reluctance of the flux M=B=
A
path, thereby increasing the magnetic flux. However, only
stepper motors with a non-magnetized rotor are known as
“variable reluctance” stepper motors. Stepper motors with We determine the magnetic flux φ from the relationship be-
permanently magnetized rotors fall into two types, on the tween the rate of change of the flux linkage, λ, and the volt-
basis of how the rotor is magnetized. The rotor of a “per- age induced on the motor’s coil, when the permanent magnet
manent magnet” stepper motor has magnetic domains which rotor is spun at a known angular frequency
reverse polarity 48 times in the circumferential direction,
Fig. A5.32. These motors are the least expensive type. The
dλ dφ
rotor of “hybrid” stepper motor is polarized in the axial di- v1g = → v1g = N
rection, Fig. A5.33. The term, “hybrid,” refers to the castel- dt dt
lation of the rotor, which increases the variation in reluctance
hence the torque produced by the motor. We are given the permeability of the permanent magnet as
A schematic of one pair of poles of a permanent magnet µ m. Therefore, we can calculate the magnetic field intensity,
stepper motor is shown in Fig. A5.26. We will express the H, which the equivalent coil must provide to replace the per-
magnetic flux φ through the rotor as a function of current i. manent magnet in the magnetic circuit.
The analysis can be broken into three steps. First, we must
determine the mmf created by the permanent magnet. The
m
magnet’s mmf is then summed (superimposed) with the mag- = M = B = µ m H equiv
wd
netomotive force created by the coils to yield the mmf which
drives the flux. Second, we create a magnetic circuit with the
reluctances of the gaps, salient poles, stator, and the mag- Recall the applied magnetic field intensity, H, is the mmf
netic core to calculate the flux through the system. Third, we with drops in the direction of the magnetic flux over a length
calculate the reluctances from the geometry and magnetic l in a material with permeability, μ
permeabilities of the motor, and express the magnetic flux
Chapter 5 Appendix 329

5
i
i
φ N
N 4
3 is
S Ns is
2
i 1
i N

Fig. A5.26   Magnetic flux paths of the electromagnetic stepper motor 6


model
Fig. A5.27   Permanent ­magnet of the rotor replaced by the equivalent
solenoid
mmf mmf lB
H= → B=µ → mmf =
l l µ
We can now establish the fictitious solenoid’s current using
Substituting in the permeability, length, width, depth, and our previous result.
pole strength of the permanent magnetic yields
l2m
mmf mag = ( Ni )S ≡ N S iS = 2 NiS =
l2 m µm w d
mmf mag = ( Ni )S =
µ m wd
1 l2 m
iS =
where the subscript, “S,” on the equivalent mmf represents 2 N µ m wd
the fictitious solenoid, Fig. A5.27.
The magnetic flux density in the permanent magnet, when The single equivalent mmf is:
it is the magnetic circuit, is not equal to magnetization M! The
 1 l2 m 
magnetization, M, is the flux density, when the flux path of mmf equiv = 2 N (i + iS ) = 2 N  i +
the magnet is through air. The flux density in the permanent  2 N µ m wd 
magnet, when it is in the magnetic circuit, will be higher, even
when there is no current through the coils, because the flux The equivalent mmf source will be placed in series with the
from the magnet flows through a ferromagnetic core of high magnetic reluctances anywhere along the central path of the
magnetic permeability, rather than through air. We must calcu- magnetic circuit. An mmf node must be added to the circuit
late the flux through the magnet in the next step. to accommodate the mmf source element. Let us add node
Step 2 Derive the equivalent magnetic circuit using the 7 between nodes 6 and 1, resulting in the magnetic circuit
mmf of the two physical coils and the fictitious coil to drive shown in Fig. A5.28.
the magnetic flux, φ. First, identify the magnetic flux paths. Due to the symmetry of the motor, all of the reluctances,
Next, identify mmf nodes where the magnetic flux divides except that of the permanent magnet, R m , appear twice. The
or joins; or, where there is a change in the magnetic perme- reluctances of the poles and the gaps are in series. The reluc-
ability or cross-section of the flux path. (We will add mmf tances of the two sides of the stator are in parallel. We will
nodes to account for the sources of mmf, after we combine reduce this circuit “graphically,” by combining reluctances
the coils, to simplify the magnetic circuit.) in series and reluctances in parallel, until there is just one
We will now simplify the magnetic circuit further, by de- equivalent reluctance and mmf source in the circuit. Recall
termining the single source of mmf equivalent to the three that reluctances can be combined with the same rules as with
coils. The physical coils which carry the same current i can resistors. Reluctances in series sum
be summed
R pole + R gap + R m + R gap + R pole ≡ R equiv
series
mmf1 + mmf 2 = Ni + Ni = 2 Ni
2R pole + 2R gap + R m ≡ R equiv
series
To include the fictitious solenoid representing the permanent
magnet in the single equivalent mmf, we must let NS equal Reluctances in parallel sum as inverses:
twice N
1 1 2 1 Rstator
+ = ≡ → R equiv =
NS = 2N Rstator Rstator Rstator R equiv parallel 2
parallel
330 5  Fluid, Electrical, and Thermal Systems

R pole R gap Rm R
1 2 2R pole+2R gap+R m+
_____
stator
7 3
2

mmf
R stator R pole R gap 7 6
6 5 4
mmf
R stator
Fig. A5.30   Final magnetic circuit model of the motor with a single
Fig. A5.28   Magnetic circuit model of the permanent magnetic stepper equivalent magnetic reluctance
motor
Flux lines exit Fringing Flux lines enter
2R pole+2R gap+R m North pole South pole

7 5 North South
φ φ
mmf
R
_____
stator
6
2
Fringing
Fig. A5.29   Simplified magnetic circuit model of the motor
Fig. A5.31   Flux fringing at an air gap

Redrawing the linear graph using the values of the two


equivalent reluctances yields a series circuit, Fig. A5.29.
Length g Length l1
The final simplification is to sum the two equivalent re- Rg = = R pole = =
luctances in series to yield the circuit shown in Fig. A5.30. µ · Area µ0 · w · d µ · Area µC · w · d
Step 3 Calculate the reluctances from the geometry of the
motor and the magnetic permeability of the materials. Length l2
R magnet = =
We will approximate the cross-sectional area of the gap as µ · Area µ m · w · d
a plane. Note the cylindrical surfaces of the gap faces, which
are necessary to allow rotation, also increase the surface The flux path through the annular stator is assumed to be at
area. In reality, there is some magnetic flux which crossed the mean circumference.
the gap from the sides of the ferromagnetic core near the
gap, rather than the face of the gap, further increasing the ef- π O.D. + I .D. π (O.D. + I .D.)
fective cross-sectional area of the gap. This phenomenon is lstator = =
2 2 4
called “fringing”, Fig. A5.31. We will neglect it.
Length lc
Reluctance calculations: R stator = =
µ ⋅ Area µC · w · d

Fig. A5.32  a Rotor of a perma-


nent magnet stepper motor show-
a Non-Ferromagnetic b
Rotor Core or “Yoke”
ing circumferential magnetic
domains, b Permanent magnet N S Magnetized
S N
N S N
S
stepper motor showing rotor and N N Rotor Shell N
S
N
S S
stator with “claw” poles S

N N
S S
N N
S S
N S N
Polarity of the
Magnetic Domains
References and Suggested Reading 331

Fig. A5.33  a Rotor of a hybrid


stepper motor with axial mag- a b
netic polarization showing castel-
lation, b Hybrid stepper motor North
showing rotor and stator poles
without their coils

Permanent
Magnet

South

We can now express the flux through the motor as a function References and Suggested Reading
of motor current i.
Bird RB, Steward WE, Lightfoot EN (2006) Transport phenomena,
mmf 2nd edn. Wiley, New York
mmf = R φ → φ = Fitzgerald AE, Higginbotham DE, Grabel A (1981) Basic electrical
R
engineering. McGraw-Hill, New York
2 N (i + is ) Fitzgerald AE, Kingsley C, Umans SD (2003) Electric machinery,
φ= 6th edn. McGraw-Hill, New York
R stator Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
2R pole + 2R gap + R m + Cliffs
2
Rosenhow WM, Cho YI (1961) Heat, mass, and momentum transfer.
Prentice-Hall, Englewood Cliffs
φ= Rosenhow WM, Hartnett JP, Cho YI (1998) Handbook of heat transfer,
3rd edn. McGraw Hill
 1 l2 m 
2N  i + Rowell D, Wormley DN (1997) System dynamics: an introduction.
 2 N µ m w d  Prentice-Hall, Upper Saddle River
Sears FW (1951) Electricity and magnetism, 2nd edn. Addison-Wesley,
 π (O.D. + I .D.) 
Reading
 4  Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
 
 l1   g  l2  µC ⋅ w ⋅ d  dynamics. Addison-­Wesley, Reading
2 + 2 +
 µ ⋅ w ⋅ d  µ ⋅ w ⋅ d +
 µc ⋅ w ⋅ d  0 m 2
Power Transmission, Transformation,
and Conversion 6

Abstract 
A “simple machine,” such as a lever, transforms the magnitudes of the forces and veloci-
ties at its points of action but does not change the mechanical power, product of force and
velocity. Modern machinery both transforms power and converts power from one type to
another. Linear graphs of energetic models of machinery include interfaces which both
couple and separate the subsystems. The interfaces consist of two branches, one for each
subsystem. There is no direct action or flow of the through variable over the interface into
the other subsystem. Hybrid systems are systems which contain subsystems of different
types. Electric and hydraulic motors, generators, and pumps are hybrid systems. The linear
graph method couples similar and dissimilar subsystems using interfaces, which differ only
in the equations that describe their effect.

6.1 Introduction to Power Transmission, It is essential to keep in mind that the energetic models
Transformation, and Conversion we create represent only the dominant energetic attributes
of a real system. We must keep our models simple enough
Two important classes of machine components are those to be useful. All physically real machine components which
which (1) transmit and transform power and those which (2) transmit, transform, or convert power have the energetic at-
convert power from one form to another. Gear trains, chain tributes of energy storage and dissipation, in addition to the
and belt drives, and levers and linkages transmit and trans- attribute of transmission, transformation, or conversion of
form power. Electric motors and solenoids convert electrical power. If the physical component also stores or dissipates a
power to mechanical power. Similarly, hydraulic motors and significant amount of energy, then storage and dissipation el-
actuators convert fluid power to mechanical power. Mechani- ements must also be added to the model. Similarly, view the
cal power is converted into electrical and fluid power by elec- power transmission, transformation, or conversion attribute
trical generators and pumps. of a physical component as a distinct element of the model,
The terminology used in system dynamics for these independent of the component’s other attributes.
classes of machine components are “transformers” and
“transducers.” Transformers and transducers are interfaces
which transmit power between subsystems in dynamic sys- 6.1.1 Transformers
tem models. They are the dynamic system elements that per-
mit us to model useful systems, since most machines are a In common usage, the term “transformer” refers to an elec-
combination of interacting subsystems. A transformer is a trical transformer which changes the magnitudes of voltage
machine element which links or interfaces two subsystems and current in alternating current (AC) electrical power. Two
of the same type of energy. A transducer interfaces subsys- electrical subsystems are created, because electrical current
tems of dissimilar types. Note that the terms transformer on one side of a transformer does not flow into the circuit
and transducer have specific definitions in system dynamics on the other side. There is a physical coupling between the
which differ from, but are based on, their common engineer- two subsystems but no direct electrical connection. The trans-
ing usage. former is an interface between the two subsystems. The most

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_6, © Springer Science+Business Media New York 2014 333
334 6  Power Transmission, Transformation, and Conversion

common electrical transformer is a “step-down” transformer Lever Subsystem II


which decreases voltage and increases current, allowing AC
K
electrical power to be distributed at high voltage and low cur- 2 g
rent to reduce resistive energy dissipation but used at safer, Pivot
LA
lower voltages. We use electrical transformers as the basis of
our system dynamics terminology and define a “transformer” Power
to be the interface between subsystems of the same type of LB b
energy. Be aware that in most engineering communication F(t)
g
a “transformer” is an electrical transformer. It is only in a g
1
Systems Dynamics context that “transformer” has a broader
meaning. Subsystem I
As it happens, the physical coupling between the subsys-
Fig. 6.1   Levers and linkages interface translational mechanical sub-
tems in an electrical transformer is magnetic and requires a
systems
changing magnetic flux linkage to induce a voltage in the
secondary circuit. The time-varying magnetic flux linkage
in an electrical transformer is produced by a time-varying Power
current in the primary circuit. Consequently, electrical trans- Pinion
N1 teeth
formers transform the sinusoidal current of AC power but
do not transform direct current (DC) power. Other devices Fluid Clutch
Damping b1
such as DC to DC “converters” or a DC motor driving a DC Bearing
Damping b 2
generator must be used. We will defer our discussion of AC
power, until we have investigated the responses of systems Ω(t)
Inertia J
to sinusoidal inputs, known as their “frequency response.”
Angular velocity source. Gear
N2 teeth
Subsystem I Subsystem II
6.1.2  Ideal Transformers
Fig. 6.2   Gear sets interface rotational mechanical subsystems
An ideal transformer neither dissipates nor stores energy. All
of the energy which flows in must immediately flow out. Since
there is no storage or dissipation of energy, we can express • Double-ended pistons which interface fluid subsystems,
energy conservation as equal power flows into and out of a Fig. 6.4
transformer The subsystems linked by transformers are comprised of en-
ergetic elements which handle the same type of power. For
 Pin = Pout (6.1) example, the subsystems identified in the system with a lever
above are both translational mechanical subsystems Fig. 6.5.
Magnitudes are used in Eq. 6.1 to defer dealing with the sign Likewise, subsystems in a system with the gear set and the
convention for the time being. system with the belt drive are rotational. Transformers cou-
Naturally, real machine components which function as ple subsystems of the same energy “domain” or type (i.e.,
transformers are made of real materials and have less than translational, rotational, fluid, or electrical domain).
ideal bearings or electrical conductors. Consequently, the Although the subsystems are linked by a transformer, they
energetic model of machine elements which function as are also separated by it. A transformer is an interface between
transformers may need energy storage and dissipation ele- the subsystems which transmits power from one subsystem
ments, in addition to the element, which represents energy to another. In an ideal transformer, all of the power which
transformation. leaves one subsystem is transmitted to the other subsystem
Examples of mechanical transformers include the follow- without energy storage or loss. The power is transformed
ing: twice during the energy transfer, from its original form to
• Levers and linkages which interface translational mechan- an intermediate form and then back to its original form. In
ical subsystems, Fig. 6.1 the translational system with the lever, power is applied to
• Gear sets which interface rotational mechanical subsys- the lever at node one, transmitted to node two as rotational
tems, Fig. 6.2 power, and then applied to the element acting at node two as
• Belt drives which interface rotational mechanical subsys- translational power, Fig. 6.6.
tems, Fig. 6.3 It is this double conversion of power from the power type
of the subsystem to a different form and then back again
6.1  Introduction to Power Transmission, Transformation, and Conversion 335

Fig. 6.3   Belt drives interface Compliant Shaft Flywheel


rotational mechanical subsystems Spring Constant K Rotational Inertia J1
Pulley
Diameter D1
Pulleys Diameters Flywheel
D2 and D3 Rotational Inertia J2
Subsystem I
Subsystem II

Angular Velocity
Input Ω(t)

Power Subsystem III


Power Fluid Coupling
Damping b
Pulley
Diameter D4

Fluid Accumulator Subsystem II


Doubled-ended Capacitance C Power Lever Translation
massless piston ΤΩ = FBv2g
Area A1 Area A 2 K
2 g
Resistance R Pivot
LB
Intermediate
p(t)
Rotational
Pump
Subsystem
Subsystem II LA b
Subsystem I F(t)
g
g
Fluid 1
Reservoir Power ΤΩ = FAv1g
Power Subsystem I
Translation
Fig. 6.4   Double-ended pistons interface fluid subsystems
Fig. 6.6   Transformers rely on a second energy domain. A lever trans-
fers translational power via rotational power
+x,v

FA Actual transformers always dissipate some energy as heat


due to friction, electrical resistance, or magnetic hysteresis
1
and store some amount of energy. As always, the question we

L1
FB must address when deciding how to model a real transformer
is “How much energy?” where “a significant amount” or “a
2
L2 negligible amount” is relative to the magnitudes of the power
flows and energy storages elsewhere in the system. When
a single machine component possesses multiple energetic
properties, each energetic property is represented by a differ-
+Θ,Ω ent energetic element in the model. For example, since a real
lever has compliance and friction in addition to its transfor-
Fig. 6.5   Lever annotated to show the positive directions for transla-
tional and rotational velocity and displacement
mation property, the schematic Fig. 6.6 containing a spring
and a dashpot could be the energetic model of lever.
The linear graph symbol for a transformer, Fig. 6.7, has
which separates the two subsystems from each other. The two branches, one for each subsystem, separated by a gap so
double conversion of power also permits a transformer to that the branches connect to different nodes. A dashed ellipse
change the relative magnitudes of the power variables. The links the branches on either side to indicate the physical cou-
magnitude of the power flow, i.e., the product of the power pling which transmits power across the gap.
variables, is unchanged if the transformer is ideal.
336 6  Power Transmission, Transformation, and Conversion

LB
v1g = - ___ v2g PowerA Transformer Power B
LA or
1 2 Transducer

Fig. 6.8   A “block model” showing the positive power flows of a trans-
former or transducer. Note that both power flows are positive into the
FA FB transducer or transformer

• Pumps interface fluid and rotational or translational


g mechanical subsystems
Real transducers have dynamic properties, in addition to the
Fig. 6.7   The linear graph symbol for a transformer consists of two ability to transduce power. Similarly, a real motor or genera-
branches, one in each subsystem, linked by a dashed ellipse which de- tor stores kinetic energy in its rotational inertia, as well as
notes the power transfer across the interface
magnetic energy in the magnetic field created by the current
through its windings. It also dissipates energy as friction in
6.1.3 Transducers its bearings and electrical resistance its contacts and conduc-
tors. In a dynamic model, each significant energetic property
In mechanical engineering, the term “transducer” most com- must be represented by a different element. If the mass of
monly refers to a sensor, such as a load cell, which emits an a real lever is significant to the dynamic performance of a
electrical signal in response to a non-electrical input, such as system, then a mass element and a transformer element must
a force, in the case of a load cell. Sensors typically interface both be included in the lumped parameter dynamic model to
two dissimilar energy systems. In system dynamics termi- represent both the energy storage and transformation proper-
nology, transducers interface subsystems of different types ties of the lever. If the rotational inertia and friction is sig-
of energy. There is a significant difference between a sensor nificant in a real electric motor, then an inertia element, a
and a system dynamics transducer. Sensors produce signals, damping element, and a transducer element must be included
which are information, not power. An ideal sensor’s signal to represent these three independent energetic properties.
is a time-varying voltage or current, not power, which is the
product of current and voltage. (Real signals require a finite
amount of power to be transmitted and processed.) Likewise, 6.1.5  Block Model of Power Flows
the mechanical power needed to produce a force signal from
a load cell is extremely small, since a load cell actually sens- The transformation or transduction of power across an inter-
es elastic displacements, which, given the size of a load cell, face between subsystems can be represented schematically
are very small. Although the use of term “transducer” as the as a rectangle, or “block,” with the power flows on either
classification of machine elements which interface dissimi- side of the interface both shown in the positive direction,
lar types of energetic subsystems is logical, it is also unfortu- which is defined as into the interface, Fig. 6.8. We will refer
nate, because it causes misunderstanding between engineers. to this as the “bank account” sign convention for power flow.
It is best to use the term “transducer” only outside of system Power flow in is positive. Power flow out is negative.
dynamics to mean sensor to avoid misunderstanding. Because a transformer or transducer neither stores nor dis-
sipates energy, power which flows in one side must flow out
of the other. Consequently, although both flows are shown in
6.1.4  Ideal Transducers the positive direction as flowing into the element, this cannot
be. One power flow must be out. In the ideal case, power can
System dynamic transducers are similar to transformers in flow in either direction across the interface. Either PA or PB
that they interface the power flow between two subsystems. can flow in and the other out at a given instant. Showing both
They differ from transformers, because they interface dis- flows into the transformer or transducer underlines this fact.
similar subsystems. Examples of transducers include these: Unfortunately, it also requires a negative sign to reverse the
• DC motors interface electrical and rotational mechanical flow direction of one of the two power flows
subsystems
• Hydraulic pistons interface fluid and translational 
mechanical subsystems PA + PB = 0 (6.2)
• Racks and pinions interface translational and rotational
 PA = −PB (6.3)
mechanical subsystems
6.1  Introduction to Power Transmission, Transformation, and Conversion 337

Assuming both flows are into the interface is a useful, general Fig. 6.9   Linear graph symbol L
purpose assumption, particularly for systems in which power
of the lever as a transformer. v1g = - ___
B
v2g
Note that the arrowheads on both LA
can oscillate back and forth across the interface. The negative branches of the symbol orient the 1 2
sign introduced so that power flows sum to zero works nicely positive direction of the through
with gear sets, levers, belt drives, and other transformers. It variables, FA and FB towards
ground. Also note that the
is less than helpful for motors, pumps, and other transducers, transformer is identified by one
where signs are a problem, due to incompatible sign conven- of the two transformer equations, FA FB
tions in the two subsystems. Ultimately, we must return to which relate a power variable on
the linear graph and the schematic model of the system to in- one side of the interface to the
corresponding power variable on
terpret the meaning of a sign. We will not be above “fixing,” the other side
or, more politely, redefining, signs when necessary. A system g
equation which describes a system placed under closed-loop
feedback control cannot have an inversion between the input
and the output variables’ signs. A positive input must yield a
positive output, or the closed-loop system will be unstable. tution. Consequently, instead of a single equation with four
However, for the time being, we will accept whichever sign variables, we will derive two equations formulated in terms
the linear graph system yields. of a power variable from each side of the interface of the
form

6.1.6  Transformer and Transducer Equations v2 g = α v1g


(6.4)

and
Transformer and transducer equations are derived, beginning
with energy conservation expressed in terms of power flows. 1
 FB = − FA (6.5)
α
PA + PB = 0
where α is the “transformer” constant in this case, since the
This expression represents the power flow into the two ports variables on either side of the equation are from the same
of the interface, but it does not provide the equations we type of energetic system, translational mechanical. We will
need to derive a differential system equation. We have two include transformer and transducer equations with the ele-
power variables on each side of the transformer or transducer ment equations in our list of equations extracted from the
interface for a total of four variables. Say we were to work linear graph. Notice that they have a form similar to a dis-
with an electrical transformer where sipative element, i.e., power variable = constant × power
variable.
PA = FA v1g and PB = FB v2 g Derivation of the transformer constant can be done via
simple engineering analysis. On the other hand, test data are
Expressing energy conservation in terms of power yields usually needed to determine a transducer constant. We will
introduce the techniques required with the relevant transducer
FA v1g + FB v2 g = 0 or transformer.

This equation can be rearranged to express one variable in


terms of the other three, say 6.1.7 Transformer and Transducers Linear
Graph Symbol
 FA v1g 
FB = −   The linear graph symbol for transformer and transducers
 v2 g  consists of two branches, one for each side of the interface,
Fig. 6.9. The through variables on each side of the interface
Although it is a valid equation, is not a particularly useful must be named, such as FA and FB, so that they can be used
addition to our equation list, because we derive a system in equations. The transformer or transducer is identified by
equation by substituting to eliminate unwanted power vari- an equation which relates a power variable on the left side
ables. This equation would introduce three variables in place of the interface to a power variable on the right side, written
of one. above the transformer or transducer symbol.
We prefer equations which contain only two variables, so A dashed ellipse encircling the two branches of the in-
that we do not add variables to the equation upon substi- terface indicates that there is a power flow across interface.
338 6  Power Transmission, Transformation, and Conversion

The through variables do not act or flow across the inter- +x,v
face! There is no branch connecting nodes 1 and 2. This is
(or should be) obvious in a transducer, because the through FA
variables on either side of the interface are different. The 1
transducer element of a model of a DC motor separates the
electrical and rotational mechanical subsystems. The respec- L1
FB
tive through variables are current and torque. There cannot 2
be a branch connecting the nodes across the interface in a L2
transducer, because the continuity equation summing cur-
rents and torques would not make sense. They are different
physical quantities. Although it is less obvious in a trans-
former, because the subsystems on either side have the same Fig. 6.10   A lever is a translational mechanical transformer. Note that
units, a branch across the interface is just as meaningless. the forces acting on the lever at nodes 1 and 2 are both shown acting in
FA and FB are different forces, because they act at different the positive direction
points on the lever. They are coupled due to moment equi-
librium about the pivot, not a force balance which includes
A = FA v1g Lever B= FB v2g
a mystery force flowing between nodes 1 and 2. There is no
transformer
branch connecting nodes 1 and 2. interface
The power flowing in any branch of a linear graph is the
product of the through and across variables of that branch. Fig. 6.11   Block model of power flow through the transformer. Note
Referring to the lever’s linear graph transformer symbol that both power flows are positive into the transformer
above, PA = FA v1g flows down the left branch, and PB = FB v2 g
flows down the right branch. Each of these power flows is
equal in magnitude to the power flow across the transformer Within this assumption, the lever is a translational mechani-
interface. Energy conservation for the transformer, expressed cal transformer. Further, we will assume that the only prop-
in terms of power, because an ideal transformer neither stores erty of the device is transformation. The lever does not store
nor dissipates energy, is or dissipate energy. For this to be reasonably true, the lever
would need low mass (small kinetic energy storage), high
PA + PB = 0 → PA = −PB stiffness (small strain energy storage), and low friction at the
pivot (small energy dissipation). Define an arbitrary positive
This equation indicates that one of the two flows is oppo- direction, as shown. Assume that there are two translational
site the assumed positive direction, towards ground. For the velocity nodes on the lever, nodes 1 and 2, and show forces
example lever, one of the forces, FA and FB, must act in the FA and FB acting at nodes 1 and 2 in the assumed positive
negative direction and the other in the positive direction, be- direction.
cause both velocities, v1g and v2g, always have the same sign, There is a physical coupling between the power flow into
since they must be in the same direction. the lever at node 1 and the power flow into the lever at node
In this text, transformer and transducer branches will 2. However, there is no force which acts directly from node 1
often be connected directly to ground. However, there is no to node 2. Our linear graph symbol will have a gap between
need for them to connect to ground. It is perhaps more com- these two nodes to represent the interface.
mon in machine design that they do not. For example, many Energy conservation for the lever can be expressed in
motors are attached to compliant fittings which incorporate terms of power, since no energy is stored or dissipated by
some damping. The motor “mounts” in an automobile in- an ideal transformer. The sum of the power flowing into the
clude a compliant elastomer (a rubbery polymer) to reduce lever equals zero, Fig. 6.11.
the vibration transmitted to the passenger compartment. A
motor mount is modeled as a spring and dashpot in parallel. PA + PB = 0

The power flowing into the lever at node 1 and node 2 is


6.2 Transformers
PA = FA v1g and PB = FB v2 g
6.2.1  Mechanical Transformers Levers
since
We will assume that the rotation and resulting horizontal
displacement of the lever shown in Fig. 6.10 is small, so PA + PB = 0 → PA = −PB
we do not have to consider motion in the vertical direction.
6.2 Transformers 339

Then, All transformers and transducer equations have a fractional


and sign inversion of the transformer or transducer constant,
FA v1g = − FB v2 g when they are written with the power variable from one side
or branch of the interface on the same side in both equations.
If the lever can be modeled as rigid, then the translational ve- In the transformer equations above, the power variables for
locity of nodes 1 and 2, relative to ground, are related. The PB , v2g and FB, are on the left side of both equations.
translational velocities, v1g and v2g, are both functions of the Although it is recommended, we did not need to derive
angular velocity of the lever, Ω, and the distance of the node the transformer equation using the relationship between the
from the pivot velocities of nodes 1 and 2. If we had preferred, we could
have used a statement of moment equilibrium about the
v1g = L1Ω and v2 g = L2 Ω pivot. Defining counterclockwise as positive, with the forces
acting in the positive x-direction, as shown in Fig. 6.10
We eliminate the angular velocity, Ω to relate v1g and v2g
directly L 
FA L1 + FB L2 = 0 → FB = −  1  FA
 L2 
v1g v2 g L1
=Ω= → v1g = v2 g
L1 L2 L2 It is generally preferable to use geometric compatibility of
velocities, rather than equilibrium of forces or moments,
Having established a relationship between v1g and v2g from when possible, for two reasons. First, equilibrium equa-
the geometry of the lever, we can now substitute into the tions are more prone to sign errors of action versus reac-
equation of energy conservation, expressed in terms of tion. It is easier to see the direction of a velocity than a force
power to determine a relationship between FA and FB or moment. Second, it is often more difficult to formulate
the equilibrium statement than the geometric compatibility
L  statement. Our lever analysis is a good example. Note that
PA = −PB → FA v1g = − FB  2  v1g
 L1  summation of the horizontal forces shown is not a statement
of equilibrium
L  L 
FA v1g = − FB  2  v1g → FB = −  1  FA FA + FB ≠ 0
 L1   L2 
because we did not include the horizontal reaction force at
The two transformer equations for this lever are the pivot. That is why we wrote a moment equilibrium state-
ment about the pivot. It was obvious in this case that we were
 L  missing the reaction force, although there are many mecha-
v2 g =  2  v1g (6.6)
 L1  nisms where the relevant equilibrium condition or all of the
reaction forces and moments may not be as obvious.
 L1 
 FB = −   FA (6.7) 6.2.1.1 Mechanical Transformers Belt and Chain
 L2  Drives
Belt drives and chain drives are used to both distribute and/
Note that there is a fractional and sign inversion in the con- or alter rotational mechanical power. They allow multiple
stant term between these two equations. The fractional in- shafts at different locations and in different orientations to be
version is a result of energy conservation, PA = PB . The driven by the same motive source, Fig. 6.12. They also allow
sign inversion is a result of the linear graph method con- the magnitude of angular velocity and torque to be changed
vention to assume both PA and PB flow into the transformer, as needed, within the constraint that their product does not
PA + PB = 0 . The sign inversion between FA and FB tells us exceed the power available.
that one of the forces is acting in the negative direction.
It is easier to see the fractional and sign inversion of the 6.2.1.2  Belt Drives
transformer ratio, when the transformer ratio is renamed, Belt drives are used to drive the water pump, alternator,
L  power steering pump, and air conditioning compressor in
n≡  2 automobiles. Automobile engine fans used to be driven by
 L1 
belt (or directly from the crank (main) shaft. They are now
more commonly driven electrically, so that they can be
 1
v2 g = n v1g and FB = −   FA turned on and off as needed. An automobile’s air condition-
 n
340 6  Power Transmission, Transformation, and Conversion

Fig. 6.12   A nineteenth-century


belt drive with pulley shafts at vsurface
90°

Ω 1g Ω 2g

vsurface
Fig. 6.13   A belt drive schematic showing the nodes and sense of direc-
tion of rotational velocity and the surface speed of the belt

er compressor is turned on and off as needed but, because it 6.2.1.4 Belt and Chain Drive Transformer
is belt driven, an expensive, magnetically engaged clutch is Equations
required. A number of different style belts are used in ma- The analyses of a chain drive and belt drive are identical.
chine design, but there are only two basic types: (1) smooth Chains are engaged by sprockets and cannot slip without
and (2) “cogged” or “toothed” belts. The cross-section of a breaking. Although smooth belts can slip against a pulley,
smooth belt may be planar, “v” shaped, or ribbed. A belt is producing a “squeal,” they are not supposed to. Therefore,
“smooth” if the cross-section is constant in the longitudinal one presumes that the belt, like a chain, does not slip against
direction along the belt. Conversely, the cross-section of either driving or being driven by a pulley. If that is true, then
toothed belts varies periodically in the longitudinal direc- the surface speed of the pulley equals the surface speed of
tion. Rectangularly shaped, transverse projections (teeth) the belt, Fig. 6.13.
are spaced uniformly along the belt to engage pockets on We will assume that we have ideal massless belts and
thee pulleys, so that the belt cannot slip. Cogged or toothed chains which are perfectly flexible, but completely inexten-
belts are commonly called “timing belts,” since they are sible, meaning that they conform to the sprocket or pulley
used to synchronize the rotation of an automobile’s cam but do not stretch. If they are massless, then they cannot store
shaft(s) and crank shaft in place of chain drives or gear kinetic energy. If they cannot stretch, then they cannot store
sets. Belt drives have applications in computer printers and elastic energy. We will assume that their perfect flexibility
scanners. Toothed belts have replaced chain drives in many allows them to conform to the sprocket or pulley without
applications, including as the main drive of motorcycles. dissipating energy in deformation. Lastly, we will assume
They have also replaced wire rope to lift elevators. that the chain or belt slips on and off the sprocket or pulley
without friction. Although belts and chains are designed to
6.2.1.3  Chain Drives minimize energy storage and dissipation, none of these as-
Although chain drives can be more expensive and require sumptions is true. If a significant amount of energy is stored
more maintenance than belt drives, they can also carry heavi- or dissipated in any of these mechanisms, then elements
er loads and have longer service lives. A current application must be added to the system to account for that energy.
of a chain drive is as a component in the automatic trans- At any instant, one side of the belt between the pulleys
mission for a front wheel drive car. The automatic transmis- will be “slack” and the other side “taut.” The taut side of the
sion of a front wheel drive car must be shorter than one for belt conducts the tensile force from the driving pulley to the
a rear wheel drive car, because the engine and transmission driven pulley needed to transmit power. The slack side has
of a front wheel drive car is mounted “transverse,” facing enough tensile force to maintain static friction between the
the passenger’s side, rather than lengthwise, facing front. A pulleys and the belt, as well as to keep the belt from flapping.
chain drive is used to transfer the power from the “torque The slack side of the belt does not transmit power. Recall
converter” pump and turbine shafts to a second, parallel shaft the engineering adage, “You can’t push a rope.” Belt drives
with the clutch plates and gear sets, essentially splitting the transmit power by pulling against the pulleys with the net
transmission into two to shorten it. tensile force, FTaut − FSlack = FNet . The power transmitted is
PBelt = FNet vsurface.
6.2 Transformers 341

Expressing energy conservation in terms of power,


A= TA Ω 4g Belt Drive B= TB Ω 5g
Fig. 6.14, the power PA flowing into or out of the belt drive transformer
at node 1 is interface

PA = TA Ω 1g Fig. 6.14   Block model of positive power flows into the belt drive
transformer interface

and the power PB flowing into or out of the belt drive at node
r 
2 is TA Ω 4 g = −TB  A  Ω 4 g
 rB 
PB = TB Ω 2 g
 r  (6.12)
TA = −  A  TB
Although we could formulate the transformer equations for  rB 
a belt drive using moment equilibrium in terms of FNet, it is
less error prone to use geometric compatibility. The surface Note that we derived the transformer equations using the re-
speed of the pulleys can be expressed in terms of their angu- lationships between the geometry of the belt drive and the
lar velocities and radii velocities of the pulleys. We could have used the fact that
belt tension force is equal at the two pulleys to derive the
 vsurface = rA Ω 1g (6.8) transformer equations, starting with pulley torque expressed
in terms of the belt tension force. However, equilibrium anal-
and yses are prone to sign errors. The sign convention of the lin-
ear graph method requires the unknown torques acting on a
(6.9)
vsurface = rB Ω 2 g body to be positive (i.e., in the positive direction). Engineers
have the tendency to assign forces and torques, so as to place
where rA and rB are the radii of the pulleys. Equating the an object in equilibrium, which inadvertently reverses one of
surface speeds yields the two torques.

rA Ω 4 g = rB Ω 5 g
6.2.2 Gears Mechanical Transformers
and, hence, the relationship between the two angular veloci- and Transducers
ties
Gear sets are used to either distribute or alter rotational mechani-
r
(6.10)
Ω 2 g = A Ω 1g cal power. They allow multiple shafts at different locations and
rB in different orientations to be driven by the same motive source.
They also allow the relationship between angular velocity and
In practice, pulleys are identified by their diameter, not by torque to be changed as needed. Typically, gear sets “step down,”
their radius. The relationship between the angular velocities or reduce angular velocity, and “step up,” or increase torque.
can be expressed in terms of diameters, if convenient “Overdrive” refers to the event, when a gear set increases angu-
lar velocity above the velocity of the input shaft. For example, an
 DA automobile may use overdrive at highway speed, when the trans-
Ω 2g = Ω 1g (6.11)
mission steps up the angular velocity of the drive (or propeller)
DB
shaft, so that it exceeds the angular velocity of the crank shaft.
The relationship between the torques acting on the two pul- Gears are classified by the shape of the blank, the loca-
leys can now be calculated using energy conservation ex- tion of the teeth on the blank, the cross-sectional shape of the
pressed as power, since the belt drive’s transform attribute teeth, the line or curve a tooth makes on the gear blank, and
only transforms the power variables. If energy is stored in the orientation of the gears’ shafts.
the belt drive, then kinetic or strain energy storage elements The overall shape of a gear is generally either a cylinder,
must be added to the model. Remember the linear graph sign the frustum of a right cone (a truncated cone), or a rectan-
conventions require that the power flowing into and out of gular prism (or bar), Fig. 6.15. More exotic shapes are used
the belt drive sum to zero. Equating the power flows leads to in some highly engineered applications, such as automotive
a sign error, Fig. 6.14 transmissions and differentials.
Gears can be combined in a wide variety of ways, since
the only fundamental requirement is that the teeth mesh as
PA + PB = 0 → PA = −PB → TA Ω 4 g = −TB Ω 5 g the gears either rotate or translate.
342 6  Power Transmission, Transformation, and Conversion

Fig. 6.15   Gear blank shapes are cylinders, truncated right cones and
bars

Fig. 6.18   A rack and pinion ( left) and a worm and gear ( right). The
axes of motion are not coplanar

Ω1 Ω2

Fig. 6.16   External ( left) and internal ( right) gears

Pinion
N1 Teeth Gear
N2 Teeth

Fig. 6.19   A pair of spur gears. “Pinion” means smaller gear, generally
on the input shaft

against each other, rather than sliding, thus reducing the fric-
Fig. 6.17   Bevel ( left) and miter ( right) gears. The axes of rotation of tional losses and tooth wear. Consequently, excessive wear
miter gears intersect at right angles. If the angle of intersection is not on gear teeth indicates that the gears are either misaligned,
90°, then the gears are bevel gears
with their shafts incorrectly spaced or not parallel, or the
gears have different “pressure angles,” defined below, and
The left-hand gear set in Fig. 6.16 is a set of external spur incompatible shapes.
gears. The larger gear of the right-hand set is a “ring” gear. A positive direction for the angular velocities and shaft
Ring gears can have teeth on their cylindrical or planar sur- torques must be defined vectorially using the right-hand rule.
faces. The pinion rotates at angular velocity Ω1g and the gear ro-
Bevel and miter blanks are frustums of right cones, tates at Ω2g. Torque TA acts on the pinion shaft and torque TB
Fig. 6.17. Varying the included angle at the tip of the cone on the gear shaft, both in the positive direction.
allows the gears’ shaft to intersect at various angles. Miter We can see from the schematic of the two gears, Fig. 6.19,
gears’ shafts intersect at 90o. that they must rotate in opposite directions (counter-rotate).
A rack and pinion, Fig. 6.18, is a transducer between ro- Both gears must have the same surface velocity of the pitch
tational and translational motion, making it a transducer in surface, so that the teeth mesh. The smaller pinion must ro-
system dynamics terminology. A worm and gear has a high tate faster than the larger gear to feed teeth into the nip at the
“gear ratio” and is “self-locking,” meaning the gear cannot same rate. For gears to mesh, their teeth must be the same
drive the worm. The worm must drive the gear. size. Consequently, a tooth and the adjacent space can be
used as a convenient unit of measure, when comparing the
6.2.2.1  Spur Gears circumferences of two meshing gears. The formal measure-
Two intermeshing spur gears are shown in cross-section in ments require identifying where on the surface of the tooth
Fig.  6.19. The name “spur” refers to the similarity of the the gear force acts. A circle passing through this location on
cross-section to a horse rider’s spur. A spur gear has a cy- each tooth and centered on the axis of the gear is called the
lindrically shaped blank. Its teeth have the “involute” shape, pitch circle, Fig. 6.20. Its radius and diameter are the pitch
which has the benefit of the two mating tooth surface rolling radius and pitch diameter, respectively. The circular pitch is
6.2 Transformers 343

Pitch Circle

Circular Pitch

o o
N Teeth 0 Helix 30 Helix

Pitch Diameter Fig. 6.22   External spur gears of two 0° and 30° helix angles

Fig. 6.20   The pitch circle, pitch diameter, and circumferential pitch
of a spur gear A= TA Ω 1g Gear Set B = TB Ω 2g
transducer
interface
Addendum
Dedendum Fig. 6.23   Block diagram of a gear set’s positive power flows

Fig. 6.24   Linear graph symbol N


Pitch Circle of gear set Ω 1g = - ___
2
Ω 2g
Tooth Force FT N1
FT 1 2

FT

Pressure Angle TA TB
14-1/2 o, 20o, 25o

Fig. 6.21   Tooth forces, pitch circles, and the addendum and dedendum g

the circumferential pitch. The circular pitch is the distance ing gear teeth appear as mirror images of each other, when
from the surface of one tooth to the next, along the pitch looking at the “nip,” where they mesh. Helical gears carry
circle. heavy torques, because more than one tooth is meshed. Heli-
The equilibrium condition for gears can be somewhat cal gears also place a component of the tooth force parallel to
counterintuitive. Levers have moment equilibrium about the the gears shaft, which must be carried by the thrust compo-
pivot. One can think of the teeth of meshed spur gears as a nent of the shaft’s bearings.
series of levers making and breaking contact with one an- The relationships between the input and output torques
other. The contact force between the meshed gear teeth (the and angular velocities are calculated using energy conserva-
tooth force) must be equal and opposite (action and reac- tion but expressed in terms of power flow, since the trans-
tion). The torques on the respective shafts are not equal and former attribute of the gear set neither stores nor dissipates
opposite, unless the gears have the same pitch radius, where energy. If the actual gear set stores or dissipates a significant
the pitch radius is the distance from the axis of a gear to amount of energy, where “significant” is always relative to
where the tooth forces act. the other elements in the dynamic system, then elements
The geometry of gears is described by the nineteenth- must be added to the model of the gear, to account for these
century terminology, Fig. 6.21. The “addendum” and “de- other power flows.
dendum” are the length of the gear tooth above and below Proceeding as with the lever, power PA flowing into or
the pitch circle. The “pressure angle” is the inclination of the out of the pinion is PA = TA Ω 1g and the power PA flowing
tooth force relative to the normal to the line between the two into or out of the gear is PB = TB Ω 2 g , Figs. 6.23 and 6.24.
gears’ centers of rotation. The sum of the power flows into the transformer interface
The tops of spur gear teeth can be either parallel to is zero:
the axis of the gear or trace a helix about the gear’s axis,
Fig. 6.22. The latter are known as helical spur gears. Mesh- PA + PB = 0 → PA = −PB → TA Ω 1g = −TB Ω 2 g
344 6  Power Transmission, Transformation, and Conversion

Fig. 6.25   60° Bevel gears

60 o
Shaft Angle

Pitch radius is an inconvenient and difficult dimension to power for the gear set to derive the relationship between TA
measure. It is far easier to count the number of teeth on a and TB
gear, Nteeth, as a measure of its size of relative gears which
will mesh with it. The relative speeds of two gears’ pitch  N 
surfaces must be equal for the teeth to mesh. The smaller TA Ω1g = −TB Ω 2 g → TA  − 2  Ω 2 g = −TB Ω 2 g
 N1 
diameter gear must rotate faster to feed teeth into the nip at
the same rate as the larger gear N
(6.15)
TA = 1 TB
N2
 v pitch = rp Ω1g = − rG Ω 2 g (6.13)
surface

6.2.2.2  Bevel Gears


Note the negative sign. Two meshing external gears must The pitch surface of bevel gears is the frustum of a right
counter-rotate (i.e., rotate in opposite directions). circular cone, Figs. 6.25 and 6.26. They are used to change
The geometric relationship between radius, diameter, and the orientation of the shafts. If the shaft angle is 90o, then
circumference are the bevel gears are called “miter” gears.

diameter 6.2.2.3  Worm and Gear


radius = and circumference = π diameter
2 Worm drives provide a large increase in torque and reduc-
tion in angular velocity, Fig. 6.27. Rotation of the worm can
circumference drive the mating gear in either direction. However, geometry
radius =
2π prevents rotation of the gear from driving the worm. A worm
drive is known as a “self-locking” gear set. As a result, a
Using dtooth as the circumferential distance along the gear worm drive is an extremely non-linear system, since power
from a point on a tooth to the same point on the adjacent can only flow in one direction, from the worm to the gear.
tooth, the circumference and radius of a gear is A good example of the importance of an understanding of
system dynamics in machine design is to consider the con-
dtooth N teeth sequence of using a worm and gear to drive a large inertial
circumference = dtooth N teeth → radius =
2π load. What could happen if the motor driving the worm were
to suddenly fail?
When two gears mesh, we can use the number of teeth on
each gear as a relative measure of its pitch radius 6.2.2.4  Gear Trains or Sets
Gear “trains” and gear “sets” are synonyms. Gear trains
 N2 transmit power between the input and output shafts with
Ω 1g = − Ω 2g (6.14)
N1 some frictional loss in each nip (or mesh). Our analysis of
a pair of spur gears can be extended to any number of gears
Once the angular velocities of the two gears are related by arranged in a set. A particular gear set may be intended either
their respective teeth numbers, we substitute back into the to (1) step up or step down (i.e., transform), the angular ve-
equation for energy conservation, expressed in terms of locity (including direction) or the torque between the input
6.2 Transformers 345

Fig. 6.26   120° Bevel gears

120o
Shaft Angle

Fig. 6.27   Worm and gear

shaft and the output shaft of the gear set or (2) change the Ω1g
location or orientation of the output shaft relative to the input
shaft.
Determining the direction of rotation of the output shaft Ω 4g
is an elementary but error-prone step. Mechanical engineers N1
like to think that they can look at a gear set of any complex- N2
ity, and see the relationship between the input and output
shafts’ rotations. This engineering hubris often yields need- N3
less, embarrassing errors. A reliable method of determining
the relative rotation of gears is to indicate the surface veloc- N4
ity at either side of every mesh, Fig. 6.28.
The rotation reverses with each mesh from input Ω 1g to Fig. 6.28   The surface velocities are indicated on both sides of each
output Ω 4 g . The relationship between angular velocities Ω 1g mesh to determine the directions of rotation
and Ω 4 g is
Notice that the presence of gears two and three does not
 N1   N 2   N 3  change the magnitude of the output angular velocity, but
Ω1g  −   − N   − N  = Ω 4 g their meshes do contribute to its direction. Since each mesh
 N 2  3 4
reverses direction, an even count of meshes yields the di-
 N1   N 2   N 3  rection of rotation for the input and output shafts. We can
Ω1g  −  −  −  = Ω4 g confirm this by removing gear three from the gear train,
 N 2   N3   N 4 
Fig. 6.29.
N1
− Ω1g = Ω 4 g
N4  N1   N 2   N   N2 
Ω1g  − − = Ω4 g → Ω1g  − 1   −  = Ω4 g
 N 2   N 4   N2   N4 
346 6  Power Transmission, Transformation, and Conversion

Ω1g
N1
Ω 4g N3
N1
N2
TA Ω 1g
N2
N4 N4

Fig. 6.29   Removing gear three of the gear train of Fig. 6.28 reverses
the direction of gear four TB Ω 4g

Fig. 6.31   Gears fixed to a common shaft are known as compound or


cluster gears. Cluster or compound gears affect the gear ratio of a gear
set

Flow In is Positive
PowerA Transformer Power B
or
Transducer
Driving Idler
Flow Out is Negative
Pinion
Driven Gear Fig. 6.32   The power flows on both sides of the transformer or trans-
ducer are assumed to flow in the positive direction, into the transformer
Fig. 6.30   Idler gears reverse rotation and change the location of the or transducer. One flow must actually be in the opposite direction
output shaft but do not change the gear ratio of a gear set

surface speed varies with radius or number of teeth. Since


N1 meshing gears have the same surface speed, this results in
Ω1g = Ω4 g
N4 the gear ratio of each mesh contributing to the input–output
gear ratio, Fig. 6.31.
Gears two and three in Fig. 6.30 are called “idler” gears.
They serve to transmit power and change the location of  N1   N 3  N1 N 3
the output shaft but have no effect on the magnitudes of the  − N   − N  Ω1g = Ω 4 g → N2 N4
Ω1g = Ω4 g
2 4
power variables. Idler gears do not affect the gear ratio of
a gear set. Again, idler gears do affect the direction of the
output shaft’s rotation.
6.3 Transformer and Transducers Sign
6.2.2.5  Transforming Angular Velocity or Torque Conventions
As derived above, idler gears do not affect the magnitude
of the output torque or angular velocity, just their sign and The linear graph method uses a sign convention for transduc-
the location of the output shaft. The largest gear ratio which ers and transformers, which presumes that power may flow
can be achieved by a pair of gears is limited by geometric in either direction across the interface, Fig. 6.32. The sign
compatibility and the space available. For a given tooth size, convention has two parts. First, the assumed positive direc-
the pinion has a minimum circumference that will allow its tion of the through variable in a branch of the transformer
teeth to mesh with the gear. The maximum allowable diam- or transducer is oriented in the direction of the drop of the
eter of a gear is limited by the space available in the machine across variable of that branch, i.e., towards ground, Fig. 6.33.
and the machine’s budget. If the gear ratio needed cannot be
achieved with two gears, the alternative is to affix interme- PA = TA Ω 2 g and PB = TB Ω3 g
diate pairs of gears to common shafts. This arrangement is
known as a “cluster” gear or a “compound” gear. Each gear
PA + PB = 0 → PA = −PB
on a common shaft has the same angular velocity, but its
6.3  Transformer and Transducers Sign Conventions 347

N contrived situations. Consequently, in fluid mechanics, posi-


Ω 2g = - ___
3
Ω 3g tive fluid pressure is compressive. Negative pressure (ten-
N2
2 3 sion) yields fluid cavitation, which develops gas bubbles or
vacuum within the fluid. In mechanical design, bubbles in a
fluid are generally bad news. Air entrained in hydraulic fluid
PowerA Power B reduces its bulk modulus to an unknown value. The bubbles
TA TB themselves may impede the function of a machine, such as
when froth or foam limits the flow into a pump. Further, high
stresses can act on the surfaces of machine parts, when a
bubble collapses, as the flow enters a region of positive pres-
g sure. Cavitation can erode hardened steel.
When the positive direction is arbitrary, in the sense that
Fig. 6.33   The linear graph sign convention orients the power flows the variable is equally likely to have either state, engineers
towards ground down both of branches of a transformer or a transducer.
One power flow must actually be in the opposite direction. The dashed
tend to define the state we prefer as positive. For example,
ellipse indicates transmission of power across the interface in solid (continuum) mechanics, we define the normal force
acting on the face of an infinitesimal cube to be positive,
when it acts outwards. The sign convention corresponds to
TA Ω 2 g + TB Ω3 g = 0 → TA Ω 2 g = −TB Ω3 g the positive direction on one face, out of two normal to an
axis. Note that the sign convention places the element in ten-
The linear graph sign convention for transformers and trans- sion. This jibes with the preference of mechanical designers,
ducers assumes that the power flows on both sides are posi- although that may be a coincidence. Engineers prefer posi-
tive, where positive power flow is defined as flow into the tive stress. Given two mechanical components of the same
interface. This is impossible, because an ideal transducer or length, the component loaded in tension would be smaller
transform is an interface between subsystems which cannot in cross-section, requiring less material, than the component
store energy. Hence, the second part of the sign convention loaded in compression. Further, tensile rupture is typically
is that the two power flows must sum to zero. Physically, at preceded by some amount of plastic deformation, which can
any instant, power flows into the interface on one side yields signal the overload and forewarn of an impending failure.
power flow out of the interface on the other side. Reversal of Compressive failure is often catastrophic buckling, which
one of the two power flows inverts the sign of one of the four provides no early warning.
power variables. Unfortunately, this sign inversion often Unfortunately, solid mechanics and vector mechanics do
leads to nonsensical mismatches between the signs of differ- not have the same sign convention. Positive external forces
ent subsystems. and internal forces are defined differently. A positive exter-
Signs are a problem in many mechanical engineering nal force is a force which accelerates a mass in the positive
calculations. Sign conventions used in the various subdis- direction. A positive internal force yields tensile stress. What
ciplines within mechanical engineering arose “organically” is the state of stress of an element subjected to acceleration?
to suit the specific subdiscipline. The state of a variable that If the acceleration is large enough, so that the acceleration
was defined as positive was either (1) the same as a different of gravity is negligible, the element is compressed due to in-
vector variable in the same system, (2) the more common ertia, regardless of the direction of the acceleration. A given
state, or (3) the preferred state. An example of the first case force may produce positive acceleration (in the positive di-
is the definition in vector mechanics, that a positive force rection) and negative (compressive) stress in the same ma-
accelerates a mass in the positive direction. It is possible to chine element.
invert the signs between displacement and force vectors, but As one would expect, signs become arbitrary, when we in-
we would rather not. First, why bother? Further, sign inver- terface dissimilar subsystems. For example, what is the posi-
sions lead to errors. Negative signs are dropped or assigned tive direction of rotation for the shaft of an electric motor?
to the wrong variable. Therefore, we define a force accel- There is some consensus that positive rotation is clockwise,
erating a mass as positive, if the acceleration occurs in the but there is no consensus where the observer stands, in front
positive direction. of the motor or behind the motor. The only way to determine
Examples of the positive direction as the more common the positive direction of a motor is to apply a positive voltage
state arise, due to the nature of the materials in the subdis- (if it is a DC motor) and observe its rotation. The positive
cipline. Particulates such as soil cannot carry tension. They direction of a motor depends on its manufacturer.
must be loaded in compression. Therefore, in soil mechan- Our solution to the sign problems inherent in system dy-
ics, compressive force and compressive stress are positive. namics will be to accept the signs of the linear graph method,
Likewise, fluids cannot carry tension, except in carefully even when some are nonsensical. We will follow the linear
348 6  Power Transmission, Transformation, and Conversion

Pinion in this system are not assigned energetic properties. This


N1 teeth Inertia J
implies that the shafts are very stiff, have negligible inertia,
Fluid Coupling and, consequently, can be modeled as ideal shafts having no
Damping b energetic properties. Ideal shafts have a single angular veloc-
ity, because they are perfectly rigid.
The first step in drawing the linear graph is to find the
nodes of distinct values for the across variable angular veloc-
T(t)
ity, Ω. The torque source is shown as a vector. A real torque
source must react against something. Since no reaction
Torque source T(t). is shown, we assume that the torque source reacts against
Positive rotation direction Gear
defined by the right-hand rule N 2 teeth ground. To transmit torque through the fluid coupling, the
drag and the cup (inner and outer elements) of the fluid
Fig. 6.34   Rotational mechanical system. The torque source defines the coupling must shear the fluid between them. Therefore, the
positive direction for the system drag and cup and the shafts attached to them have differ-
ent angular velocities. The output shaft of the fluid coupling
drives the pinion. The pinion and gear must counter-rotate.
graph method’s sign convention during the reduction of a Consequently, the pinion and gear will have different angular
system to a system equation. We will then define the positive velocities, even if they have the same number of teeth. The
state for the output variable on a case-by-case basis, when inertia is fixed to the ideal shaft driven by the gear and has
we interpret our results. Attempting to correct signs before the same velocity as the gear.
one has completely derived the system equation leads to er- We have identified three distinct angular velocities plus
rors and confusion. It is better to have nonsensical but con- ground. The transformer relationship we derived above will
sistent signs than uncertainty over what a sign means. be used here, revising the subscripts for the angular veloci-
A common use of the linear graph method is to derive ties to be Ω2g for the pinion and Ω3g for the gear
the system equation (mathematical model) of a system to be
placed under closed-loop feedback control. In any feedback N2 N
Ω2g = Ω 3 g and TA = 1 TB
control application, we must define signs such that a posi- N1 N2
tive input yields a positive output. Otherwise, the resulting
closed-loop control system will be unstable. Both of these equations are added to our equation list with
In summary, the linear graph method sign convention for the other element equations. One of these equations must be
mechanical transformers and transducers requires us to written above the transformer symbol on the linear graph to
1. Assume both power flows are positive, which is defined identify the transformer. One equation is sufficient, because
as into the transformer or transducer. the second transformer equation can be derived from it and
2. The sum of power flows into the transformer or trans- the linear graph. The two branches which form the interface
ducer must equal zero, because it cannot store energy, represent the power flows into the transformer:
PA + PB = 0. This inverts one of the power flow signs,
PA = −PB . PA + PB = 0 → TA Ω 2 g + TB Ω3 g = 0
3. Show the forces or torques acting on the transducer or
transformer as a free body diagram in the positive direc-
N  N
tion. TA Ω 2 g − TB  1  Ω 2 g = 0 → TA = 1 TB
4. Orient the through variable flow on both branches of the  N2  N2
linear graph symbol towards ground.
We need to define the positive direction for torques and an-
gular velocities. If a positive direction is not indicated ex-
6.3.1 Example 1: The Linear Graph plicitly, then use the direction of the source. Because the
for a Rotational System gear set is an interface between two subsystems, it has two
with a Transformer. branches in its linear graph symbol. The torques shown on
The linear graph sign conventions are illustrated, using the the branches of the linear graph symbol for the transduc-
rotational mechanism shown in Fig. 6.34, which consists of er are the torques acting on the gears. The positive direc-
a torque source, T( t), a fluid coupling or drag cup with ro- tions of these torques is towards ground. The linear graph is
tational damping b, a spur gear set, and a mass moment of Fig. 6.35.
inertia or rotational inertia J. The positive direction is de- Now, we shall check that our sign conventions for the
fined as the direction of the input torque. Note that the shafts transformer lead to signs that make sense. Torques TA and TB
6.3  Transformer and Transducers Sign Conventions 349

N
Ω 2g = - ___
2 Mass-moment
Ω 3g of Inertia J
b N1
1 2 3

TJ
T(t) TP TG J TB
3

Fig. 6.37   Torque Tb acting on the gear shaft and torque TJ acting to
g g g accelerate the mass moment of inertia J

Fig. 6.35   Linear graph of the rotational system shown in Fig. 6.34.


The positive direction of each branch is defined by its arrowhead. Both
branches of the transformer interface are oriented towards ground TA
Tb
Fluid Coupling
Damping b
2
TJ
TB Tb
3
T
TA 1
Tb
2 Fig. 6.38   Action and reaction torques shown on the fluid coupling

Fig. 6.36   Positive torques TA and TB shown on the shafts of the gear There are two aspects of this drawing which may look
set. Reaction torques Tb and TJ shown on nodes 2 and 3 wrong. First, why is the angular velocity of node 2 in the
opposite direction of torque Tb? In order for torque Tb to
act through the fluid coupling, the fluid coupling must be
are shown on the linear graph as positive torques, where the in quasi-static equilibrium. (The prefix “quasi” means “as
positive direction for a through variable in the linear graph if” or “in a sense.” In this context, it means we can formu-
is towards ground. Torques TA and TB are shown in the free late torque equilibrium, even though the fluid coupling is in
body diagram acting on shafts of the gear set, Fig. 6.36. The motion.) The torques acting on the fluid coupling must sum
positive direction for this free body diagram was defined as to zero. The elemental equation for the fluid coupling with
the direction of the input torque. damping b
The reaction torque from the fluid coupling against the
pinion’s shaft, Tb, at node 2 and the reaction torque from Tb = bΩ12 = b ( Ω1 − Ω 2 )
the rotational inertia against the gear, TJ, at node 3 are also
shown. Torques Tb and TJ shown are the torques acting on states there must be a difference in angular velocity between
the fluid coupling and the rotational inertia, respectively. the two ends shafts of the coupling, for a torque to exist. The
These are the torques, whose signs must make sense physi- two components of fluid coupling need not move in opposite
cally. Starting with the inertia, we see it will rotate in the directions. In fact, in this system, they will rotate in the same
same direction as the gear, counterclockwise for the given direction, expect during the transient period after the input
input torque. Using the right-hand rule on the vector TJ, we torque changes direction.
see that it produces the correct counterclockwise rotation of Second, if we have assumed that the applied torque, T,
the inertia, Fig. 6.37. acts away from ground, why is torque T at node 1 acting
Considering the fluid coupling, Fig. 6.38, we must have towards ground? This is also a case of action and reaction.
equal and opposite torques acting on its two shafts. Torque TA If we were to draw a free body diagram of the device serv-
at node 2 is the torque acting on the pinion. It is in the posi- ing as our torque source, Fig. 6.39, we would see that, for
tive direction. The reaction torque is the torque Tb, acting in the torque source to be in quasi-static equilibrium, and for
the negative direction. Similarly, at node 1, torque Tb is the torque T to act away from ground at the ground (node g), the
torque acting on the fluid coupling. The reaction torque T is torque T acting at the other end of the torque source at node 1
the torque of the fluid coupling on the torque source. must be in the opposite direction.
350 6  Power Transmission, Transformation, and Conversion

Tb Diametrically opposed graphite


brushes pressed against the Cross-section of
T(t) commutator by soft springs rolling-contact
bearing
1 energize the rotor windings -i
Torque Source
Rotor, winds, shaft
Rotational inertia J
T(t)
Tg
g Commutator,

Fig. 6.39   Action and reaction torques shown on the torque source
φ +i
Rotating electrical
contacts
Magnetic flux φ emerging
from the rotor’s north pole Energizing
current i
In summary, the linear graph sign conventions allow us
Fig. 6.40   DC motor rotor showing the commutator, brushes, and one
to yield the correct signs for the output variables of this sys-
set of windings
tem. This is true in the general case for a linear graph which
contains transformers but not transducers. Transformers in-
terface subsystems of the same type. If we respect the lin- 6.4.1 DC Electric Motors: Electrical
ear graph sign conventions, the signs of the power variables to Mechanical Transducers
are consistent. We avoid the necessity of drawing free body
diagrams for every mechanical element in a system, saving Motor action results from the interaction of a motor’s mag-
effort and the possibility of an action versus reaction error. netic fields. Large numbers of different designs lead to no
We will not enjoy the same success, when we work with standard terminology. In general, motors consist of a “stator”
transducers. Incompatibility of the signs between systems of which is fixed to the motor’s frame, and, in the case of rota-
different energy types requires further interpretation. tory motors, a rotor supported on bearings which revolves
within the frame. The magnetic field of the stator may be
created by a permanent magnet or by “field” windings. The
6.4 Transducers magnetic field of the rotor is created by a set of windings
energized progressively as the rotor spins.
In systems dynamics, “transduction” is the conversion of
power between subsystems with different energy domains 6.4.1.1  Permanent Magnet DC Motor
or types. It is a reasonable term, but its meaning in system DC motors interface electrical and rotational mechanical
dynamics does not correspond to its general engineering subsystems. A permanent magnet DC motor rotates, because
usage, where a transducer is a sensor. Consequently, outside the magnetic field, created by passing a current through a
systems dynamics, it is best to use the name of the type of set of windings on its rotor, is not aligned with the magnetic
transducer, such as DC motor, or gear pump, rather than the field of the permanent magnets adhered to the stationary
classification, transducer, to avoid confusion. housing. The moment created by the misalignment rotates
As an aside, there is weak analogy between the two uses of the rotor to align the two magnetic fields. Before the fields
transducer which is the root of its system dynamics usage. A are aligned, the stationary contacts, the brushes, and the pair
pressure transducer, for example, senses a fluid variable, and of moving contacts on the shaft connected to the winding
produces a signal which is an electrical variable. Sensors in- break contact, de-energizing the winding. The stationary
terface between dissimilar systems. However, in the ideal, the brushes make contact with the next pair of contacts on the
signals produced by sensors are not power flows but informa- shaft, as they move under them, energizing the next wind-
tion flows. Ideally, the pressure sensor’s output voltage could ing, which too is out of alignment with the stationary mag-
be read by the receiving instrument without a current flow into netic field, repeating the process. The ring of contacts on the
the receiver. In reality, that never happens. All receivers draw shaft is called the commutator, since it moves with the shaft.
some power from the signal. Consequently, all signals contain The brushes are actually blocks of graphite, pressed against
some power. Nevertheless, we model signals as being devoid the commutator by a spring to maintain contact. A drawing
of power, except when we are designing the “fan-out,” the of the rotor of a DC motor and schematic for a DC motor
number of receivers we can send a signal to, and have to deal model, neglecting the induction of the windings, are shown
with the realities of our electronics. in Figs. 6.40 and 6.41.
6.4 Transducers 351

R A= iMv2g DC Motor B= TMΩ 3g


transducer
interface

v(t) MOTOR JM Fig. 6.42   Block diagram showing the positive power flows into the
DC motor transducer element, which interfaces the electrical and me-
bM chanical subsystems

Fig. 6.43   Linear graph N


Fig. 6.41   DC motor energetic schematic. The motor’s induction is not symbol of the DC motor Ω 1g = - ___
2
Ω 2g
included in this model, because it occurs on a much smaller time scale transducer interface N1
than the mechanical response of the motor 1 2

Other than the voltage source, all elements shown in this


schematic are energetic properties of the DC motor. The cir-
cle labeled “motor” represents only the transducer property TA TB
of the motor. The conversion of electrical power to mechani-
cal power is known as “motor action.” There are other en-
ergetic properties in a DC motor, which are included in the
g
model as additional individual energetic elements.
The copper wire of the motor windings has electrical
resistance. There is also electrical resistance in the graphic
brushes. Finally, there is electrical resistance, due to the dis- tate the de-energized motor through a complete rotation. All
continuity between the graphite brushes and the copper com- of these energy losses are lumped into a single ideal viscous
mutator, known as contact resistance. These three electrical friction element, and modeled as rotational damping b.
resistances are in series. If the motor is poorly designed or The transducer property of the DC motor (its motor ac-
damped, arcing, an additional electrical resistance, can occur, tion), indicated in the schematic as a circle labeled “motor,”
when electrical contact is broken between a pair of contacts is the interface between the electrical and mechanical sub-
on the commutator and the brushes. Lastly, there is magnetic systems. It is important to emphasize that the ideal transduc-
hysteresis, which results in some of the energy stored in the er property neither stores nor dissipates energy. Any energy
magnetic field not returning to the winding as electric cur- storage or dissipation in a real transducer must be represent-
rent, but being lost as heat, when the coil is de-energized ed by elements in addition to the transducer element. The
and the magnetic field collapses. All of these energy losses transduction can be represented using the transformer block,
are summed and represented as a lumped property, an ideal with both power flows flowing in, as shown in Fig. 6.42, al-
linear electrical resistance R. though one flow must be negative. In the linear graph sym-
The rotating element in the motor is made of a stack of bol, both through variables, current iM and torque TM, flow
stampings of “electrical steel” (alloyed with silicon for high to ground, Fig. 6.43. Both across variables, voltage v2g and
magnetic permeability, to enhance the magnetic field created angular velocity Ω3g, are referenced to the ground.
by the current through the windings) supported by a steel The product of current iM and voltage between node 2 and
shaft. The surfaces of the steel stampings are insulated to ground on the electrical side of the interface is the electrical
limit “eddy currents” created by the voltage, induced in the power, which flows into the transducer
steel by the movement of the rotor in the field of the perma-
nent magnets. The windings are many turns of thin copper  PElectrical = iM v2 g (6.16)
wire. Copper is denser than steel, and the windings contrib-
ute substantially to the mass moment of inertia of the rotor, Likewise, the product of torque TM and angular velocity be-
J. In addition to kinetic energy storage in rotational inertia tween node 3 and ground on the electrical side of the inter-
J, there is frictional energy loss in the bushings, or bearings, face is the mechanical power, which flows into the transducer
which support the shaft, as well as between the graphite
brushes and the commutator. Finally, the electrical steel of  PMechanical = TM Ω3 g (6.17)
the rotor becomes magnetized. There is magnetic interaction
between the rotor and stationary permanent magnets, when We must experimentally determine (or be provided) one of
the motor is de-energized. This phenomenon is known as the two transducer equations. Assume that we experimen-
“cogging,” because of the variation in torque needed to ro- tally determined the relationship between the current flow
352 6  Power Transmission, Transformation, and Conversion

through the motor, and the torque produced by the motor to T


be reasonably approximated by this linear relationship: [T ] = [ KT i ] → [ K ] =  i 
T
 
and
 TM = KT iM (6.18)

v
where KT is the motor’s “torque constant.” [ v ] = [ KT Ω ] → [ KT ] =  Ω 
Note that both power flows are shown flowing into the  
transducer interface even though one must flow out, since
the interface cannot store or dissipate energy. The terms “gyrating” and “non-gyrating” transducers are
If we know the experimentally determined relationship, definitions that academics use, which carry no practical sig-
TM = KT iM , then we can use conservation of energy, ex- nificance. A “non-gyrating” transducer is one whose power
pressed in terms of power, to derive the relationship between equations relate the same-type variables to one another.
the voltage drop across the transducer, v2g, and the angular The DC motor is “non-gyrating,” because it relates torque
velocity, Ω3g, to current, and these are both through variables. A gyrating
transducer relates a through variable in one subsystem to an
PA + PB = 0 → PA = −PB across variable in the other.
The sign conventions for electromechanical transducers
(motors) are arbitrary. Although one would expect a conven-
iM v2 g = −TM Ω3 g → iM v2 g = − KT iM Ω3 g tion to exist, there is none. The direction of rotation of a mo-
tor’s shaft depends on position, relative to the front of the
motor. Consequently, how one looks at the motor must be
v2 g = − KT Ω3 g
(6.19) defined for clockwise or counterclockwise to have meaning.
The motion of a motor in response to an applied voltage is
The voltage, v2g, across the electrical branch of the transduc- consistent, relative to the housing of the motor, but indepen-
er, is known as the motor’s “back EMF,” where EMF is the dent of the orientation of the motor in space. Pick a motor
abbreviation for electromotive force. Electromotive force is up and turn it around. It will rotate in the opposite direction,
a nineteenth-century term for magnetically induced voltage. relative to a fixed reference frame. Some motors are “dual
A conductor moving across a magnetic field, such that it cuts shafted” (which is really a single shaft that extends out the
the field lines (Chap. 5 Appendix), creates a voltage in the “front” and the “back” of the housing), to further complicate
conductor. This is the principle of operation of electrical gen- signs. The only useful definition of the positive direction of
erators. The windings on the rotor of a motor cut through the a motor’s rotation is the direction it rotates, when the voltage
stator’s magnetic field lines create a voltage which opposes applied is positive.
the voltage from the source driving the motor. This principle
is Lenz’s law, proposed by Heinrich Lenz in 1833. The sym- 6.4.1.2 Solenoids
bol L for inductance is for Lenz. Induction acts to limit the Ampere coined the term solenoid, Greek for channel, to de-
speed of electric motors. Even a frictionless electric motor scribe the effect of multiple turns of a coil at intensifying the
has a maximum speed, which is the angular velocity at which strength of magnetic flux φ, created by a current i. The term,
the motor’s back EMF is equal in magnitude to the voltage solenoid, is now commonly understood to describe a linear
applied by the power source. electric motor, which is actuated by a coil around a cylindri-
Checking the units of a system equation is more difficult, cal ferromagnetic core, Fig. 6.44. The block model of the so-
if there is a transducer in the system, because the transducer lenoid and its linear graph are shown in Figs. 6.45 and 6.46.
constant appears to have different units in each of the trans- Another style of solenoid is shown in Fig. 6.47. Ener-
ducer equations. In fact, the two sets of units are the same, if gizing the coil moves the ferromagnetic plunger, thereby
expressed in fundamental units. However, do not check the maximizing magnetic flux φ by minimizing magnetic “reluc-
units of the system equation, by expressing the transducer tance” R. All motors can be understood from the perspective
constant in fundamental units. It is easier to simply check the of either motion, resulting from interacting magnetic forces
system equation units, by expressing the transducer constant or torques or motion to maximize magnetic flux φ.
units in terms of both sets of power variable units, and then
making the substitution for the transducer constant units last, 6.4.1.3 Example 1: The Linear Graph of the DC
when it will be clear which set to use, rather than by express- Motor Model
ing all of the parameters of the system in fundamental units, The first step in drawing a linear graph is to find the nodes
which often produces more errors than it reveals. of distinct values of the across variables on the schematic,
Example: Express the units of motor torque constant KT in Fig. 6.48. What are the across variables of the two subsystems
terms of the power variables
6.4 Transducers 353

FM Normally-closed ____
contact 1CR
Spring contact
Logic Output
R iron R axial
Normally-open 1CR shell air gap
contact
Logic Input φ
Coil 1CR

R iron R radial
Fig. 6.44   Schematic of a solenoid in an electromechanical relay. En- plunger
air gap
ergizing the solenoid creates a magnetic force FM which “transitions”
(moves or changes) the spring contact “breaking” the normally-closed
contact and “making” the normally-open contact Fig. 6.47   Cross-section of a solenoid with a free-floating plunger.
Energizing the coil moves the plunger to close the axial air gap, mini-
mizing the magnetic reluctance of the “magnetic circuit”
A= i Mv2g Solenoid B = FM v3g
transducer
interface
1 R 2
Fig. 6.45   Block diagram of positive power flows into a solenoid linear 3
electrical motor transducer element. The symbol for voltage and trans-
lational velocity are the same. The node subscripts identify the element
that the variable is associated with, thus, distinguishing voltages and v(t) MOTOR JM
velocities
bM
Fig. 6.46   Linear graph symbol FM = KF i M g
of the solenoid (linear electric
motor) transducer interface 2 3
Fig. 6.48   DC motor model showing nodes of distinct values of voltage
and angular velocity

iM FM TM = K T i M
R
1 2 3

g
v(t) iM TM J b
of the DC motor? The across variable of the electrical system
is voltage. The across variable of the rotational mechanical
system is angular velocity.
Although the electrical reference voltage (called ground) g g g
and the mechanical non-accelerating inertial reference ve-
locity (also called ground) are two different physical quanti- Fig. 6.49   Linear graph of the DC motor model of Fig. 6.48
ties, they are both represented by the “ground plane” of the
schematic for graphical convenience.
Draw the linear graph, by first drawing and identifying one of the two branches. Do not assign a branch more than
the nodes, and then adding the linear graph elements between one property by adding transduction to it. One way to avoid
the corresponding nodes, Fig. 6.49. There must be a trans- making this error is to draw the interface first and assign
ducer interface symbol in the linear graph to separate the two unique variable names to the through variables in the two
subsystems. Remember, the linear graph symbol for a trans- branches of the transducer’s linear graph symbol. Assigning
ducer has two branches which represent the two sides of the through variable names makes it less likely that a parameter
interface. A branch in the linear graph can only represent one will also be mistakenly assigned to a transducer branch. Fi-
energetic property. A common error is to use a branch in the nally, remember to identify the transducer with one of its two
linear graph, which happens to be parallel to the interface, as transducer equations.
354 6  Power Transmission, Transformation, and Conversion

R R
TM = K T i M
K
1 2 3 2

K
v(t) MOTOR JM JL v(t) iM TM JM bM JL bL
bM bL

g g g g
Fig. 6.50   Schematic of a DC motor driving an inertial load through a
compliant shaft Fig. 6.52   Linear graph of the system shown schematically in Figs. 6.50
and 6.51

1 R 2 Generator
Output External
3 4 Terminal Electrical
K Rwinding Load
v(t) MOTOR JM JL
R Load 2
bM bL T(t) DC
J Generator RLoad1 L
g g g g
b
Fig. 6.51   Schematic with nodes of distinct values of the across vari-
ables voltage and angular velocity
Fig. 6.53   Schematic of a DC generator driven by a torque source

6.4.1.4 Example 2: The Linear Graph for a DC


Motor Driving a Load through will perform a partial reduction to a set of three first-order
a Compliant Shaft differential equations using the “state-space” method.
A dynamic system with a permanent magnet DC motor driv-
en by a voltage source is shown schematically in Fig. 6.50.
The torque produced by the motor is proportional to the cur- 6.4.2 Generators: Mechanical to Electrical
rent flow through the motor, TM = KT iM. The motor has elec- Transducers
trical resistance R, rotational inertia JM, and viscous friction
bM. The motor drives an inertial load, JL, through a shaft with The back EMF of a DC motor is the voltage produced on the
appreciable compliance, described by torsional stiffness K. rotor’s windings as they move through the stator’s magnetic
The inertial load rotates on hydrodynamic bearings with vis- field. A DC motor functions as a DC generator, when the
cous friction bL. motor’s shaft is driven by a mechanical power source with
Again, the first step in drawing a linear graph is to find the intent of converting the mechanical power into electrical
the nodes of the distinct values of the across variables, and power, Figs. 6.53 and 6.54.
identify them on the schematic. What are the across variables
of the two subsystems on either side of the transducer? The
across variable of the electrical system is voltage. The across 6.4.3  Pumps: Mechanical to Fluid Transducers
variable of the rotational mechanical system is angular ve-
locity. Mechanical to fluid transducers are machines which convert
Draw the linear graph, Fig. 6.52, by drawing and identify- translational or rotational mechanical power to fluid power.
ing the nodes, and then add the elements between them. Re- The system dynamics terminology is awkward, since fluid
member that the transducer property of the motor interfaces systems are mechanical systems in that they obey Newton’s
the electrical and mechanical subsystems. The interface is laws.
between the electrical resistor R and the rotational inertia JM. Pumps are classified by the type of fluid they move (either
One would extract the energetic equation from the linear gas or liquid); the motion of the mechanical element(s) which
graph of this system as usual. This system has three inde- moves the fluid; either rotational or reciprocating transla-
pendent energy storage elements and two rotational inertias tional motion; and whether fluid is moved by trapping it in a
separated by the rotational spring. We will not reduce a sys- volume, or by establishing a pressure gradient. Pumps which
tem as complex as this to a single differential equation. We expel fluid trapped in a volume are “positive displacement
6.4 Transducers 355

Fig. 6.54   Linear graph of a DC v2g = KG Ω1g


generator and electrical load R winding R Load 2
driven by a torque source 1 2 3 4

T(t) J b TG iG R Load L
1

g g g g

pumps.” Pumps which move fluid by establishing a pressure tion avoids negative fluid pressure. Negative pressure sig-
differential in the fluid include fans, turbines, and diaphragm nals possible fluid cavitation (vaporization), which is very
pumps. Fluid can flow through a turbine pump, when the destructive to machines. The existence of bubbles caused by
pump is not driven. Conversely, fluid cannot flow through a vaporization is not the problem, although, in general, you
positive displacement pump unless it is driven. want to avoid creating foam or froth when pumping fluid.
The clash of sign conventions between the solid, vector, The damage caused by cavitation is created by the high
and fluid mechanics is beyond awkward. It can lead to con- stresses which occur when the bubble collapses. Unfortu-
tradictory motions between solid elements and fluids. Unfor- nately, the linear graph sign convention often leads to vol-
tunately, the sign conventions used in solid mechanics and ume flow rate Q in a non-sensical direction. It is a no-win
in fluid mechanics are opposite. Positive stress in a solid is situation. Resist the temptation to “fix” the signs until after
tensile stress, whereas positive pressure compresses a fluid. the system equation has been derived.
Negative (absolute) pressure places a fluid in tension. This
state of affairs is a result of sign conventions being estab- 6.4.3.1  Reciprocating Piston Pumps
lished separately in each engineering specialty. The general Reciprocating piston pumps are positive displacement
notion in each field is reasonable enough. Typically, if oppo- pumps. Their simple geometry allows the transducer equa-
site states are possible, then the positive sign indicates what tions to be determined from a drawing. The simplest are a
an engineer prefers. If only one state is possible or common, single piston with a check valves on the inlet and outlet to
then it is assigned the positive state. For example, tensile maintain pressure during the pumping phase, and to prevent
force is preferred in mechanical design, because buckling backflow during the suction phase. Reciprocating piston
cannot occur. Conversely, fluids can support tension in only pumps are capable of enormous pressures. A fluid accumula-
a few, special circumstances. Fluids are usually in compres- tor or capacitor is needed on the output to reduce the pressure
sion. Hence, compressive pressure is positive. The sign re- fluctuations from the reciprocating action.
versal between stress in solids and pressure in fluids is not
the end of the trouble. The sign conventions of solid mechan- 6.4.3.2  Gear Pumps
ics and vector mechanics also differ. In vector mechanics, a A gear pump consists of two, counter-rotating, meshed spur
positive force is a force which acts in the positive direction, gears within a housing, Fig. 6.55. One of the two gears is
relative to a coordinate reference frame. A force which is act- driven by a shaft. The fluid between the gear teeth is moved
ing in the positive direction can either push or pull, putting from the intake side to the discharge side of the pump by
an element in either tension or compression. Added to this the rotation of the gears. The meshing of the gear teeth pre-
confusion is the unfortunate linear graph sign convention vents pressure driven flow from the high-pressure discharge
of assuming both power flows are in the positive direction side to the low-pressure intake side fluid. Similarly, the tight
into the transducer. This is impossible because a transducer’s clearance between the outer surfaces of the gears’ teeth and
only energetic property is to interface two subsystems. Ideal the housing also limits the pressure-driven back flow.
transducers can neither store nor dissipate energy. All energy Since rotational mechanical power drives the pump, and
which flows in must immediately flow out. This sign con- fluid power is delivered by the pump, a gear pump is a trans-
vention leads to the power flows having the same magnitude ducer which interfaces two dissimilar subsystems (Figs. 6.56
but opposite signs. and 6.57). The transducer equations for a gear pump are de-
When working with fluid system transducers, such as rived from the volume of fluid moved by the pump per rota-
a piston-cylinder, define your sign convention, such that a tion of the gears. The volume moved per revolution or cycle
compressive force transmitted to a fluid produces positive is the pump’s displacement:
pressure. Then, use PA + PB = 0 to derive the relationship be-
tween displacement and volume flow rate. This sign conven- volume
displacement ≡ ≡ Vol
revolution
356 6  Power Transmission, Transformation, and Conversion

A = TP Ω1g Gear Pump B = QP p2g


Qp transducer
interface

-Ω1g Ω1g Fig. 6.56   Block diagram of the positive power flows into a gear pump
transducer interface

gallons per minute at 200 psi. Fire engine pumps run of the
truck’s diesel engine through a “split” gear box, which can
Qp transmit the engine’s power to either the pump or the drive
shaft for the wheels.
Fig. 6.55   Gear pump. One of the two gears is shaft driven. The gears
counter-rotate. The fluid moved forward between the gear’s teeth and
the housing is forced out when the gears mesh 6.4.4 Hydraulic Motors Fluid to Mechanical
Transducers
The flow rate of the pump Qp is the volume moved per revo-
lution times the angular velocity of the gears. We need to Hydraulic power is replacing many mechanical drive sys-
work in SI units for volume (m3) and in units of radians per tems, due to the convenience of running a hydraulic line,
second for angular velocity, so we must divide the volume rather than constructing a mechanical transmission to move
per rotation by 2π rad: power from point A to point B, and because the price of the
hydraulic components has decreased significantly. Hydraulic
Vol motors and actuators have “high power density,” meaning
Qp = Ω 1g
2π rad that they can deliver a lot of power using a small motor. Al-
though high power actuators can be small, the entire hydrau-
The second transducer equation for a gear pump is derived lic system can be large. The hydraulic fluid must be pressur-
from the rotational mechanical power and fluid power flow- ized by the fluid power unit consisting of the prime mover,
ing through the pump: either an electric motor or an internal combustion engine; a
gear pump; a fluid reservoir; and a means of cooling the hy-
PA + PB = 0 → TP Ω 1g + QP p2 g = 0 draulic fluid to remove energy dissipated due to shear.
Hydraulic motors are classified by type of motion of the
mechanical elements within the motor (rotation or transla-
 1  tion) and type of motion output by the motor. Internally, the
TP Ω1g + Vol   Ω1g p2 g = 0
 2π  pressurized working fluid can either turn a rotating rotor or
vane within a housing, or displace a piston within a cylin-
 1  1 der. Externally, power is output as either a rotating shaft or
TP + Vol   p2 g = 0 → TP = −Vol   p2 g
 2π   2π  a translating rod. There is a remarkable variety of clever de-
signs.
Again, note that both positive power flows are shown flow-
ing into the transducer interface, even though one must flow
out, since the interface cannot store or dissipate energy. 6.4.5 Linear Hydraulic Motor or Hydraulic
Piston-Cylinder
6.4.3.3  Fans and Turbopumps
Fans and turbopumps are non-positive displacement, axial Hydraulic power systems run at pressures which range from
flow pumps. Their transducer relationship cannot be deter- 750 psi to 15,000 psi, with 2,500 psi as a typical pressure.
mined by simple geometry. Their performance is best char- A hydraulic piston-cylinder, Fig. 6.58, also called a hydraulic
acterized experimentally. The torque driving a fan or pump ram, is a widely used linear “actuator.” It is capable of produc-
is usually unknown, but the shaft speed can be measured. ing large forces, given the high working pressures. Important-
Typically, the pump is spun up to steady-state at a known ly, hydraulic actuators can provide force without motion. The
speed. Then, the pressure and quantity of fluid discharged hydraulic pump driving the actuator circulates the working
during a known time are recorded. An example of a powerful fluid back to the reservoir through a fluid resistor, when there
turbopump is a fire engine’s pump. The current standards for is no flow from the pump, allowing the pump to maintain the
New York City require fire engine pumps to discharge 2,000 rated pressure in the working fluid.
6.4 Transducers 357

Vol
Q P = - _____ Ω 1g
+x,v
2π rad 1
1 2 2 FP
Ap1g
g

TP QP Piston area A

Fig. 6.59   Free body diagram of the piston. The piston force FP is
shown in the positive direction. Seals on actual pistons can produce
shear forces against the cylinder of up to 10 % of the piston force
g

Fig. 6.57   Linear graph symbol of the gear pump transducer interface
Piston with diameter D
Piston with diameter D and area A +x,v
and area A +x,v
FP 2 1 Qp
1 2 FP p1g
Qp
p1g g
g
Vent to atmosphere
Vent to atmosphere Piston or “operating” rod
Piston or “operating” rod
Fig. 6.60   Reversing the orientation of the hydraulic cylinder changes
the transducer signs, because the positive direction of force remains
Fig. 6.58   Cross-section of a single acting hydraulic cylinder and pis-
unchanged
ton. An internal spring or an external mechanism is needed to move the
piston towards the hydraulic fluid port

fluid side, and the piston force in the positive direction for
A hydraulic piston-cylinder is a transducer, since it is translational velocity.
the interface between a fluid system and a translational me- Equilibrium yields
chanical system. A hydraulic ram is described as having high
“power density,” because of the large amount of power it can Ap1g + FP = 0 → − FP = Ap1g
transduce in its relatively small size, due to the high pressure
of the hydraulic fluid. Use the power relationship for the transducer to derive the
Although it is generally best to derive the first of the two second transducer relationship
transducer or transformer equations using a geometric rela-
tionship between the two systems, the transducer equation PA + PB = 0 → QP p1g + FP v2 g = 0
for a hydraulic piston-cylinder is an exception. Using the
force balance between the net pressure force acting on the QP p1g − Ap1g v2 g = 0
fluid side of the piston and the corresponding translational
force on the mechanical side, as the first of the two trans- QP − Av2 g = 0 → QP = Av2 g
ducer equations, allows us to keep the sign of the pressure
in the cylinder positive, when the force acting on the piston
compresses the fluid. Negative pressures are disturbing, be- Although the signs of both transducer equations make sense
cause they lead to fluid cavitation, the development and sub- in this case, there is no reason that they should and they often
sequent collapse of bubbles in the hydraulic fluid. Although do not. If we reverse the hydraulic cylinder from left to right,
it would seem inconsequential, cavitation damages solid sur- Fig. 6.60, and we again define the relationship between pis-
faces, including high strength steel, because of the extremely ton force and pressure, such that pressure is positive, when
high local stresses produced, when the bubbles collapse. the fluid is in compression, we have a sign incongruity be-
The free body diagram of the piston, Fig. 6.59, shows tween volume flow rate into the cylinder, Qp, and the veloc-
pressure acting normal to the surface of the piston on the ity of the piston, v2g.
358 6  Power Transmission, Transformation, and Conversion

A = QP p2g Hydraulic Piston B = FP v3g +x,v


transducer
interface 1 3 2 FP
A αp13
Fig. 6.61   Block diagram of positive power flows into a hydraulic
piston-cylinder Aβ

g
FP = Ap1g Qα Qβ
1 2
Fig. 6.63   Double-acting hydraulic piston-cylinder. Area Aα > Aβ due to
cross-sectional area of the piston rod

QP FP High and Low Pressure Port Pair

Vane
g

Fig. 6.62   Linear graph symbol of a hydraulic piston-cylinder


Stator Rotor

The linear graph sign convention, that both power flows


are positive into a transducer or transformer, Figs. 6.61 and High and Low Pressure Port Pair
6.62, leads to sign problems between the volume flow rate
into the cylinder, Qp = v2g, and the velocity of the piston, v3g. Fig. 6.64   Cross-section of a vane-type hydraulic motor. The vanes are
We can’t make the piston area, A, negative, so the sign of forced out of their pockets in the rotor by fluid pressure. Vane-type
either Qp or v2g must be inverted. Mechanical signs are often hydraulic motors can produce torques of 2,000 ft-lbs
a problem, since we have two independent sign conventions.
A positive force accelerates a mass in the positive direction, inlet port towards the drain port by a drop in pressure, with
or a positive force produces positive stress. Interfacing a me- the pressure difference between the two creating a net force
chanical system with a fluid system increases the sign dif- on the rotor and a resulting torque. A vane motor’s vanes
ficulty, since positive pressure is compressive, and positive are carried in pockets or slots in the rotor, and then forced
normal stress is tensile. Ultimately, one has to return to the out against the smooth but non-cylindrical stator by a spring,
signs defined on both the schematic and the linear graph to fluid pressure behind the vane, or a combination, assisted by
understand what the sign of a variable means. centrifugal force. A rotor-type motor’s vanes are fixed pro-
Hydraulic, pneumatic, and vacuum cylinders are widely jections or lobes on both the rotor and the stator. There is
used in machine design, and there are many variations of one fewer vane on the rotor than the stator. The rotor’s vanes
their design. A fundamental distinction is whether only one or lobes slide past the stator’s with the pinch point moving
side or both sides of the piston can be pressurized. If only along the stator, as the rotor rotates.
one side can be pressurized, the piston-cylinder is “single” Both vane- and rotor-type motors are reversible. Some
acting. If both sides of piston can be pressurized, then it is designs are reversed by reversing the pump and drain con-
“double” acting, Fig. 6.63. nections at a spool valve outside of the motor. Other designs
have a pair of high-pressure inlet ports and control which
port is pressurized. The torque of vane motors is doubled,
6.4.6  Rotational Hydraulic Motors and the side force on its shaft’s bearings eliminated, by hav-
ing two pairs of high pressure and drain ports diametrically
The four broad classes of rotational hydraulic motors design opposite on the stator, Fig. 6.64. A rotor-type motor with a
types are vane, rotor, gear, and piston. single high-pressure inlet is a unidirectional motor. A pair of
high-pressure inlet ports, located on either side of the diame-
Vane- and rotor-type hydraulic motors  The terminology ter through the drain port, allows rotation in either direction.
overlaps between the vane and rotor designs, since both have An interesting hydraulic motor which combines aspects of
rotors and vanes. Their rotors rotate from the high-pressure a rotational motor and a gear motor is a “gerotor” type motor
6.4 Transducers 359

Two High Pressure Ports cantly higher than the 2,000 psi typical of internal combus-
tion engines. There are an odd number of pistons in a rotor
that resembles the cylinder of a large caliber revolver hand-
gun, Fig. 6.66. The operating rods of the pistons have a ball
Rotor end which is captured in a socketed plate. If the socketed
plate can be tilted, it is called a “swash” plate. If it is fixed
in position, it is called the “angle” plate. The axis of sock-
eted plate is inclined, relative to the axis of rotor. The dis-
placement of a piston in a pressurized cylinder increases, as
Roller Vane the rotor turns towards the top position, and then decreases,
Stator as the rotor turns through the drained side. The tangential
Drain Port component of the force, exerted against the angle plate by a
pressurized piston’s operating rod end, creates a torque about
Fig. 6.65   Cross-section of a rotor-type hydraulic motor the axis of rotation. The base of each cylinder has a pressure
port which aligns with a stationary valve plate, as the rotor
revolves. The top cylinder (at the maximum piston displace-
shown in Fig. 6.65. A gerotor resembles a rotor-type pump, but ment) and cylinders on one side of the rotate are pressurized.
the fixed peak-like vanes of the stator are replaced by rollers. The cylinders on the other side are drained. The direction of
Note that there are more lobes on the housing than there are on rotation is reversed by pressurizing the cylinders that were
the rotor. (The lobes in this design are steel rollers into the plane drained, and draining those which were pressurized, except
of the view.) High-pressure fluid introduced from a port, where the top cylinder that is always pressurized.
the volume between the lobes of the housing and rotor is small,
causes the rotor to rotate towards the low pressure of the drain,
located where the volume is large. It is called a “roller-vane” 6.4.7 Example 3: Linear Graph
type, because the lobes on the rotor function as the vane in a of a Fourth-Order
conventional pump. The vanes contact rollers, rather than fixed Fluid-Mechanical System
surfaces, to reduce wear.
The order of a system is determined by the number of inde-
Gear motors Unidirectional gear motors are, essentially, pendent energy storage elements in the system. The system
gear pumps driven by fluid pressure. They are capable of shown schematically in Fig. 6.67 consists of a torque source
speeds up to 10,000 RPM, but have relatively low efficiency, driving a gear pump. The gear pump has rotational inertia J,
due to flow through the clearance between the gears and the and a by-pass flow path with resistance R1, which allows the
housing. Bi-directional gear motors need a pair of input ports pump to run and maintain pressure, when there is no flow
to create a preference pressure drop in either direction. out of the pump. The fluid discharged by the pump flows
through a fluid resistance R2 and an inertance I, before reach-
Piston hydraulic motors  The operating pressure of piston ing a fluid capacitance (or fluid accumulator) C immediately
hydraulic motors ranges from 3,000 to 15,000 psi, signifi- before a hydraulic piston-cylinder. The actuator moves mass

Fig. 6.66   A cross-section of the Angle plate rotates the output shaft
rotor and output shaft and a bot- Rolling-contact
tom view of the rotor of a piston- bearings Piston rod ends captured
type rotational hydraulic motor
in ball and socket joints
Output shaft
These fluid ports pressurized
for counter-clockwise rotation

Rotor rotates from pressure side through


maximum piston displacement to drain side Bottom View of Rotor
360 6  Power Transmission, Transformation, and Conversion

Fig. 6.67   Fourth-order fluid- Hydraulic


mechanical system Piston
Area A
C
T(t) R2 I
J Pump M
b R1

Fig. 6.68   Velocity and pressure Hydraulic


nodes of the fourth-order fluid- Piston
mechanical system Area A
C
1 4
T(t) 2 R2 3 I 5
J Pump M
g
b R1
g g g

Fig. 6.69   Linear graph of the Vol


fourth-order fluid-mechani- Q P = ___ Ω 1g R2 I FA= Ap4g
cal system 2π
1 2 3 4 5

T(t) J TP QP R1 C QA FA M

g g

M. This system is fourth order, because there are four inde- 6.4.8 Rotational to Translational Mechanical
pendent energy storages, J, I, C, and M. Transducers
The first step in drawing a linear graph is to find the nodes
of the across variables and identify them on the schematic of 6.4.8.1  Rack and Pinion
the system. What are the across variables of the three sub- A rack is a “linear” gear. Pinion means “small gear.” A rack
systems? The across variable of the rotational mechanical and pinion is a transducer, since they interface rotational and
system is angular velocity Ω. The across variable of the fluid translational subsystems, Fig. 6.70.
system is pressure p. The across variable of the translational The analysis of a rack and pinion requires either knowing
mechanical system is translational velocity v, Fig. 6.68. or calculating the pitch radius of the pinion, so that the an-
Draw the linear graph by first drawing and labeling the gular velocity of the pinion can be related to the translational
nodes, Fig. 6.69. Next, add the two transducer interfaces be- velocity of the rack
tween their respective nodes. Then add the remaining ele-
ments between the corresponding nodes. Remember that v2 g = rp Ω1g
each branch in the linear graph represents a distinct power
flow. Never assign two properties to one branch, by using a or knowing either the circular pitch of the pinion, ppinion and
branch belonging to an element such as the inertia, fluid ca- its number of teeth, Npinion, or the lineal pitch of the rack, prack
pacitance, or mass as one side of a transducer interface. The and the number of pinion teeth, so that translational displace-
two branches of the transducer interface represent the power ment x and rotational displacement θ, and, hence, velocities
flow passing across the interface. They cannot represent any can be related:
other power flow in the system.
p pinion N pinion prack N pinion
x2 g = Ω1g or x2 g = Ω1g
2π 2π
6.4 Transducers 361

Pin constrained by slot


to translate in contact with
+θ,Ω the root of power screw.
-x,v +x,v

+x,v
1
2

Fig. 6.70   Rack and pinion showing positive directions for translation +Θ, Ω
and rotation 2π
__
∆x
+x,v
Fig. 6.72   Pin riding on the root of a power screw
2
1
n threads n revolutions n 2π rad
≡ =
in. in. in.
+θ,Ω
Although it is possible to perform system dynamics calcula-
Fig. 6.71   Power screw and nut. The nut is constrained from rotation tions in US customary units, it is inadvisable, due to errors
and must translate along the power screw in converting units of weight to units of mass. It is preferable
to calculate using SI units, even when the results are to be
expressed in US customary units:
The analysis then uses energy conservation expressed in
terms of power, since only the transducer attribute of the rack n 2π rad n 2π rad  1 in.  rad
= ≡N
and pinion is being considered, and it neither stores nor dis- in. in.  0.0254 m  m
sipates energy:
Since the power nut cannot rotate, it must translate to accom-
PA + PB = 0 → TP Ω1g + FR v2 g = 0 modate the rotation of the screw. The translation x of the nut
is related to the angular displacement θ as
TP Ω1g + FR rp Ω1g = 0
1 m
x2 g =  θ
 N rad  1g
TP = − rp FR
To express this in terms of the power variables, v and Ω, we
differentiate both sides with respect to time
6.4.8.2  Power Screw
A power screw interfaces rotational power and translational dx2 g 1 m  dθ1g 1 m
= → v2 g =  Ω
power. The nut is constrained from rotation and can only dt  N rad  dt  N rad  1g
translate, Fig. 6.71. The relative motion of the power screw
and nut may be easier to visualize in Fig. 6.72 showing a The second transducer equation is derived from energy con-
pin, which represents a point on the thread of the power nut, servation expressed in terms of power, because, again, the
following just one axis of the helix of the root of the power transducer represents only the transduction of the power
screw thread. The nut cannot rotate. It is constrained to trans- from one type of power (or energy domain) to the other:
lation. The screw cannot translate. It is constrained to rota-
tion. PA + PB = 0 → TS Ω1g + FN v2 g = 0
The transducer equation is based on the distance the
power screw advances the nut with one rotation. Hence, we 1 m 
TS Ω1g + FN  Ω =0
must know the pitch of the screw. Say screw has n threads  N rad  1g
per inch. A “thread” in this usage represents a full rotation
or revolution. We must represent angular displacement using
radians, not revolutions or degrees. Note that all “units” of 1 m
TS = − FN 
angular displacement are, in fact, dimensionless  N rad 
362 6  Power Transmission, Transformation, and Conversion

Crank
Follower Connecting rod
+x,v
Slider
Θ1 r1 1

Cam
Θ2
+θ,Ω +x,v 2
+Θ,Ω θ
r2
Fig. 6.74   Crank and slider. This is an in-line or centric crank and
Fig. 6.73   Cam and roller follower slider, because the line of motion of the slider intersects the center of
rotation of the crank.

Power screws are designed to transduce rotational power


into translational power. The pitch of the power screw makes In the past, cams were relied on to perform some tasks,
them “self-locking,” meaning that they cannot be driven by such as automotive fuel injection, which are now accom-
translating the nut. Consequently, power screws are non- plished using solenoids. The disadvantage of cams is that
linear transducers, since power can flow from the rotational they are “hard” automation, meaning that to change the mo-
power into translational power, but not in the reverse direc- tion or timing of a mechanism it must be redesigned and re-
tion. fabricated, rather than simply reprogrammed. The sophisti-
cation of feedback control in an automobile engine, and the
6.4.8.3  Cam and Follower speed of modern “power train control modules” (micropro-
A cam and follower, Fig. 6.73, is used to drive a translation- cessors which control both the engine and automatic trans-
al mechanism repeatedly through identical displacements. mission), makes it inevitable that future internal combustion
For example, cams are used to open the intake and exhaust engines will use solenoids, rather than cams to control the
valves of automobile engines, synchronized with the motion valves, to yield greater fuel efficiency and lower pollution.
of the pistons in the cylinders. The cams, which activate the
valves, are machined onto the cam shaft. The cam shaft is
driven by a “timing” chain or belt powered by the “crank” 6.4.9 Translational to Rotational Mechanical
shaft, which, in turn, is driven by pistons during the combus- Transducers
tion phase of the cycle.
The displacement of the follower equals the change in the 6.4.9.1  Crank and Slider
radius of the cam. The function of angular displacement θ Depending on the design, a crank and slider mechanism can
which describes radius of a cam, r (θ ), should be a smooth- either convert circular motion to reciprocating linear motion
ing varying function. Otherwise, the profile will wear at the or vice versa, Fig. 6.74. The relationship between the angular
discontinuities in the slope of the cam’s surface. displacement of the crank and the translational displacement
In automotive design, a “tappet” is used as the element in of the slider is non-linear, and is described by trigonometry.
contact with the cam’s surface. A particularly clever design, An example of a crank and slider mechanism being used to
which dates from the 1930s, is a “hydraulic” tappet, used to convert translation to rotation is a piston engine, where the
reduce the effect of wear in the contact surfaces of the “valve reciprocating translational motion of the pistons within their
train” (consisting of the cam, tappet, lifting rod, “rocker cylinders is converted to rotational motion by the crank shaft.
arm,” a lever which engages the end of the valve stem). A
hydraulic tappet uses the engine oil, bathing the cam shaft, as
hydraulic oil pressure to extend the tappet, so that it remains 6.5  Multiport Transformers and Transducers
in contact with the cam and lifter rods. The design can be
viewed as a fluid capacitor in series with a non-linear fluid Thus far, we have considered transformers and transducers
resistance created by a check valve. The time constant for with two “ports” through which energy passes from one sub-
the tappet to extend and fill with oil (when lightly loaded) system to another. It is common in machine design to have
is smaller than the time constant for the tappet to compress mechanical transformers and transducers with one input port
and expel oil. Hydraulic tappets are now obsolete, having and many output ports. For example, the “serpentine” belt
been replaced by cam followers with rollers as the contact. drive in an automobile is driven by a pulley on the crank
Although a roller-type follower is more complicated, its re- shaft and, in turn, drives the water pump, the alternator, and,
duced friction increases gas mileage.
6.5  Multiport Transformers and Transducers 363

PowerB LA LB
b
F(t)
LC

Pivot
Power A Transformer PowerC
or
Transducer K

Fig. 6.75   Three-port transformer or transducer block model showing


the positive power flows
Fig. 6.76   Translational mechanical system

often, a power steering pump and air conditioner compres-


sor. g LA LB
Extending the linear graph of a two-port transducer or b
transformer element to a multiport element is straightfor- F(t) 3 g
ward. Adding a third port to the “block” model of a trans- LC
ducer, Fig. 6.75, identifying that power flow as C, and, again, 2
Pivot
recognizing that the transformer or transducer element’s only 1
attribute is transformation or transduction (i.e., no energy g
K
storage or dissipation) allow us to write energy conservation
at an instant, in terms of power, yielding

g
∑ Power = Power A + PowerB + PowerC = 0
Fig. 6.77   Schematic of translational mechanical system with velocity
The power flow into the transformer or transducer at a nodes
port must flow out of one or both of the other two. The book-
keeping is marginally more involved than that of a two-port 4. Mechanical subsystems are restricted to a single direc-
transducer, because the outward (or negative) flow of power tion, so that the power variables can be represented as
can divide between the available ports. The division of the scalars, rather than as vectors.
outward power flows results from either the summation of
the through variable (force or torque) at a node, or across Within these constraints, we are free to define the positive
variable (velocity) around a loop, as illustrated in the follow- direction as we wish. If the input is represented as a vector,
ing examples. we can choose to use the direction of the input as the positive
direction. Alternatively, we can choose to define a coordinate
system. In any event, we can choose to define the positive
6.5.1 Example 1: A Lever with Three direction. We will establish signs graphically on the system’s
Attachments schematic, as needed to formulate the system equation, and
then interpret the results of calculations, with reference to
The “L” lever shaped as shown in Fig. 6.76 is pivoted at the the linear graph and the schematic.
intersection of its two arms. Recall that levers are modeled as perfectly rigid and as
This system presents two challenges. First, how does one undergoing “small” displacements, where small means
define the positive direction? Second, how does one model “small enough,” that the difference between the arc, which
three elements acting on a lever? a node transverses, and its horizontal or vertical compo-
Signs are a problem in system dynamics. The only con- nent can be neglected, allowing the motion to be considered
straints we must respect are the sign conventions of the lin- translational. We will establish the signs for the translational
ear graph method. motion of nodes on the lever to be positive, when the applied
1. Sources raise the across variable, translational velocity in force, F( t), acts in the direction shown, Fig. 6.77. In other
this case, in the positive direction indicated by the arrow- words, counterclockwise rotation of the lever produces posi-
head orienting the element. tive translational displacement of nodes on the lever.
2. The through variables of both branches of a transducer
or transformer are oriented towards ground, and power Multiport Transformer Model Now consider the lever.
flows in the two branches sum to zero. The key is to view the system as both a mechanical system
3. In all other elements, the positive direction is arbitrary. and an energetic system, meaning that we will use equilib-
364 6  Power Transmission, Transformation, and Conversion

2 1 3 2 1 3

K Fβ Fα F(t) Fγ Fδ b

g g g g g g

Fig. 6.78   Transformer nodes of the L-shaped lever system of Fig. 6.77 Fig. 6.80   Remaining elements added to the linear graph

2 1 3 2 1 3

Fβ Fα Fγ Fδ Fβ Fα

g g g g g g
Fig. 6.79   Transformer interface branches added between the interface Fig. 6.81   Transformer interface between nodes 1 and 2
nodes of the L-shaped lever system

rium, geometric compatibility, and conservation of energy to v1g v2 g LA + LB


Ω= = → v1g = v2 g
develop a model. We will start with conservation of energy. ( LA + LB ) LB LB
Recall that the energetic attribute of the transformer (and
its linear graph element) only transforms power variables. Now, look at the linear graph. The branches on either side
There is no energy storage or dissipation in the transforma- of the interface define the power flows into and out of the
tion. Consequently, we can express conservation of energy interface. Remember that power is the product of the through
at any instant in terms of power flows. The lever divides the and across variables of an element. Hence
power entering it between the other elements connected to it.
We will draw a linear graph of this system by following Pα + Pβ = 0 → Fα v1g + Fβ v2 g = 0
our usual procedure. First, identify the nodes of discrete ve-
locities on the schematic of the system, which we have al- Eliminate the velocities by substitution:
ready done. Next, start the linear graph by drawing and iden-
tifying the nodes, Fig. 6.78. Then, add the interface branches  L + LB 
of the transformer, Fig. 6.79. Remember that an interface has Fα v1g + Fβ v2 g = 0 → Fα  A v2 g + Fβ v2 g = 0
 LB 
two sides, i.e., two branches. Each interface (i.e., each pair
of interfaces branches) can only share one node, which is  L + LB 
ground, in this system. Finally, add the remaining elements Fβ = −  A Fα
 LB 
to the linear graph, Fig. 6.80.
All that remains is to determine the transformer equations
for both interfaces. They are determined individually, as no Repeat the process to determine the transform equations for
other element in the system existed. Starting with the inter- the interface between nodes 1 and 3, Fig. 6.82.
face between nodes 1 and 2, Fig. 6.81, we first use geometric Geometric compatibility ( LA + LB )Ω = v1g and LC Ω = v3 g
compatibility to write an equation relating the velocities, v1g
and v2g. The rigid lever rotates with a single angular velocity v1g v3 g LA + LB
Ω= = → v1g = v3 g
Ω. The translational velocities vary with the distance of the ( LA + LB ) LC LC
nodes from the center of rotation:
Energy conservation in terms of power
( LA + LB ) Ω = v1g and LB Ω = v2 g Pγ + Pδ = 0 → Fγ v1g + Fδ v3 g = 0

Substitute to eliminate velocities


6.5  Multiport Transformers and Transducers 365

2 1 3 LA + L B LA + L B
v1g= ____ v2g v1g= ____ v3g
LB LC
1 2 3
Fγ Fδ

F(t) Fα Fβ K Fγ Fδ b
g g g

Fig. 6.82   Transformer interface between nodes 1 and 3 g g g

Fig. 6.84   Equivalent linear graph of the L-shaped lever system,


LA + L B LA + L B Fig.  6.77. This linear graph has the same “topology” or connections
v1g= ____ v2g v1g= ____ v3g between elements as the linear graph in Fig. 6.83
LB LC
2 1 3

Ω2 Ω1 Ω3
K Fβ Fα F(t) Fγ Fδ b

g g g
Gear A Pinion Gear B
Fig. 6.83   Linear graph of the L-shaped lever system, Fig. 6.77, with N 2 Teeth N 1 Teeth N3 Teeth
transformer interfaces identified with one of the two transformer equa-
tions
Fig. 6.85   A pinion driving two gears

 L + LB 
Fγ v1g + Fδ v3 g = 0 → Fγ  A v3 g + Fδ v3g = 0
 LC  6.5.2  Example 2: A Pinion Driving Two Gears

 L + LB  The linear graph for the gear set shown in Fig. 6.85, is drawn
Fδ = −  A Fγ by, as always, first identifying the nodes of distinct values
 LC 
of the across variable angular velocity, and then adding the
elements, Fig. 6.86. We have not identified what type of ele-
The linear graph of the bent lever system is completed by ment is powering the pinion. It could be a torque or velocity
identifying the transformers with a transformer equation, source, it could be a compliant shaft modeled as a spring,
Figs. 6.83 and 6.84. or a fluid coupling modeled as a damper. (It could not be an
There are two transformer equations for each transformer inertia. If you are unsure why not, then review rotational ele-
interface. The transformer equations added to the list of en- ments.) Let us keep the model general, and simply identify
ergetic equations are the torque as TA.
Leverage between nodes 1 and 2 Since we are dealing with a mechanical system, our
first preference is to use geometric compatibility to derive
LA + LB  L + LB  a relationship between the angular velocities. The surface
v1g = v2 g and Fβ = −  A Fα
LB  LB  speed of the pinion and both gears A and B is the same, or
they would not mesh. They would shear off teeth.
Leverage between nodes 1 and 3 The number of teeth in each gear of a gear set is a measure
of its relative pitch circumference, and, hence, its relative
LA + LB  L + LB  pitch radius:
v1g = v3 g and Fδ = −  A Fγ
LC  LC 
rp ∝ N1 rA ∝ N 2 rB ∝ N 3

The surface speed is the product of the angular velocity of a


gear and its pitch radius. Hence
366 6  Power Transmission, Transformation, and Conversion

Mesh between Mesh between rA


Pinion and Gear A Pinion and Gear B Ω3g = __ Ω
2 1 3 r C 1g
2 1 3
TA
Tβ Tα Tγ Tδ
Tγ Tδ

g g g
g g g
Fig. 6.86   Linear graph of pinion driving gears A and B, Fig. 6.85
Fig. 6.88   Transformer interface between the pinion and gear B

Mesh between
Pinion and Gear A
2 1 3 also require more computation. Again, our preference, when
we can choose, is to use geometric compatibility, rather than
equilibrium, because it is easier and less error prone.
The power flows down either side of each gear mesh in-
Tβ Tα terface are equal

Pα + Pβ = 0 → Tα Ω1g + Tβ Ω 2 g = 0
g g g
Eliminate the velocities by substitution
Fig. 6.87   Transformer interface between the pinion and gear A
 N 
Tα Ω1g + Tβ Ω 2 g = 0 → Tα Ω1g + Tβ  − 1  Ω1g = 0
 N2 
surface speed = rp Ω 1g = rA Ω 2 g = rB Ω 3 g
N1
Tα = Tβ
N2
N1Ω 1g = N 2 Ω 2 g = N 3 Ω 3 g

where the magnitude bars are needed, because speed is a sca- The process is repeated to derive the transformer equations
lar, but velocity is a vector. We must inspect the schematic for the mesh between the pinion and gear B:
to determine the direction of rotation of the gears, yielding

  N 
Pγ + Pδ = 0 → Tγ Ω1g + Tδ Ω3 g = 0
N1 
Ω2 g =  −  Ω1g and Ω3 g =  − 1  Ω1g
 N2   N3   N 
Tγ Ω1g + Tδ Ω3 g = 0 → Tγ Ω1g + Tδ  − 1  Ω1g = 0
Now, examine the linear graph. The interfaces between the  N3 
pinion and gear A and the pinion and gear B required us to
N1
divide the through variable torque supplied to node 1 as TA Tγ = Tδ
into two torques, which we named Tα and Tγ, Figs. 6.87 and N3
6.88. The node equation for node 1 is not used to derive the
transformer equations. It is certainly essential to describe the The two transformer equations for the two interfaces,
system and will appear in the equation list, but the trans- Fig. 6.89, would be added to the list of energetic equations as
former equations for each interface are independent of the Mesh between pinion and gear A
other elements in the system.
Recall that the transformer equations for gears can be de-  N1  N
Ω2g =  −  Ω1g and Tα = 1 Tβ
rived from any two of the three equations, which describe  N 2 N 2
the interaction of the gears in terms of geometric compat-
ibility, expressed as equal surface speeds. Energy conserva- Mesh between pinion and gear B
tion is expressed in terms of power flows, and equilibrium
of tooth forces. Tooth force equilibrium calculations require  N1  N1
Ω3g =  − Ω1g and Tγ = Tδ
the pressure angle and pitch radii of the pinion and gear and  N 3  N3
6.5  Multiport Transformers and Transducers 367

N N rA rA
Ω 2g = - __1 Ω1g Ω 3g = - __1 Ω1g Ω 2g = __ Ω3g = __ Ω
N2 N3 r Ω1g
B r C 1g
2 1 3 2 1 3
TA TA
Tβ Tα Tγ Tδ Tβ Tα Tγ Tδ

g g g g g g
Fig. 6.89   One of the two transformer equations for each interface is Fig. 6.91   Transformer linear graph elements of belt drive shown in
added to identify the power transformations of the pinion driving two Fig. 6.90
gears, Fig. 6.85

rA
Pulley B Ω 2g = __
r Ω1g
Radius rB B
Ω 2g 2 1 3
Idler
Idler

Tβ Tα
Pulley C
Ω3g Radius rC
Pulley A Ω1g
Radius rA g g g
vsurface
Fig. 6.92   Transformer interface of pulleys A and B
Fig. 6.90   A belt drive with one driving pulley, two driven pulleys, and
two idlers
where the magnitude must be used, because the direction of
the angular velocity depends on the direction of the belt’s
6.5.3  Example 3: A Belt Driving Two Pulleys wrap around the pulley, i.e., clockwise or counterclockwise.
The angular velocities of the pulleys are related by the
The belt drive shown in Fig. 6.90 has five pulleys, two of inverse ratio of their radii, with the sign of the rotation of a
which are idler pulleys that turn freely. Idler pulleys are used pulley determined by inspection of the schematic, and with
in belt drives to direct the belt to a driven pulley and increase the direction of the rotation input pulley generally defined as
the “wrap” or contact area between the belt and a pulley. Idler positive. In the belt drive above, pulleys A, B, and C rotate
pulleys also help to provide sufficient shear stress against the in the same direction. Therefore,
face of the pulley in order to provide the required torque on
the pulley’s shaft. Belt drives typically include a “tensioner” Ω4 g Ω5 g Ω6 g rB r
= = → Ω4 g = Ω5 g and Ω 4 g = C Ω 6 g
which is a spring-loaded, pivoted arm that presses an idler rA rB rC rA rA
pulley against the belt.
The derivation of the transformer equations for a serpen- The linear graph of the belt drive is drawn, as always, by
tine belt drive is very similar to a gear set. Both are rota- first identifying nodes of distinct values of the across vari-
tional systems, where the geometric compatibility between able angular velocity, as we have, then drawing the nodes,
interfaces is surface speed. In a belt drive system, the surface adding the two branches of each transformer interface, and,
speed of the belt is constant along the belt. If the belt is mod- lastly, inserting the remaining branches, Fig. 6.91. We will
eled as inextensible, then it can bend but cannot stretch. The let torque TA from an unidentified element drive the pulley A.
belt’s surface speed equals the product of the magnitude an- Having used geometric compatibility to derive one of the
gular velocity of a pulley times the pulley’s radius: two transformer equations for each interface, we now pro-
ceed to derive the second transformer equation for each, by
surface speed ≡ vsurface = rA Ω 4 g = rB Ω5 g = rC Ω6 g working with only one interface at a time. We will start with
the interface between pulleys A and B, Fig. 6.92.
368 6  Power Transmission, Transformation, and Conversion

rA 6.6 Floating Sources, Transformers,


Ω3g = __ Ω and Transducers
r C 1g
2 1 3
The term “floating” describes an energetic element which
is normally connected to a constant ground reference, but,
in a particular case, is not. Floating sources, transformers,
Tγ Tδ and transducers are widely found in machine design, in large
part due to the need to damp out vibrations, or limit “shock”
loading in mechanical systems. Transformers or transducers
g g g in series are used to provide larger power variables than a
single element can, as well as to provide two or more “taps”
Fig. 6.93   Transformer interface of pulleys A and C of different magnitudes of the power variable. An example of
transducers connected in series is a two-stage air compres-
sor, in which the first stage draws air from the atmosphere
Conservation of energy across the interface between pul- and supplies pressurized air to the second stage.
leys A and B, expressed instantaneously in terms of power The only energetic elements which must be referenced to
flows, is the sum of the products of the through and across a “ground” are the mechanical inertial elements, mass and
variables of the two branches of the interface mass moment of inertia (or rotational inertia), and fluid ca-
pacitances. Inertial elements need to be referenced to a non-
Pα + Pβ = 0 → Tα Ω 4 g + Tβ Ω5 g = 0 accelerating velocity node. (By the way, the most recent es-
timate of your translational velocity, as you orbit the Milky
Eliminate the angular velocities by substitution Way’s galactic center, is 600,000 mph). Fluid capacitances
must be referenced to the environment pressure, usually at-
rB mospheric pressure. All other energetic elements can act be-
Tα Ω 4 g + Tβ Ω5 g = 0 → Tα Ω5 g + Tβ Ω5g = 0
rA tween any two nodes in a system.
Although we most often model source elements as ref-
rB erenced to (or attached to) a ground node, they need not be.
Tβ = − Tα
rA Sources, transformers, and transducers do not need to be ref-
erenced to ground, either. However, keep in mind that real
The procedure is repeated for the interface between pulleys devices which act as sources, or perform energy transforma-
A and C, Fig. 6.93. tion or transduction, have mass or mass moment of inertia.
These energetic attributes must be referenced to ground.
Pγ + Pδ = 0 → Tγ Ω 4 g + Tδ Ω 6 g = 0 Modeling systems with floating sources, transformers, or
transducers is no more difficult than modeling systems in
Eliminate the angular velocities by substitution which these elements are connected to ground. The assumed
positive directions of the through variable of both branches
rC of the transformer or transducer interface must be in the di-
Tγ Ω 4 g + Tδ Ω 6 g = 0 → Tγ Ω 6 g + Tδ Ω 6 g = 0
rA rection of a drop of their corresponding across variable, when
the input to the system is a positive step. If the through vari-
rC able is in the direction of an increase of the across variable,
Tδ = − Tγ
rA you are violating energy conservation. Sources are the only
energetic elements, in which the across variable increases in
The transformer equations added to the energetic equations the direction of the through variable flow.
are:
Interface between pulleys A and B
6.6.1  Floating Mechanical Sources
rB r
Ω4 g = Ω 5 g and Tβ = − B Tα Floating mechanical sources are very common, due to the
rA rA
need to control vibrations. A typical motor mount is a bolt-
Interface between pulleys A and C ed connection in which an elastomeric pad (i.e., rubbery

rC r
Ω4 g = Ω 6 g and Tδ = − C Tγ
rA rA
6.6  Floating Sources, Transformers, and Transducers 369

Concentric torsion spring and drag cup b 1


b K
g T(t) J 2
K
T(t) J

Rigid mounting plate Rigid shaft

Rigid mounting plate Rigid shaft


Fig. 6.95   Schematic of an automobile engine model with angular ve-
locity nodes
Fig. 6.94   Schematic of an automobile engine modeled as a torque
source driving a rotational inertia and mounted on a concentric torsion-
al spring and drag cup
1 T(t) 2

polymer) carries some of the load in compression. A motor


mount is modeled as a spring and damper in parallel. Motor b K J
mounts carry two simultaneous (or superimposed) loads, the
“dead” or constant load, which is the weight of the motor, and
the “live” or variable load, due to the transient and steady-
state reaction forces and torques of the motor created while g
it powers the machine. The dead loads are not included in
Fig. 6.96   Linear graph of the automobile engine model with floating
the energetic model, because they have no effect on the dy-
torque source
namic response of a linear system. They are important when
modeling a non-linear system. The elemental parameters of
a non-linear system are functions of the power variables. For
example, the stiffness of a non-linear spring changes with the C2
force or torque it carries. Consequently, dead loads must be R2
p2(t)
included in non-linear models, because they determine the
Pump 2
M
“operating point” at which the transient response occurs. R3
An example of a system which includes a floating source p1(t) R1
is an automobile. If you have ever watched an automobile Pump 1
C1 Hydraulic Piston
engine as it is “revved” (or its angular velocity is acceler- Area A
ated), you will have noticed that the engine rocks, due to the
reaction torque rotating the engine in the opposite direction
of the accelerating rotating parts. One or two sets of motor
Fig. 6.97   Schematic of a third-order fluid-translational mechanical
mounts will isolate the passenger compartment from the vi-
system with two pumps. Pump two is “floating”
bration created by the engine, depending on whether the en-
gine is mounted to the “frame,” comprised of the main struc-
tural elements, which runs the length of both sides in a rear 6.6.2  Floating and Multiple Fluid Sources
wheel drive automobile, or to a “subframe” in front wheel
drive automobiles. Any number of sources can be used in a machine, if they are
The engine mounts are lumped into one concentric (or controlled properly. A fluid-translational mechanical system
coaxial) torsion spring K and rotational damper, or drag cup, with two pumps modeled as pressure sources is shown in
b, see Figs. 6.94, 6.95 and 6.96. The schematic symbol repre- Fig. 6.97. Energy can be stored independently in two fluid
sents a cross-section through a coaxial drag cup and torsion- capacitances and the mass, making it a third-order system,
al spring, but is easy to mistake for two drag cups and one Figs. 6.98 and 6.99.
spring. If you are uncertain, check. It should be referenced in
call-out, parts list, or other documentation, or in this course,
the problem statement. The engine is modeled as a torque 6.6.3  Floating and Multiple Electrical Sources
source.
Torsion spring K and damper b provide the reaction nec- 6.6.3.1  RC Circuit with Two Voltage Sources
essary for the torque source to apply torque to the mass mo- Electric batteries are connected in series to increase voltage
ment of inertia J. and in parallel to increase current. Standard A, AA, AAA, C,
370 6  Power Transmission, Transformation, and Conversion

2 R2
C2
4 5 + Switch 2
i R2
p2(t) 3 R2 iB 2
1.5 VDC
Pump 2 M
R3 - R1
p1(t) 1 R1 C1 1 3
Pump 1 2
Hydraulic Piston + Switch 1
i R1
Area A
1.5 VDC iC
iB C
g 1
-
g
g
Fig. 6.98   Schematic of a third-order fluid-translational mechanical
system with two pumps showing nodes of distinct values of pressure Fig. 6.100   Electric circuit with two voltage sources (batteries) in series
and translational velocity

Table 6.1   Circuit configurations


FP = Ap4g Energized Switch 1 Switch 2
R2
configuration
3 4 5 One Closed (conducting) Open (non-conducting)
Two Open (non-conducting) Closed (conducting)
Three Closed (conducting) Closed (conducting)
p2(t)

R1
1 2
R3 C2 QP FP M R1
1 3
+ Switch 1
i R1
p1(t) C1 iC
1.5 VDC iB
1 C
-
g g
Fig. 6.99   Linear graph of the third-order fluid-translational mechani- Fig. 6.101   Energized configuration one
cal with two pumps

2 R2
and D “cell” chemical batteries provide a voltage increase
from their negative terminals to positive terminals of 1.5 + Switch 2
i R2
VDC. The schematic symbol for a battery is intended to re-
1.5 VDC iB 2
semble a voltaic “pile.” The end with the long bar is positive,
the “anode.” The negative end is the “cathode.” -
The RC circuit shown in Fig. 6.100 has two batteries in 1 3
series. Node 1 and node 2 can be switched independently. +
Consequently, there are three possible energized configura- 1.5 VDC iC
iB C
1
tions, Table 6.1.
Each of the energized configurations yields a different
-
system equation. Energized configurations one and two are g
simplified, by identifying the circuit elements with no cur-
rent, during the transient portion of the step response, yield- Fig. 6.102   Energized configuration two
ing straightforward RC circuits, Figs. 6.101 and 6.102.
The third energized configuration with both Switch One
and Switch Two closed, Fig. 6.103, has not been considered cally, is the step response of configuration three the sum of
previously. the step responses of the configurations one and two? No,
How do we determine its system equation for the voltage summing the step responses of configurations one and two
across the capacitor, v3g? Can we use superposition? Specifi- do not yield the step response of configuration three.
6.6  Floating Sources, Transformers, and Transducers 371

2 R2 v21 + v1g − v3 g 1.5 VDC + 1.5 VDC − 4.5 VDC


= = iR2
R2 R2
+ Switch 2
i R2
−1.5 VDC
1.5 VDC iB 2 iR2 =
R2
- R1
1 3 Hence
+ Switch 1
i R1
1.5 VDC iC −3.0 VDC −1.5 VDC
iB C iR1 + iR2 = iC → + = iC
1
- R1 R2

g Although steady current flow through the resistances is ac-


ceptable in steady-state, the steady-state capacitor current
Fig. 6.103   Energized configuration three must be zero, or the energy in the capacitor would change,
Fig. 6.104. For the voltage across the capacitor, v3, and,
1
To understand the response of configuration three, we hence, the energy in the capacitor EC = Cv32g to be con-
2
best proceed methodically and write the energetic equations. stant, the steady-state capacitor current iC must be zero.
Superposition works, when additional inputs are applied
Energetic Equations to the same input node, or “driving point.” Superposition
fails in this system, because the two voltage sources act on
Node Eqs: iB1 = iB2 + iR1 iR1 + iR2 = iC different input nodes.
We will now reduce the equation list to derive the system
Loop Eqs: v1g = v13 + v3 g v21 + v1g = v23 + v3 g equation for the voltage across the capacitor in configura-
dv3 g tion three. The only difference between this reduction and
Elements Eqs: v13 = iR1 R1 v23 = iR2 R2 iC = C our previous is that we now keep two inputs. We will write
dt
the inputs v1g and v21 as v1g (t ) and v21 (t ) to identify as inputs.
1
Energy Eqs: E System = EC = Cv32g
2 Reduction:  Inputs v1g (t ) and v21 (t ) Output v3g
Superimposing (or summing) the step responses of con-
figurations one and two yields a steady-state voltage across
v21 (t ) + v1g (t ) = v23 + v3 g → v21 (t ) + v1g (t ) = iR2 R2 + v3 g
the capacitor of v3 g = 4.5 VDC. This cannot be the equilib-
rium state. If the steady-state voltage in the capacitor were
v3 g = 4.5 VDC, then there would be current flowing from the (
v21 (t ) + v1g (t ) = iC − iR1 R2 + v3 g)
capacitor back through both resistances in steady-state
dv3 g
v21 (t ) + v1g (t ) = R2 C − iR1 R2 + v3 g
dt
v1g = v13 + v3 g → v1g − v3 g = v13

v1g − v3 g dv3 g v 
v1g − v3 g = iR1 R1 → = iR1 v21 (t ) + v1g (t ) = R2 C −  13  R2 + v3 g
R1 dt  R1 

dv3 g
v1g − v3 g
=
1.5 VDC − 4.5 VDC
= iR1 → iR1 =
−3.0 VDC v21 (t ) + v1g (t ) = R2 C
dt
(
− v1g (t ) − v3 g ) RR
2
+ v3 g
R1 R1 R1 1

R  dv3 g  R2 
v21 (t ) + v1g (t ) +  2  v1g (t ) = R2 C + v3 g + v3 g
and  R1  dt  R1 

 R  dv3 g  R2 
v21 + v1g = v23 + v3 g → v21 + v1g − v3 g = v23 v21 (t ) + 1 + 2  v1g (t ) = R2 C + 1 +  v3 g
 R1  dt  R1 
v21 + v1g − v3 g
v21 + v1g − v3 g = iR2 R2 → = iR2
R2
 R + R2  dv3 g  R1 + R2 
v21 (t ) +  1  v1g (t ) = R2 C + v3 g
 R1  dt  R1 
372 6  Power Transmission, Transformation, and Conversion

2 R2 2 R2
+ Switch 2
i R2 + Switch 2
i R2
1.5 VDC iB 2 1.5 VDC iB 2
- R1 - R1
1 3 1 3
+ Switch 1
i R1
Switch 1
i R1
1.5 VDC iB iC(ss)=0
1 C Fig. 6.105   Steady-state current loop of the system in configuration
- three

Fig. 6.104   No current flows through the capacitance during steady- iR1 ( ss ) + iR 2 ( ss ) = iC ( ss ) → iR1 ( ss ) = −iR 2 ( ss )
0
state of the step response when the system is in configuration three
A current loop through the two resistors and battery B2 is cre-
 R1   R1   R1 + R2  ated in steady-state, Fig. 6.104.
 R + R  v21 (t ) +  R + R   R  v1g (t )
1 2 1 2 1 The applied voltage drops over each resistance, in propor-
tion to its resistance divided by the total, because the resis-
 R R  dv3 g  R1   R1 + R2 
= 1 2 C + v3g tors have the same current. We will calculate voltage v31 ( ss )
 R1 + R2  dt  R1 + R2   R1  across resistor R1 from node 3 to node 1, and add it to voltage
v1g ( ss ) across battery B1 to determine the voltage across ca-
The System Equation is: pacitor v3 g ( ss ) in steady-state

 R1   R1 R2  dv3 g v21 ( ss ) = v23 ( ss ) − v13 ( ss )


 R + R  v21 (t ) + v1g (t ) =  R + R  C dt + v3 g
v21 ( ss ) = iR2 ( ss ) R 2 − iR 1 ( ss ) R1
1 2 1 2

τ

 RR  We have established that in steady-state iR1 ( ss ) = −iR2 ( ss ).


τ = 1 2 C Hence
 R1 + R2 
v21 ( ss ) = −iR1 ( ss ) R2 − iR1 ( ss ) R1
 R1 
 R + R  v21 ( ss ) + v1g ( ss ) = v3 g ( ss )
1 2 v21 ( ss ) = −iR1 ( ss ) ( R2 + R1 )

The resistance in the time constant has the same form as the v21 ( ss )
equivalent resistance of two resistors in parallel = −iR1 ( ss )
R2 + R1
R1 R 2 1 Therefore,
Requiv = =
R1 + R 2 1 1
+
R1 R 2 v13 = iR1 R1 → − v13 = v31 = −iR1 R1

The circumstances are similar but not identical to the equiv- R1


alent resistance, which replaces two resistors connected v31 = v21 ( ss )
R 2 + R1
between the same two nodes. In the present case, the two
resistors have only node 3, where they connect to the capaci-
tor, in common. Nevertheless, we can conclude that the time When we add this result to the voltage of battery B1, v1g, it
constant of configuration three is less than the time constants agrees with steady-state value of the step response we de-
of either configurations one or two, because the resistance to rived from the system equation
two parallel resistances is less than that of either individual
resistance.  R1 
The steady-state voltage across the capacitor can be un-  R + R  v21 ( ss ) + v1g ( ss ) = v3 g ( ss )
1 2
derstood, if one recognizes that absence of current into the
capacitor at node 3 yields
6.6  Floating Sources, Transformers, and Transducers 373

3.0 LA LB
2.5
Configuration 2 F(t)
2.0
M
Pivot
v3g (t) Configuration 3
1.5
VDC
1.0
Configuration 1 K b
0.5

0.0
0 5 10 15 20 25 30
Fig. 6.107   Schematic of a lever with its pivot on a shock mount
t, sec
Fig. 6.106   Step responses for the system in Configurations 1, 2, and 3 g
LA LB

The step responses for the system in configurations one, F(t)


two, and three, using R1 = 1, R2 = 2, and C = 3, are shown in M
Pivot
Fig. 6.106.
1 2 3

6.6.4  Floating Transformers and Transducers K b

So far, we have modeled transformers and transducers which g


were referenced to ground. In the general case, transformers
and transducers can be located anywhere within a dynamic Fig. 6.108   Schematic of a lever with pivot on a shock mount showing
system. A transformer or transducer which is not referenced nodes of distinct translational velocities
directly to ground is known as “floating.” The transformation
or transduction relationship is not affected. The across vari- LA
v32 = - __ v
able differences are those across each branch in the positive. LB 12
Regardless of where in a linear a transducer or transformer 1 3
interface is located in the system, the positive direction of
the through variables on both interface branches (sides) of
the interface must be in the direction of the drop in the across Fα Fβ
variable; otherwise, you both create power as you transform
or transduce it, violating energy conservation. F(t) 2 M
Whether in parallel in a translational mechanical system,
or coaxial in a rotational mechanical system, as illustrated as
a motor mount above, a spring and dashpot are used in many K b
different physical configurations and a wide range of capaci-
ties and applications in machine design. In some cases, the
device is intended simply to provide the compliance neces- g
sary to accommodate misaligned parts, which occur due to
Fig. 6.109   Linear graph of a lever with pivot on a shock mount
the practical limits of tolerances of parts and assemblies. If
the primary purpose were to limit the magnitude of impulse
loadings, then the device would be named a shock mount, a 2. Draw and label the nodes.
shock coupling (for rotation), or, if the forces or torques are 3. Add both branches (sides) of transformer or transducer
very large, a “snubber.” elements.
The mechanical schematic in Fig. 6.107 shows the pivot 4. Add the remaining elements, Fig. 6.109.
of a lever mounted on a spring and damper in parallel. The In this case, because the transformer is floating, we must
linear graph for this system is drawn following the usual check after we add all of the remaining elements that the as-
method. Again: sumed directions for the through variables in the transformer
1. Find the nodes of the across variables on the schematic, branches is in the direction of a drop in the across variable.
Fig. 6.108.
374 6  Power Transmission, Transformation, and Conversion

+x,v L2
b1 V2g = _____ V1g
L 1+ L 2
F(t) 1
g 1 2

L1 K
2 M F(t) b1 Fα Fβ b2 M K
L2
b2
g
g
Fig. 6.111   Linear graph of the mechanical system shown in Fig. 6.110
Fig. 6.110   Schematic of a translational mechanical system with a lever
showing nodes of distinct value of the across variable, translational ve-
locity +x,v

F(t)
g 1

6.7 Equivalent Elements in Systems with L1


Transformers and Transducers
2
L2 b2
We often wish to simplify a machine (or the dynamic model
of a machine) by replacing both a mechanical transformer or
transducer (gears, lever, or linkage) and an energy storage or g
dissipation element with a single “equivalent” element. We
must size the equivalent element, so that it will store (or dis- Fig. 6.112   Auxiliary system constructed by removing elements from
the original system of Fig. 6.110
sipate) the same amount of energy (or power) as the mechan-
ical transformer plus the element it replaces. To determine
+x,v
the equivalent element, we will use an “auxiliary,” simpli-
fied system, consisting of just the source driving the original F(t) 1
g
system, the transformer or transducer, and the energetic ele-
ment. This auxiliary system will be reduced to an equivalent b2 equiv
system, consisting of the source and an equivalent energetic
element. The method is best presented using an example.

6.7.1 Equivalent Elements in a System with a


Transformer
Fig. 6.113   Equivalent system created by removing lever and changing
the parameter value of damping b2
A translational mechanical system consisting of a force
source, a lever, two dampers, a spring, and a mass is shown
in Fig. 6.110 and its linear graph is Fig. 6.111. The auxiliary system is created by removing all of the
We wish to eliminate the lever but maintain the dynamic energetic elements from the original system, except the force
relationship between applied force F( t) and the velocity of source, the lever, and damper b2, Fig. 6.112. The desired
node 1, the point of application of force F( t). Consequently, equivalent damper is shown in an equivalent system consist-
we need to determine the equivalent mass, spring constant, ing of just the force source acting directly on the equivalent
and damper coefficient b2, so that when the lever is elimi- damper b2equiv , Fig. 6.113.
nated, and the elements are connected between node 1 and For damper b2equiv to be equivalent to damper b2 in the orig-
ground, they provide energetic properties equivalent to the inal system, the effect of the leverage acting on damper b2
original system. We will use an “auxiliary” system consist- must be included in the parameter value b2equiv. This is done by
ing of the force source, the lever, and one of the dynamic drawing the linear graph for the auxiliary system, shown in
elements connected between node 2 and ground in order to Fig. 6.114a. Extract the energetic equations, and reduce them
determine the equivalent element parameter. We will first de- to equate the force Fα acting through the transformer branch
termine the equivalent element for damper b2. on the node on one side of the interface, to a force written
6.7  Equivalent Elements in Systems with Transformers and Transducers 375

a L2 b a L2 b L2
V2g = _____ V1g V2g = _____ V1g V2g = _____ V1g
L 1+ L 2 L 1+ L 2 L 1+ L 2
1 2 1 1 2 1 2

F(t) Fα Fβ b2 F(t) b2 equiv F(t) Fα Fβ K F(t) Fα Fβ M

g g g g

Fig. 6.114  a Linear graph of the auxiliary system consisting of the Fig. 6.115  a Auxiliary system to determine the equivalent spring.
force source, the lever, and damper b2. b Linear graph of the equivalent b Auxiliary system to determine the equivalent mass
force source acting directly on the equivalent element b2

We have eliminated the unwanted power variables. Now, re-


in terms of the original damping coefficient b2, and the lever arrange the equation into the form of the elemental equation
ratio. This effectively removes node 2 from the system, yield- of a damper
ing the system shown in Fig. 6.114b.
If the auxiliary and equivalent systems have the same Fα = n 2 b2 v1g
dynamic properties, then the power flows from the forces
sources in the two systems must be identical. Consequently, The equivalent damping coefficient is
the power flows from node 1 to ground in both systems are
equal:  2
 L2 
b2equiv 2
= n b2 =  b2 (6.20)
Fα v1g = Fb2 v1g  L1 + L2 
equiv

This is consistent with the objective of the equivalent sys- Check the physical reasonableness of this result. A larger
tem to have the same “feel” as the original system, since it damping coefficient yields a larger force acting through a
requires damper, which is needed for a unit velocity drop across a
damper. Our equivalent damper will be connected between
Fα = Fb2 node 1 and ground, and will have a larger velocity drop than
equiv
the original damper, which was connected between node 2
We will write the energetic equations for the auxiliary sys- and ground. In order to have the same force from the source,
tem (with a single energy storage or dissipation element). the equivalent damping coefficient must be smaller than the
Then, reduce either the element equation or energy storage original damping coefficient, which is our result.
equation, so that it is written in terms of the power variables, The analysis is repeated for the spring and the mass,
Fα and v1g. The term in that equation comprises of the ele- Fig. 6.115.
ment parameter and the transformer ratio, n, is the equivalent
L2 1
parameter. Transformer Eqs: v2 g = v1g ≡ nv1g Fβ = − Fα
L1 + L2 n
 L2  1
Transformer Eqs: v2 g =   v1g ≡ nv1g Fβ = − Fα . Node Eqs: F (t ) = Fα − Fβ − FK = 0
 L1 + L 2  n
Node Eqs: F (t ) = Fα − Fβ − Fb2 = 0 Loop Eqs: (both trivial) v1g = v1g v2 g = v2 g
dFK dv2 g
Loop Eqs: (both trivial) v1g = v1g v2 g = v2 g Element Eqs: = Kv2 g FM = M
dt dt
FK2
Element Eqs: Fb2 = b2 v2 g Energy Eqs: E Auxillary = E K E K =
system 2K
1
Reduction  Start with the element equation for the damper, and E Auxillary = E M E M = Mv22g
since there is no energy equation. Substitute to eliminate Fb2 System 2
and v2 g The reduction for the equivalent spring will be demonstrated
twice: first, starting with the energy equation and next, start-
1
Fb2 = b2 v2 g → − Fβ = b2 nv1g → Fα = b2 nv1g ing with the element equation.
n
376 6  Power Transmission, Transformation, and Conversion

Reduction starting with the energy equation for a spring 1

(− F )
2
F2 β
EK = K → EK =
2K 2K F(t) b1 b2 Mequiv Kequiv
2
equiv
1 
 Fα  Fα2
n
EK = → EK =
2K (
2 n2 K ) g

Fig. 6.116   Equivalent system


yielding

2 a 1 b 1
2  L2 
K equiv =n K = K
 L1 + L2 

Reduction starting with the element equation for a spring F(t) b1 b2 F(t) b total
equiv

1 
dFK (
d − Fβ ) d  Fα 
n  g g
dt
= Kv2 g →
dt
= K nv1g → (
dt
)
= nKv1g
Fig. 6.117  a Auxiliary system of the force source and dampers b1 and
b2 equiv. b Equivalent system of force source and damper btotal
dFα
dt
(
= n 2 K v1g )
Auxiliary system:
also yields
2
Node Eq: F (t ) = Fb1 + Fb2
 2  L2  equiv
K equiv =n K = K (6.21)
 L1 + L2  Loop Eq: (trivial) v1g = v1g

The physical reasonableness check of this result parallels Element Eqs: Fb1 = b1v1g   Fb2 = n 2 b2 v1g
the damper. Moving the spring up on the lever reduces the equiv

mechanical advantage of the lever on the spring. The spring Equivalent system:
must be softer, if it is higher on the lever for it to “feel” the
same as the force source. Node Eq: F (t ) = Ftotal
The reduction for the equivalent mass will start with the
energy equation for the mass Loop Eq: (trivial) v1g = v1g

1 1 Element Eq: Ftotal = btotal v1g


( )
2
EM = Mv22g → E M = M nv1g
2 2
Reduction
yielding
F (t ) = Fb1 + Fb2 → F (t ) = b1v1g + n 2 b2 v1g
 2 equiv
 L2 
M equiv = n 2 M =  M (6.22)
 L1 + L2  (
F (t ) = b1 + n 2 b2 v1g )
The last step is to draw the linear graph of the system with and
equivalent elements on node 1, Fig. 6.116, then combine any Ftotal = btotal v1g → F (t ) = btotal v1g
like elements. One further simplification is possible. The two
dampers in parallel can be combined, Fig. 6.117. Therefore,

(
btotal = b1 + n 2 b2 )
6.8  Example Problems 377

1 velocity source, Ω( t), which acts on a compliant shaft K,


shown schematically as a torsion spring. Shaft K drives a
pinion with N2 teeth, engaged in a rack with N1 teeth per inch.
The rack is connected to mass M, which slides on a lubricant
F(t) b1+b2 Mequiv Kequiv film with damping b2. Gear N3 engaged in the rack drives
equiv
the input shaft of a fluid coupling (drag cup) with damping
b1. The output shaft of the fluid coupling drives flywheel J.
g
Example Problem One  Draw a linear graph of this system
Fig. 6.118   Simplified equivalent spring, mass, and damper system and determine the transformer or transducer equations.
The key step is to identify the nodes of the across vari-
ables in the system. This system consists of three subsys-
Summarizing the equivalent elements, Fig. 6.118, tems: rotational one, translational, and rotational two. The
across variables are rotational and translation velocities,
2
Fig. 6.120.
2  L2  There are two complementary perspectives on the power
K equiv =n K = K
 L1 + L2  flow through this “multiport” transducer. The first is to see
the rack as both transducing and dividing the power flowing
2 into the rack from the pinion. This perspective is presented
 L2 
M equiv = n 2 M =  M in the linear graph shown in Fig. 6.121.
 L1 + L2  Note that the rack acts as a transducer in the interface be-
tween nodes 2 and 3, and as an idler in a transformer in the
2
2  L2  interface between nodes 2 and 4.
btotal = b1 + n b2 = b1 +  b2 In the second perspective, the rack is seen as transducing
 L1 + L2 
power across the interface between nodes 2 and 3, distribut-
ing some of the power to mass M and damping b2, and trans-
ducing the remaining power across the interface between
6.8  Example Problems nodes 3 and 4, as shown in the linear graph of Fig. 6.122.
Both linear graphs are correct.
6.8.1 Example Problem 1: Linear Graph Derive the transducer and/or transformer equations. The an-
of a Hybrid Rotational Translational gular velocities of the pinion and the gear are proportional to
System the translational velocity of the rack. The pinion and gear are
described by their number of teeth. The rack is described by
A hybrid rotational and translational mechanical system is its lineal pitch, or the number of teeth per inch, which must be
shown schematically in Fig. 6.119. The input is an angular converted to teeth per meter:

Fig. 6.119   A hybrid rotational Angular Velocity Source, Ω(t)


and translational mechanical sys- Compliant Shaft
tem Torsional Spring K

Pinion, N2 teeth

Rack, N1 teeth per inch

Flywheel Gear, N3 teeth


Rotational Inertia J
Mass M

Lubricant Film
Damping b2

Fluid Coupling
Damping b1
378 6  Power Transmission, Transformation, and Conversion

Fig. 6.120   Nodes of distinct Angular Velocity Source, Ω(t)


values of the across variables, Compliant Shaft
rotational and translational ve- Torsional Spring K
locities, shown on the schematic
of the hybrid mechanical system g
Pinion, N2 teeth
1
Rack, N1 teeth per inch
2
Flywheel Gear, N3 teeth
Rotational Inertia J
5 3 Mass M

4
Lubricant Film
Damping b2

Fluid Coupling
g
g Damping b1

Fig. 6.121   Linear graph show- b1


ing the rack dividing power be-
K
1 2 v3g= XΩ 2g 3 Ω 4g= YΩ2g 4 5
tween the subsystems with mass
M and mass moment of inertia J

Ω(t) TR FR M b2 Tα Tβ J

Fig. 6.122   Linear graph show- K b1


ing the power being transferred 1 2 v3g= XΩ 2g 3 Ω4g = Z v2g 4 5
to the rack, and the remainder
flowing into the subsystem with
the mass moment of inertia J

Ω(t) TR FR M b2 Fα Tβ J

 teeth   inch  teeth  meter   N 2 teeth 


N1  = 39.37 N1 v3 g =   Ω2 g
 inch   0.0254 meter  meter 
 39.37 N1 teeth   2π rad 

N 2 meter
The transducer constant relating the angular velocity of the v3 g = Ω2 g
N1 (39.37 )( 2π rad )
pinion in radians per second to the translational velocity of
the rack in meters per second is based on equal speeds of N 2 meter
their pitch surfaces: v3 g = X Ω 2 g where X =
N1 (39.37 )( 2π rad )
 N 2 teeth   teeth 
 2π rad  Ω 2 g =  39.37 N1 meter  v3 g
6.8  Example Problems 379

Calculate the transducer equation, which relates the torque


Input Force, F(t)
of the pinion to the force acting on the rack:
L1
PA + PB = 0 → TR Ω 2 g + FR v3 g = 0
Spring K1 L2 pivot
TR Ω 2 g + FR X Ω 2 g = 0 → TR = − X FR
L3
Damper b
Likewise, the transducer constant relating the translational
velocity of the rack to the angular velocity of the gear is
based on equal pitch surface speeds:
Spring K 2

 N 3 teeth   teeth  Fig. 6.123   Translational mechanical system. Force source acting on a
 2π rad  Ω4 g =  39.37 N1 meter  v3 g pivoted beam

 39.37 N1 teeth   2π rad  Table 6.2   Parameter values


Ω4 g =     v3g L1 L2 L3 b K1 K2
 meter  N 3 teeth 
N · sec N N
5 m 2 m 3 m 400 10 20
N (39.37 )( 2π rad ) m m m
Ω 4 g = Z v3 g where Z = 1
N 3 meter

The corresponding transducer equation relating the force of


the rack to the torque on the gear is The transformer equation relating the pinion and gear torques
is
PC + PD = 0 → Fα v3 g + Tβ Ω 4 g = 0 PE + PD = 0 → Tα Ω 2 g + Tβ Ω 4 g = 0

Fα v3 g + Tβ Z v3g = 0 → Fα = − Z Tβ Tα Ω 2 g + Tβ Y Ω 2 g = 0 → Tα = −YTβ

The transformer equation, which relates the pinion’s angular


velocity directly to the gear’s velocity, results from eliminat-
ing the rack’s velocity from the proceeding transducer equa- 6.8.2 Example Problem 2: A Pivoted Beam
tions: (Lever) Acting on Three Elements

A translational system is shown in Fig. 6.123. The system


v3 g = X Ω 2 g and Ω 4 g = Z v3 g → Ω 4 g = ZX Ω 2 g
consists of a rigid, massless beam supported by a pivot and
acted on by input force F( t), springs K1 and K2, and damper
N1 (39.37 ) ( 2π rad ) b. The parameter values and the spacing of elements along
N 2 meter
Ω4 g = Ω2 g the beam are given in Table 6.2. The system is in steady-state,
N 3 meter N1 (39.37 ) ( 2π rad ) after having been subjected to a force input F (t ) = 50 N , at
some time t < 0, before the pulse input plotted in Fig. 6.124
N2
Ω4 g = Ω2 g is applied to the system.
N3
Example Problem Two, Part A  Derive the system equation
N2 that relates input force F( t) to the velocity of the location,
Ω 4 g = Y Ω 2 g where Y = where spring K2 is attached to the beam.
N3
Identify the nodes of distinct values of the across variable,
translational velocity, on Fig. 6.125.
380 6  Power Transmission, Transformation, and Conversion

F(t)
200
Po
w
newtons
150 F(t)
er
100
50 Pspring K1 Psource
Spring K1 Pdamper
0 1 2 3
t, sec Damper b Pspring K2
Fig. 6.124   Force applied to the system shown in Fig. 6.123
Spring K2
g
Fig. 6.126   Power from force source viewed as flowing into the lever,
Input Force, F(t)
and then dividing between the two springs and the damper

1 L1
L2
____ ____
-L 3
v 2g = v1g v 3g = v1g
Spring K1 L2 2 L1 +L 2
1
L1 +L 2
3
2 pivot
g L3
Damper b
g 3 b Fβ Fα F(t) K1 Fγ Fδ K2

Spring K 2
g g
g g
Fig. 6.125   Nodes of distinct values of the across variable, translational
velocity Fig. 6.127   Linear graph showing power dividing where it enters the
beam, at node 1

Three equivalent linear graphs are shown. The first two


have identical “topology,” or connectedness, and represent ____
-L 3
L2 v 3g = v1g
power from the source dividing between spring K1, at node 1 v 2g =
____ v1g 2 L1 +L 2 3
1, damper b, at node 2, and spring K2 at node 3, Figs. 6.126, L1 +L 2

6.127 and 6.128. The division of power is represented by the


product of the velocity of node 1 times the continuity equa-
tion for node 1: F(t) K1 Fα Fβ b Fγ Fδ K2

(
F (t ) v1g = FK1 + Fα + Fγ v1g )
g g g g
F (t ) v1g = FK1 v1g + Fα v1g + Fγ v1g
Fig. 6.128   Linear graph with same topology of Fig. 6.127

where Fα v1g and Fγ v1g are the power flows to the damper and
spring K2, respectively. Geometry: There is one angular velocity of the rigid
An alternative linear graph divides the power from the beam. All points on the beam have the same angular velocity
source at node 1 between spring K1 and node 2. The power but different translational velocities
transmitted to node 2 then divides between the damper and
node 2, Figs. 6.129 and 6.130. v1g v2 g −v3 g
Ωbeam = = =
Transformer Equations: Use geometry to derive the first L1 + L 2 L2 L3
transformer equation of each pair, then sum power flows
down the branches of the transformer interface to derive the L2 − L3
second transformer equation. v2 g = v1g and v3 g = v1g
L1 + L2 L1 + L2
6.8  Example Problems 381

er Name the lever ratios, A and B, to reduce effort and the likeli-
Pow hood of transcription error:
F(t) L2
Psource A≡
L1 + L2
v2 g = Av1g Fα = − AFβ
Po
Pspring K1 we L3
Pnode 2 r B≡
L1 + L2
v3 g = − Bv1g Fγ = BFδ .
Spring K1
Pnode 3 FK21 FK22
Energy Eqs: E sys = E K1 + E K2 E K1 = EK = .
Damper b 2 K1 2
2K2
Pspring K 2 Reduction:  Input F( t), Output v3g

Spring K2 F (t ) = FK1 + Fα + Fγ → F (t ) = FK1 − A Fβ + B Fδ

Fig. 6.129   The alternative is to view power from the source flowing F (t ) = FK1 + A Fb − B FK2
into node 1, the remaining power to node 2, and then what remains
flowing to node 3
dF (t ) dFK1 dv2 g dFK2
= + Ab −B
dt dt dt dt
Summation of transformer power flows
dF (t ) dv2 g
Pα + Pβ = 0 → Fα v1g + Fβ v2 g = 0 = K1v1g + Ab − B K 2 v3 g
dt dt

L2 − L2
Fα v1g + Fβ v1g = 0 → Fα = Fβ dF (t ) K1 dv1g
L1 + L2 L1 + L2 =− v3 g + A2 b − B K 2 v3 g
dt B dt

Pγ + Pγ = 0 → Fγ v1g + Fδ v3 g = 0δ dF (t ) 1 2 dv3 g  K1 
=− A b − + B K 2  v3 g
dt B dt  B 

− L3 L3
Fγ v1g + Fδ v1g = 0 → Fγ = Fδ
dF (t ) dv3 g
L1 + L2 L1 + L2 −B
dt
= A2 b
dt
(
+ K1 + B 2 K 2 v3 g )
The derivation of the system equation for the velocity of the
beam, at the point where spring K2 is attached, will use the B dF (t ) A2 b dv3 g
linear graph of Fig. 6.127. − 2
= 2
+ v3 g
K1 + B K 2 dt K1 + B K 2 dt
The system has two springs but only one independent en-
ergy storage mode. Given the deformation of one spring, the
deformation of the other can be calculated, since the beam is Substitute in the lever ratios
rigid. The system equation will be first-order.
2
 L3   L2 
Energetic Equations b
 L1 + L2  dF (t )  L1 + L2  dv3 g
− 2
= 2
+ v3 g
Continuity Eqs:  L3  dt  L3  dt
K1 +  K2 K1 +  K2
F (t ) = FK1 + Fα + Fγ − Fb − Fβ = 0 − FK2 − Fδ = 0  L1 + L2  
 L1 + L2 

Compatibility Eqs: v1g = v1g v2 g = v2 g v3 g = v3 g


dFK1 dFK2 Check the units of the system equation in terms of the power
Element Eqs: Fb = bv2 g = K1v1g = K 2 v3 g . variables and time. Transformer constants are dimensionless.
dt dt The units of the element parameters are
L2 − L2
v2 g = v1g Fα = Fβ
L1 + L2 L1 + L2 F F
[ K ] =  t v  [b] =  v 
− L3 L3    
v3 g = v1g Fγ = Fδ
L1 + L2 L1 + L2
382 6  Power Transmission, Transformation, and Conversion

v 2g =
L2
____ v1g v 3g =
____
-L 3
v1g
g
L1 +L 2 L1 +L 2 Input Force, F(t)
1 2 3

1
L1+L2
F(t) K1 Fα Fβ b Fγ Fδ K2 Spring K1 pivot
g L3

g g g 3
Fig. 6.130   The linear graph corresponding to the power flows shown
in Fig. 6.129 Spring K 2
g
Fig. 6.131   Auxiliary system for determining equivalent spring
  L3  
  L + L  
 1 2 dF (t ) 
− 2
dt  ____
-L 3
 K +  L3  K  v 3g =
L1 +L 2
v1g
 1
 L1 + L2 
2
 1 3

  L 2 
  2
 b 
  L1 + L2  dv3 g  F(t) K1 Fγ
= + v3 g  Fδ K2
 L 
2
dt   
K + 3
K 
 1  L1 + L2  2 
g g
 1 F   b v t v F   F t v v 
 K t  =  K t  + [ v ] →  F t  =  v F t  + [ v ]
Fig. 6.132   Linear graph of the auxiliary system shown in Fig. 6.131

t v F   F t v v
 =  + [v] → [v] = [v ] + [v ] The spring constant for the equivalent of spring K2 attached
F t   v F t to node 1 is
2

The units check.  L3 


K equiv = B 2 K 2 → K equiv =  K2
There are two springs in this system. Should the system  L1 + L2 
equation be second order? No. This is a first-order system,
because the deformations and forces of the two springs are The last step is to remember how springs in parallel combine,
related through the rigid beam. The state variable of the sys- Fig. 6.133. Is the equivalent spring constant stiffer or softer
tem is established by finding the single equivalent spring, than one of the original spring constants? It is stiffer, because
which stores the same energy as the two springs in the sys- the two springs divide the load leading to less deformation.
tem. We will find the single equivalent spring for node 1. The Hence, the spring constants of springs in parallel sum.
auxiliary system is shown in Fig. 6.131 and its linear graph 2
in Fig. 6.132.  L3 
K total = K1 +  K2
Express the energy equation for spring K2 in terms of  L1 + L2 
force Fγ to eliminate node 3 from the auxiliary system
Example Problem Two, Part B  Determine the velocity of
FK22 Fδ 2
the location where spring K2 is attached to the beam, as a
EK = → EK = function of time in response to the input plotted in Fig. 6.124.
2
2K2 2
2K2
Use Mathcad or MATLAB to plot that velocity.
2 Use superposition of the unit step response to formulate
 Fγ 
the response function. The system is assumed to be de-en-
 B  Fγ2
EK = → EK = ergized, prior to the application of the first step input at an
2
2K2 2
2B2 K2 unknown time but long enough prior to time t = 0, that the
6.8  Example Problems 383

a 1 b 1 L3
L1 + L 2
( )
v3 g 0+ = − 2
1N
 L2 
F(t) K1 F(t) b 
K equiv K total  L1 + L 2 

( )
2
L 3 L1 + L 2
g g v3 g 0( )= −+

b L 22
1N

Fig. 6.133  a Linear graph of the auxiliary system with the equivalent
spring in parallel with spring K1. b Equivalent spring combined with Determine the final value of the unit step response, and the
spring K1 time constant from the system equation.

 L3 
system has reached steady-state by time t = 0. We will as-  L + L  dus ( ss )
1 2
sume that the first step input was applied at time t =  − 7 τ, − 2
 L3  dt 0
seven time constants before the pulse is applied. K1 +  K2
Identify the type of step response from the initial and  L1 + L2 
steady-state values. To use superposition we must assume 2
 L2 
that the system is de-energized, prior to the input of the  L + L  b dv3 g ( ss )
unit step. Determine the value of the output variable at time = 1 2
2
+ v3 g ( ss )
t = 0+ , the instant after a unit step input was applied to the  L3  dt 0
K1 +  K2
de-energized system.  L1 + L2 
Since the system is assumed to be de-energized before
the application of the unit step input, both spring forces are v3 g ( ss ) = 0
zero. The forces acting through the springs cannot change
instantaneously, because springs store energy. The applied 2
unit step must be picked up by the damper at time t = 0+. The  L3   L2 
transient response of the system is due to the transfer of the  L + L  dF (t )  L + L  b dv3 g
1 2 1 2
force from the damper to the two springs. − 2
= 2
+ v3 g
 L3  dt  L3  dt
Calculate the force acting through damper b in order to K1 +  K2 K1 +  K2
determine the velocity of node 2 at time t = 0+ , v2 g (0+ ). Find  L1 + L2  
 L + L2 
1
v3 g (0+ ) from v2 g (0+ ). τ
2
 L2 
( ) = 1N = F (0 )
F 0 +
K1
+
0
( )
+
+ Fα 0 + Fγ 0 ( )
+
 L + L  b
1 2
τ= 2
 L3 
( )
+
1N = Fα 0 + Fγ 0 ( )
+
→ 1N = − AFβ 0 + BFδ 0 ( ) +
( ) +
K1 + 
 L1 + L2 
K2

( )
1 N = AFb 0+ − B FK2 0+ ( ) 0
1 N = AFb 0+ ( ) Notice that the denominator of the time constant is the com-
bined spring constant, Ktotal.
1 The unit step response is a decay to zero of the form
( )
Fb 0+ =
A
1N
−t

( )
v3 g (t ) = v3 g 0+ e τ
1 1
( )
bv2 g 0+ =
A
1N → ( )
v1g 0+ = −
bA 2
1N
Evaluate the constants, v3g(0+) and τ,
B
( )
v3 g 0+ = −
bA 2
1N
3m (5 m + 2 m )
2
m
( )
v3 g 0+ = −
N ⋅ sec
1 N = 0.0919
sec
u .s.
400 ( 2 m )2
m
384 6  Power Transmission, Transformation, and Conversion

− (t + (7 )( 2.39))
200 F(t)  m 
newtons v3 g (t ) = 50  0.0919 e 2.39  us (t + ( 7 )( 2.39))
150 100 N u s (t-1)  sec 
50 N u s(t+7τ)
100 − (t −1)
 m 2.39 
50 + 100  0.0919
sec
e  us (t − 1)
 
− (t − 2 )
 m 2.39 
-7τ 0 1 2
t, sec
3 − 150  0.0919
sec
e  us (t − 2)
-50  
-100

-150 6.8.3 Example Problem 3: A Serpentine Belt


Driving Two Elements
-200 -150 N u s(t-2)
A rotational mechanical system is shown in Fig. 6.136. The
Fig. 6.134   Superposition of scaled and time shifted Heaviside step in-
puts to create the input applied to the system
system consists of a motor which is modeled as a torque
source, T( t). The motor drives a serpentine belt through a
4-in. diameter pulley. Driven by the belt are a rotation-
al damper (drag cup), b = 2N · m · sec, and a flywheel,
8
J = 8kg · m 2. A 5-in. diameter pulley drives the shaft attached
to the rotational damper’s drag. The shaft on the rotational
4
v3g(t) damper’s cup is rigidly attached to the machine’s frame. The
m 0 flywheel is driven by a pulley, 6-in. in diameter. There are
___
sec three idler pulleys to route the belt, and provide sufficient
wrap of the belt around the motor, damper, and flywheel pul-
-4 leys. Assume that the belt does not slip.

-8 Example Problem 3, Part A  Derive the system equation


-20 -15 -10 -5 0 5 10 15
t, sec which relates the input torque to angular velocity of the
motor shaft. Check the units of the system equation in terms
Fig. 6.135   Response function v3g( t) of the power variables and time.
The first step is to find the nodes of the across variable
2
on either end of the energetic elements and identify them on
 2m  N ⋅ sec the schematic, Fig. 6.137. The idler pulleys route the belt but
 5 m + 2 m  400 m have no energetic properties and should be neglected.
τ= 2
→ τ = 2.39 sec There are two equivalent linear graphs for this system
N  3m  N
10 +   20 which differ on how one perceives the power flow. First, the
m  5m + 2 m m
idler pulleys have no effect on the power flow in the system.
They can be ignored. Second, the belt has no mechanical
The unit step response is properties, other than being described as not slipping on the
pulleys. No mechanical properties means the belt does not
−t
m 2.39 store or dissipate energy. It simply transmits power. An ideal
v3 g (t ) = 0.0919 e
sec belt is perfectly flexible but inextensible. The following are
the two equivalent models.
Create the input function by superimposing scaled and time- Model One. Ignore the ideal belt and view power as di-
shifted Heaviside step functions, Figs. 6.124 and 6.134. viding at the torque source into two power flows. One power
flow occurs between the motor shaft and the damper shaft.
F (t ) = 50 N us (t + 7τ ) + 100 N us (t − 1) − 150 N us (t − 2) The other power flow occurs between the motor shaft and
the flywheel shaft, Fig. 6.138.
Construct the response function, by weighting (scaling) the Model Two. Include the ideal belt, and view the power
unit step response with the magnitude of the input. Time- from the torque source as flowing into the belt, and then
shift the unit step response with the same time shift as the dividing into two power flows at the damper shaft, one
input term. Multiply by the unit step with that time shift to which flows into the damper and the other into the flywheel,
zero-out the term, until time has advanced to that term: Fig. 6.139.
6.8  Example Problems 385

Fig. 6.136   Rotational system Flywheel


driven by a serpentine belt Rotational Inertia J

Rotational Damper
Damping b
Note! End of Damper
Shaft Rigidly Attached
to Frame.

Drive Pulley D6
6 in. Diameter Motor Modeled as
Torque Source T(t)

Drive Pulley D 5
5 in. Diameter

Idler Pulleys D3
3 in. Diameter
Drive Pulley D4
4 in. Diameter

Fig. 6.137   Nodes of distinct Flywheel


values of the across variable Rotational Inertia J
angular velocity

Rotational Damper
Damping b
Note! End of Damper
Shaft Rigidly Attached
3 to Frame.
g
Drive Pulley D6
6 in. Diameter Motor Modeled as
g 2 Torque Source T(t)

Drive Pulley D 5
5 in. Diameter

Idler Pulleys D3
3 in. Diameter
Drive Pulley D4
1
4 in. Diameter

Fig. 6.138  a Model One. Power


from the source divides the
a b
motor shaft. b Linear graph of D D
__
Model One
Ω 2g = __4 Ω 1g Ω 3g = 4 Ω 1g
D5 D6
2 1 3

Pflywheel b Tβ Tα T(t) Tγ Tδ J
Pdamper
g g g
Psource
386 6  Power Transmission, Transformation, and Conversion

Fig. 6.139  a Model Two. Power


delivered to the belt and divided
a b D
D
at the damper shaft. b Linear Ω 2g = __4 Ω 1g Ω 3g = __5 Ω 2g
D5 D6
graph of Model Two
1 2 3

Pflywheel T(t) Tε Tφ b Tη Tλ J

Pdamper
g g g
Pbelt = Psource

We will work the problem using the linear graph for Energetic Equations
Model One, Fig. 6.138.
Nodes: T = Tα + Tγ − Tβ − Tb = 0 − Tδ − TJ = 0
Transformer Equations The transformer equations are
derived by using the geometry of the system to establish Loops: Ω 1g = Ω 1g Ω 2g = Ω 2g Ω 3g = Ω 3g
one of the two transformer equations for each interface. The
second transformer equation for each interface is established d Ω 3g
Elements: Tb = bΩ 2 g TJ = J
by summing the power flows down the two branches of the dt
interface and equating the sum to zero. The geometric prop-
D5 D D6 D
erty of the system used to establish the transformer ratios is Ω1 g = Ω 2 g Tβ = − 5 Tα Ω1 g = Ω 3 g Tδ = − 6 Tγ
the belt speed, vbelt. An ideal belt does not slip on a pulley. D4 D4 D4 D4
Therefore, the surface of the pulley and the belt have the
same speed. The angular velocity of the shaft is belt speed Define temporary variable names for the pulley diameter
divided by the radius of the pulley. Hence, the geometry ratios:
yields the angular velocity relationship. Then the angular D5 D6
velocity relationship and the power summation yield the A≡ and B ≡
D4 D4
relationship between the torques.
Ω 1g = AΩ 2 g Tβ = − ATα Ω 1g = BΩ 3 g Tδ = − BTγ
D4 D D
Geometry: Ω1g = vbelt = 5 Ω 2 g → Ω1g = 5 Ω 2 g
2 2 D4
1
Energy: E sys = E J EJ = J Ω 32g
D4 D D 2
Ω1g = vbelt = 6 Ω 3 g → Ω1g = 6 Ω 3 g.
2 2 D4 Reduction: Input T, Output Ω1g

Power: Tα Ω1g + Tβ Ω 2 g = 0 and Tγ Ω1g + Tδ Ω 3 g = 0 1 1 1 1


T = Tα + Tγ → T=− Tβ − Tδ → T= Tb + TJ
Torques: A B A B
D5 1 1 d Ω 3g 1 1 1 1 d Ω1 g
Tα Ω 1g + Tβ Ω 2 g = 0 → Tα Ω 2 g + Tβ Ω 2 g = 0 T= bΩ 2 g + J → T= bΩ1g + J
D4 A B dt AA BB dt

D5 2 2
d Ω 1g
Tβ = − Tα  1  1
D4 T =   bΩ 1g +   J
 A  B dt
D6 D6 1 d Ω 1g 1
Tγ Ω 1g + Tδ Ω 3 g = 0 → Tγ Ω 3 g + Tδ Ω 3g = 0 T= J + 2 bΩ 1g
D4 D4 B2 dt A
D6 A2 A2 J d Ω1g
Tδ = − Tγ T= 2 + Ω1g System equation
D4 b B b dt

Substitute in pulley diameter ratios


6.8  Example Problems 387

200 of the de-energized system to a unit step input of torque,


T(t)
1 N-m. We will then construct the input function by scaling
N • m 150 and time shifting Heaviside step functions. Finally, we will
100 construct the response function by scaling and time shifting
the unit step response with the same scaling and time shifting
50
used to construct the input function. We must also multiply
each term in the response function by a Heaviside unit step
10 20 30
function with the time shift of the input, so that each term
-50 t, sec
will be zeroed-out until its corresponding input step turns on.
-100

-150
Unit Step Response  Determine the value of the output vari-
able Ω1g, at time t = 0+ . The energy stored in an element can-
Fig. 6.140   Input torque not change instantaneously without an infinite power flow.
Hence, the state variable must remain unchanged, when a
2 2 2 step input is applied to the system. The state variable of this
1  D5   D   D4  J d Ω 1g
T = 5   + Ω 1g first-order system is the angular velocity of the flywheel, Ω3g.
  
b  D4   D4   D6  b dt
1
E sys ( 0 − ) = 0 = E sys ( 0+ ) → E sys = J Ω 23 g
2 2 2
1  D5   D  J d Ω1 g
T =  5 + Ω1g System equation Ω 3 g ( 0− ) = 0 = Ω 3 g (0+ )
b  D4   D6  b dt

τ
2
We have assumed that the belt does not slip. Consequently,
 D5  J the angular velocity of all the pulleys is proportional, includ-
 D  b ≡ τ. Time constant ing the pulley on the torque source. If the flywheel’s pulley is
6
not rotating at time t = 0+ , then neither is the torque source:

Check the consistency of the units. Ω 3 g (0+ ) = 0 → Ω1g (0+ ) = 0

 D  2 1   D  2 J d Ω1g  We now determine the steady-state value of output variable


 5  T  =  5   + Ω1g  Ω1g in response to a unit step input from the system equation:
 D4  b   D6  b dt 

 D  J d Ω1g ( ss )
2 2
T   J Ω  1  D5 
 b  =  b t  + [ Ω ]   1 N ⋅ m us ( ss ) =  5  + Ω1g ( ss )
b  D4  1  D6  b dt 0

2
 Ω Ω T t Ω   5  1N·m  rad
T =  + [Ω ] → [Ω ] = [Ω ] + [Ω ] Ω1g ( ss ) =    → Ω1g ( ss ) = 0.78
 4   2 N · m · sec  sec
 T  T Ω t  u .s. u .s.

The units are consistent. 2 2


 D  J  5 8
τ =  5 =  = 2.78 sec
Example Problem 3, Part B The system is running in  D6  b  6  2
steady-state under the action of a step input of torque,
T = −100 N · m, applied at an unknown previous time The first-order step response is a stable growth from an initial
when, at time t = 0, the applied torque is increased to value, Ω 1g (0+ ) = 0, of the form
T = +50 N · m . Ten seconds later, the torque is increased
again to T = +200 N · m. The input torque history is shown  −t

Ω 1g (t ) = Ω1g ( ss ) 1 − e τ 
in the plot of Fig. 6.140. Determine the angular velocity  
of the motor’s shaft as a function of time. Use Mathcad or
MATLAB to plot the angular velocity of the motor’s shaft rad  −t

Ω 1g (t ) = 0.78 1 − e 2.78
from 0 to t = 10 + 6 τ sec. u .s. sec  
We will determine the response of the system to this input
using superposition. First, we will establish the response
388 6  Power Transmission, Transformation, and Conversion

200 150 N •m us (t) Ideal Frictionless


T(t) Fluid Coupling
Bearing
N • m 150 Damping b N1
N3
100
150 N m us (t-10)

50 Angular
Velocity
Input Ω(t) N2
-7τ 10 20 30 N4
-50 t, sec Gear 2 and Gear 3 are keyed
Inertia J
-100 (fixed) to the same shaft.

-150 -100 N us (t+7τ) Fig. 6.143   Rotational system with compound (cluster) gears. The
input is angular velocity
Fig. 6.141   Torque input
− (t +19.5)
rad  
Ω 1g (t ) = −78 1− e 2.78
 us (t + 19.5)
200 sec  
150 rad  −t

+ 117 1 − e  us (t )
2.78

sec 
Ω1g (t) 100 − (t −10)
rad  
 us (t − 10)
rad
___ 50 + 117 1− e 2.78

sec sec  
0

-50 6.8.4 Example Problem 4: Rotational System


-100 with Compound Gears
-20 -10 0 10 20 30
t, sec A rotational mechanical system is shown in Fig. 6.143.
The angular velocity source acts on the input shaft of
Fig. 6.142   Response function Ω1g , the angular velocity of the motor’s a fluid coupling, modeled as a drag cup with damping
shaft b = 150 N · m · sec/rad . The output shaft of the fluid coupling
drives gear one with N1 = 54 teeth. Gear one drives gear two
with N 2 = 18 teeth. Gear two is keyed (fixed) to the same
Construct the input function from scaled and time shifted shaft as gear three. Gear three with N 3 = 48 teeth drives gear
Heaviside unit step functions (Figs. 6.141 and 6.142): four with N 4 = 12 teeth, which is on the shaft of the mass
moment of inertia J = 0.3 kg · m 2. Inertia J is supported by
T (t ) = −100 N ⋅ m us (t + 7τ ) + 150 N ⋅ m us (t ) an ideal frictionless bearing.
+ 150 N ⋅ m us (t − 10)
Example Problem Four, Part A Use the linear graph
method to derive the system equation, which relates the
Construct the response function with the same scaling and angular velocity input to the torque through the fluid cou-
time shifting as the input function. Multiply each term of the pling, damping b.
response function by the corresponding unit step function Identify nodes of distinct values of the across variable,
from the input function angular velocity, on the schematic of the system, Fig. 6.144.
Note that gears two and three have the same angular veloc-
− (t + 7 τ )
 rad    ity, and the only energetic property of any of the gears is
Ω 1g (t ) = −100  0.78   1 − e τ
 us (t + 7τ )
the transformation of the power variables. Real gears have
 sec  
inertia, and gear sets have significant friction. In this model,
 rad   −t
 the equivalent inertia of the gear set would be included in ro-
+ 150  0.78   1 − e τ  us (t )
 sec    tational inertia J, and the linear friction of the gear set would
− (t −10) be added to the damping of fluid coupling b.
 rad   
+ 150  0.78  1− e  us (t − 10) The most common error is to assign two energetic prop-
τ
 sec    erties to a branch of a transformer or transducer. Avoid this
6.8  Example Problems 389

Ideal Frictionless using geometric compatibility, where number of teeth on a


Bearing
Fluid Coupling gear is proportional to its pitch circumference
Damping b N1
N3 N1
N1Ω 2 g = vtooth = − N 2 Ω 3 g → Ω 3g = − Ω 2g
2 N2
Angular 1 3 4
Velocity g N
Input Ω(t) N2 Simplify notation. Define X ≡ 1 → Ω 3g = − X Ω 2 g
N4 Likewise N 2
g Gear 2 and Gear 3 are keyed
Inertia J
(fixed) to the same shaft. N3
N 4 Ω 4 g = vtooth = − N 3 Ω 3 g → Ω 4g = − Ω 3g
N4
Fig. 6.144   System schematic annotated to show nodes of distinct val-
ues of the across variable, angular velocity
and
N3
≡Y → Ω 4 g = −Y Ω 3 g
Ω3g= - X Ω 2g Ω4g= - Y Ω 3g N4
b
1 2 3 4
Sum the power flowing into the interface of the transformer
to establish the relationship between the torques

Ω(t) Tα TβTγ Tδ J PA + PB = 0 → Tα Ω 2 g + Tβ Ω 3 g = 0

Substitute to eliminate Ω 3g
g g g g
( )
Tα Ω 2 g + Tβ − X Ω 2 g = 0 → Tα = XTβ
Fig. 6.145   Linear graph including angular velocity of Gears 2 and 3, Ω3g
Similarly,

Ω4g= - Z Ω 2g PC + PD = 0 → Tγ Ω 3 g + Tδ Ω 4 g = 0
b
1 2 4
( )
Tγ Ω 3 g + Tδ −Y Ω 3 g = 0 → Tγ = YTδ

Ω(t) Tα Tδ J Energetic Equations

Continuity:

g g g TSource = Tb Tb = Tα − Tβ − Tγ = 0 − Tδ − TJ = 0
Fig. 6.146   Linear graph of equivalent system eliminating node 3
Compatibility:

error by drawing the interface elements and identifying their Ω 1g = Ω 12 + Ω 2 g Ω 3g = Ω 3g Ω 4g = Ω 4g


through variables, before drawing any other elements. A
branch identified by a through variable is less likely to have d Ω 4g N1 N3
a parameter assigned to it. Name the torques acting through Elements: Tb = bΩ 12 TJ = J ≡X ≡Y
dt N2 N4
the gear set transformer elements as you please, but keep it
simple, Fig. 6.145. Ω 3 g = − X Ω 2 g Tα = XTβ Ω 4 g = −Y Ω 3 g Tγ = YTδ
Note that there is no energetic element connected to
1
node 3, other than the branches of two transformer inter- Energy: E sys = E J EJ = J Ω32g
faces. Consequently, this linear graph can be simplified, by 2
eliminating node 3, and expressing the velocity of node 4 as Reduction: Input Ω (t ) = Ω1g ≡ Ω , Output Tb
a function of the velocity of node 2, Fig. 6.146.
Tb T 1
We will work with the linear graph, which includes Ω = Ω12 + Ω 2 g → Ω = + Ω 2 g → Ω = b − Ω3 g
node 3. Write the transformer equations for the gear set, b b X
390 6  Power Transmission, Transformation, and Conversion

Tb 1  1  d Ω 1 dTb 1 d Ω4 g Determine the unit step response. First establish the value
Ω= −  − Ω4 g  → = +
of the state variable at time t = 0+, assuming the system is
b X Y dt b dt XY dt
initially de-energized:
d Ω 1 dTb 1 TJ d Ω 1 dTb 1
= + → = − Tδ
dt b dt XY J dt b dt XYJ
E sys ( 0− ) = 0 = E sys ( 0+ ) = E J ( 0+ )
d Ω 1 dTb 1  Tγ  d Ω 1 dTb 1
= − → = + Tβ 1
dt b dt XYJ  Y  dt b dt XY 2 J E J ( 0+ ) = ( )
J Ω 42g 0+ = 0 → Ω 4 g 0+ = 0 ( )
2
d Ω 1 dTb
= +
1  Tα 

d Ω 1 dTb
=
1
+ 2 2 Tb Ω 4 g ( 0+ ) = 0
 
dt b dt XY 2 J  X  dt b dt X Y J

dΩ J dTb The unit step is Ω (0+ ) = 1(rad/sec) us (0+ ) = 1(rad/sec) . Use


2 2
X Y J = X 2Y 2 + Tb the values of the input and the state variable at time t = 0+ ,
dt b dt
with algebraic equations of the equation list, to determine the
2 2 2 2 value of the output variable at time t = 0+ :
 N1   N 3  d Ω  N1   N 3  J dTb
 N   N  J dt =  N   N  b dt + Tb
2 4 2 4
rad
2 2 Ω ( 0+ ) = 1 ( )
= Ω 12 0+ + Ω 2 g 0+ ( )
 N1 N 3  d Ω  N1 N 3  J dTb sec
 N N  J dt =  N N  b dt + Tb . System equation
 rad T b 0
+
1( )
( )
2 4 2 4

τ 1 = − Ω 3 g 0+
sec b X
Check the units of the system equation in terms of the power
variables and time: rad Tb 0
+
( )1  rad 
1
sec
=
b
+
XY
Ω 4 g 0+ ( ) 0
( )
→ Tb 0+ = b 1 
 sec 
T dΩ Tt
T = bΩ → [b] =  Ω  T=J → [ J ] =  Ω   N · m · sec   rad 
  dt   ( )
Tb 0+ = 150
  1 
rad   sec 
( )
→ Tb 0+ = 150 N · m

 N N  2 d Ω   N N  2 J dT 
 1 3  J  =  1 3  b
 + [Tb ]
 N N  dt  N N  b dt Find the time constant and the steady-state value of the unit
 2 4   2 4  step response from the system equation:
 Ω J T 
 J t  =  b t  + [T ]
2 2
 N1 N 3  d Ω  N1 N 3  J dTb
 N N  J dt =  N N  b dt + Tb
2 4 2 4

T t Ω  T t Ω T 
 =  + [T ] → [T ] = [T ] + [T ] τ
Ω t  Ω T t  2
NN  J
The units check. τ = 1 3
 N2 N4  b
Example Problem Four, Part B  The system was de-ener-
2
gized before the angular velocity pulse, shown in Fig. 6.147,  54·48  0.3 kg · m 2
τ=  = 0.288 sec
was applied to the system. Determine the response function  18·12  N · m · sec
150
for torque through the fluid coupling (damping b), for the rad
input plotted in Fig. 6.147.
2
We will solve using superposition of scaled and time-  N1 N 3  d  rad 
shifted unit step responses. Construct the input function from  N N  J dt 1 sec us ( ss )
2 4 0
scaled and time shifted Heaviside unit steps, Fig. 6.148:
 N N  J dTb ( ss )
2

= 1 3 + Tb ( ss )
rad rad  N2 N4  b dt 0
Ω (t ) = 20 us (t ) − 20 us (t − 0.5)
sec sec
Tb ( ss ) = 0
6.8  Example Problems 391

Identify the type of first-order unit step response, Tb ( ss ) = 0. TM = α iM → PA + PB = 0 → v2 g iM + TM Ω 3 g = 0


The unit step response is a decay to zero of the form
−t −t

( )
Tb (t ) = Tb 0+ e τ → Tb (t ) = 150 N · m e 0.288
v2 g iM + α iM Ω3 g = 0 → Ω3 g =
−1
v2 g
α

Construct the response function, Fig. 6.149, with the same Power Screw  We know that the lineal pitch of the power
scaling and time shifting as the input function, screw is 2 threads per inch. One thread equals 2π rad of rota-
tion:
rad rad
Ω (t ) = 20 us (t ) − 20 us (t − 0.5)
sec sec threads  1rev   2π rad   39.36 in  rad
2       = 495
inch 1thread rev 1meter m
−t
Tb (t ) = (150 N ⋅ m ) e 0.288 20 us (t )
rad Ω 3g
− (t − 0.5)
β = 495 and β≡ → Ω 3 g = β v4 g
− (150 N ⋅ m ) e 0.288
20 us (t − 0.5) m v4 g

PC + PD = 0 → TS Ω 3 g + FN v4 g = 0
6.8.5 Example Problem 5: Hybrid Electric,
Rotational, and Translational System
−1
TS β v4 g + FN v4 g = 0 → TS = FN
The system shown in Fig. 6.150 is powered by a voltage β
source which drives an electric motor. The electric motor’s
winding resistance is R = 8 Ω . The mass moment of inertia Energetic Equations
of the motor is negligible compared to the motor’s load. The
motor produces 150 ft · lbs of torque for 30 amps of current. Continuity:
The shaft of the motor turns a flywheel with mass moment
of inertia J = 3 kg · m 2 The flywheel’s shaft turns a power isource = iR iR = iM − TM − TJ − TS = 0 − FN − FM − Fb = 0
screw with a linear pitch of two threads per inch. The power
screw pushes mass, M = 20 kg, on lubricated ways. The lu- Compatibility: v1g = v12 + v2 g Ω 3g = Ω 3g v4 g = v4 g
bricating film between mass M and the ways has damping
b = 2 N · sec/m . Elements:

Example Problem Five, Part A Use the linear graph d Ω 3g dv4 g


iR = Rv12 TJ = J FM = M Fb = bv4 g
method to derive the system equation, which relates input dt dt
voltage to the velocity of the mass.
Draw the linear graph. The essential first step is to identify
−1 −1
the nodes of distinct values of the across variable in the en- TM = α iM Ω 3g = v2 g TS = FN Ω 3 g = β v4 g
α β
ergetic model of the system, Fig. 6.151. There are three sub-
systems: electrical, rotational, and translational, Fig. 6.152.
1 1
The energetic properties of the system are described in the Energy: E Sys = E J + E M EJ = J Ω32g EM = Mv42g
problem statement and called-out on the drawing. Do not 2 2
add additional energetic properties to the model. Reduction: Input v1g, Output v4g

Electric Motor  We know the motor produces 150 ft · lbs of v1g ≡ v = v12 + v2 g → v = RiR − αΩ3 g
torque for 30 amps of current. Convert the torque to SI
R
 12 in   0.0254 m   1 N  v = RiR − αβ v4 g → v= TM − αβ v4 g
150 ft · lb  = 204 N · m α
 1 ft   1 in   0.224 lb 
R R R
Calculate transducer constant
v=
α
( −TJ − TS ) − αβv4 g → v=−
α
TJ −
α
TS − αβ v4 g

TM 204 N · m N·m
α≡ → α= = 6.8
iM 30A A
392 6  Power Transmission, Transformation, and Conversion

RJ β dv4 g R Ω(t) 20
v=−
α dt
+
αβ
( − FM − Fb ) − αβv4 g
rad
___
sec
RJ β dv4 g R  dv4 g  10
v=− +  −M − bv4 g  − αβv4 g
α dt αβ  dt 
0
0.25 0.5 0.75 1.0
RJ β dv4 g RM dv4 g Rb
v=− − − v − αβv4 g t, sec
α dt αβ dt αβ 4 g
Fig. 6.147   Input pulse of angular velocity
 RJ β RM  dv4 g  Rb 
−v =  +  + + αβ v4 g
 α αβ  dt  αβ 
rad
___ u (t)
20 sec s
 RJ β 2
RM  dv4 g  Rb α β  2 2 20
−v =  + + + v
 αβ αβ  dt  αβ αβ  4 g
10

 RJ β + RM  dv4 g  Rb + α β 
2 2 2
Ω(t)
−v =   dt +   v4 g
 αβ αβ ___ 0
rad 0.25 0.5 0.75 1.0
sec t, sec
-10
αβ  αβ   RJ β + RM  dv4 g 2
rad
___
− v= 2 2   + v4 g -20 sec us (t-0.5)
Rb + α β
2 2
 Rb + α β   αβ  dt -20

αβ RJ β 2 + RM dv4 g Fig. 6.148   Two scaled step functions, one of which is time shifted, that
− v= + v4 g superpose to form the input pulse
Rb + α β
2 2
Rb + α β
2 2
 dt
τ
System equation in time constant form 3,000
2,000
Check the units of the system equation:
1,000
Tb(t)
T
TM = α iM → [α ] =  i  N •m
0
  -1,000

−1 volt  -2,000
Ω3 g = v2 g → [α ] = 
α  Ω  -3,000
0 0.5 1.0 1.5 2.0
t, sec
−1 F
TS =
β
FN → [β ] =  T 
  Fig. 6.149   Response function, the torque acting through the damper


Ω3 g = β v4 g → [β ] =  v  DC Motor Flywheel
  Rotational Inertia J

volt 
v12 = RiR → [ R ] =  Electrical
 i  Resistance R Power Screw
Mass M
d Ω 3g Tt
TJ = J
dt
→ [ J ] =  Ω 
 

dv4 g Ft
FM = M → [ M ] = 
dt  v  Voltage Source v(t) Lubricating Film
Damping b
F
Fb = bv4 g → [b] =  v  Fig. 6.150   Electromechanical system
 
6.8  Example Problems 393

DC Motor Flywheel R TM = α i M Ω 3g= βv4g


Rotational Inertia J
1 2 3 4

Electrical
Resistance R 2 Power Screw
v(t) iM TM J TS FN M b
3 Mass M
1
g g g g
g 4
Fig. 6.152   Linear graph of the system shown in Figs. 6.150 and 6.151
Voltage Source v(t) Lubricating Film
Damping b g

Fig. 6.151   Electromechanical system, annotated to show nodes of 200


distinct values of the across variables, voltage, angular velocity, and
translational velocity 100
v(t) 0
VDC 1 2 3 4
 αβ   RJ β 2 + RM dv4 g  t, sec
-100
 − Rb + α 2 β 2 v =   + v4 g 
   Rb + α β
2 2
dt   
-200

Fig. 6.153   Input voltage pulse v( t)


 αβ   RJ β 2 + RM v 
 Rb + α 2 β 2 v =   + [v]
  Rb + α β t 
2 2

200 VDC us(t)
200
  200 VDC us(t-3)
  100
 αβ
volt  v(t) 0
 volt F T volt F Ω 
 i v + i Ω  VDC 1 2 3 4 5
 T v  t, sec
-100
 volt T t 2 volt F t 
 β + 
= i Ω i v v + v -200
 volt F T volt F Ω t 
[]
 i v + i  -300
-400 VDC us(t-2)
 Ω T v 
-400

 T F   volt T t F Ω volt F t  Fig. 6.154   Three scaled and time shifted step functions, which super-
   +  pose to form the input pulse
i T i Ω T v i v v
 
volt =  + [v]
 volt F   volt F t
   
 i v   i v  Input Function:
[v ] = + [v ] + [v ]
v (t ) = 200VDCus (t ) − 400VDCus (t − 2) + 200VDCus (t − 3)

The units check. Determine the unit step response of output variable v4g, as-
suming that the system is de-energized before application of
Example Problem Five, Part B  The system is de-energized the unit step input. The output variable, the velocity of the
at time t = 0 − . Solve the system equation for the voltage input mass, v4g, and Ω3g, the angular velocity of the flywheel, are
history given in Fig. 6.153. Use Mathcad or MATLAB to dependent variables. The mass cannot move, if the flywheel
plot the velocity of the mass from time t = 0 to t = 6 τ sec. does not move. The value of one of the two variables de-
Use superposition to construct input function v( t) from termines the value of the other. Thus, either variable can be
scaled and time shifted Heaviside unit step functions, used as the state variable of the system. For this problem,
Fig. 6.154.
394 6  Power Transmission, Transformation, and Conversion

it is convenient to use the velocity of the mass as the state 5.0


variable:
2.5
1 1 v4g(t)
E sys = E J + E M → E sys = J Ω 32g + Mv42g
2 2 0
cm
___
1 1 1 sec
( ) ( )
2
E sys = J β v3 g + Mv42g → E sys = J β 2 + M v42g -2.5
2 2 2
-5.0
The system is de-energized at time t = 0 −. Hence 0 1 2 3 4 5 6
t, sec
( )
v4 g 0 − = 0 = v4 g 0+ ( ) Fig. 6.155   Response function v4g( t)

Determine the system’s time constant and the steady-state


response of the output variable from the system equation: The response is a growth from zero of the form

αβ RJ β 2 + RM dv4 g  −t

− v= + v4 g v4 g (t ) = v4 g ( ss ) 1 − e τ 
Rb + α β
2 2
Rb + α 2 β 2 dt  
  
τ
m  −t

v4 g (t ) = −2.97 × 10 −4 1 − e 0.52
RJ β + RM sec  
2
τ=
Rb + α 2 β 2
Form the response function by scaling and time shifting the
2 unit step response function with the same scaling and time
rad 
(8 Ω) (3kg ⋅ m 2 )  495  + (8 Ω )( 20 kg ) shifting as the input function (Fig. 6.155):
m
τ= 2 2
= 0.52sec
N ⋅ sec   N ⋅ m  rad 
(8 Ω)  2  +  6.8   495   m −t

m A m v4 g (t ) = 200 us (t )  −2.97 × 10 −4  1 − e 0.52
 
sec   
The steady-state value of the unit step response − (t − 2 )
 m 
− 400 us (t − 2)  −2.97 × 10 −4   1 − e 0.52 
 sec   
αβ
− 1VDC us ( ss ) − (t − 3)
Rb + α 2 β 2 1
 m 
+ 200 us (t − 3)  −2.97 × 10 −4   1 − e 0.52

RJ β 2 + RM dv4 g ( ss )  sec   
= + v4 g ( ss )
Rb + α 2 β 2 dt 0

αβ  m −t

v4 g ( ss ) = − 1VDC v4 g (t ) =  −59.4 × 10 −3   1 − e 0.52
 us (t )
Rb + α 2 β 2  sec  
− (t − 2 )
 m 
−  −118.8 × 10 −3 1 − e  us (t − 2)
0.52
 N ⋅ m  rad    
−  6.8   495  sec   
 A  m
v4 g ( ss ) = 1VDC − (t − 3)
N ⋅ sec   N ⋅ m 
2
rad 
2
 m 
(8 Ω)  2 +  −59.4 × 10−3 1 − e  us (t − 3)
0.52
 +  6.8   495    
m A m sec   

m
v4 g ( ss ) = −2.97 × 10 −4 6.8.6 Example Problem 6: Hybrid Rotational
sec
and Fluid System

Identify and formulate the unit step response A torque source T( t) drives the input shaft of the belt drive
system shown in Fig. 6.156. The belt drive consists of a 6 in.
m diameter pulley on the input shaft, a 5 in. diameter pulley
v4 g (0+ ) = 0 and v4 g ( ss ) = −2.97 × 10 −4
sec on the shaft of a rotational inertia J = 1kg · m 2, and a 4 in.
6.8  Example Problems 395

Fig. 6.156   Belt-driven hybrid Inertia J


rotational mechanical fluid
Fluid reservoir vented
system
to the atmosphere
5 in. Dia. Pump
Pulley Low pressure return

2 in. Dia. Idler Hose, Fluid


Resistance R

Input Torque T(t) High


6 in. Dia. pressure
Pulley discharge
2 in. Dia. 4 in. Dia.
Idler Pulley

Fig. 6.157   System schematic Inertia J


annotated with nodes of distinct
Fluid reservoir vented
values of the across variables,
angular velocity and pressure to the atmosphere
5 in. Dia. Pump
Pulley Low pressure return
2
g Hose, Fluid
1
Resistance R
3 4
Input Torque T(t) High
g 6 in. Dia. pressure
Pulley discharge
4 in. Dia.
Pulley

diameter pulley on the shaft of a pump. The two other pul- Belt drive transformer equations  The belt’s surface speed,
leys are idlers. The pump is rated at 15 gallons per minute vs, is related to the angular speed of a pulley by the pulley’s
at 200 RPM. The pump pulls fluid from a reservoir vented radius:
to the atmosphere, and discharges it to a hose with fluid re-
D1
sistance R = 100 MPa · sec/m3. The flow returns to the res- vs = r1Ω 1g where r1 =
2
ervoir.
D1 D D
Example Problem Six, Part A  Use the linear graph method Ω 1g = vs = 2 Ω 2 g → Ω 2 g = 1 Ω 1g
2 2 D2
to derive the system equation, which relates the input torque
to the pressure of the pump’s discharge port, relative to
D1 D D1
atmospheric pressure, and check its units. Ω 1g = vs = 3 Ω 3 g → Ω 3g = Ω 1g
Find the nodes of distinct values of the across variables, 2 2 D3
angular velocity, and pressure on the system’s schematic,
Fig. 6.157. Sum the power flows of two branches of each interface to de-
Draw the linear graph, Fig. 6.158. Draw the transducer termine the second transformer equation for each interface:
interfaces first, labeling the branches of the interfaces with
through variables, so that the branches are not subsequently Pα + Pβ = 0 → Tα Ω 1g + Tβ Ω 2 g = 0
mistaken as other elements.
396 6  Power Transmission, Transformation, and Conversion

Ω2g = f Ω1g Ω3g = g Ω1g QP = h Ω 3g


rev  2π rad   1 min  rad
1 2 3 4 200 = 20.9
min  1 rev   60 sec  sec

The transducer ratio is


T(t) Tα Tβ J Tγ Tδ TP QP R
m3
Qp 9.48 × 10 −4
volume flow rate sec
= =
angular velocity Ω3 g rad
g 20.9
sec
Fig. 6.158   Linear graph of belt-driven system
Qp m3 m3
= 4.54 × 10 −5 → Q p = 4.54 × 10 −5 Ω3 g
Ω3 g rad rad
D1 D
Tα Ω 1g + Tβ Ω1g = 0 → Tβ = − 2 Tα m3
D2 D1 Define h ≡ 4.54 × 10 −5 → Qp = h Ω 3 g
rad
Pγ + Pδ = 0 → Tγ Ω1g + Tδ Ω3 g = 0
Sum the power flows to derive the second transducer equa-
D tion
D1
Tγ Ω1g + Tδ Ω1g = 0 → Tδ = − 3 Tγ
D3 D1
PA + PB = 0 → TP Ω 3 g + QP p4 g = 0

There is no need to convert the pulley’s diameters from inch-


es to meters, since the diameters only appear in ratios. TP Ω3 g + h Ω3g p4 g = 0 → TP = − h p4 g

6in. Energetic Equations


Ω 2g = Ω 1g → Ω 2 g = 1.2 Ω 1g
5in.
Continuity:
5in.
Tβ = − Tα → Tβ = −0.833Tα
6in. T (t ) = Tα + Tγ − Tβ − TJ = 0 − Tδ − TP = 0 − QP − QR = 0

6in. Compatibility:
Ω 3g = Ω 1g → Ω 3 g = 1.5Ω 1g
4in.
Ω 1g = Ω 1g Ω 2g = Ω 2g Ω 3g = Ω 3g p4 g = p4 g
4 in.
Tδ = − Tγ → Tδ = −0.667Tγ d Ω1g
6 in. Elements: TJ = J p4 g = RQR
dt
It is easier and less error prone to write the transformer equa- D1 6 in. 1
tions using single symbols in place of ratios or values: f ≡ = = 1.2 Ω 2 g = f Ω1g Tβ = − Tα
D2 5 in. f
D1 6 in. D1 6 in. D1 6 in. 1
f ≡ = = 1.2 and g≡ = = 1.5 g≡ = = 1.5 Ω 3 g = g Ω1g Tδ = − Tγ
D2 5 in. D3 4 in. D3 4 in. g

Pump transducer equations  The pump is rated at 15 gal- m3


h ≡ 4.54 × 10 −5 Q p = h Ω 3 g TP = − h p4 g
lons per minute at 200 RPM. The volume must be converted rad
to cubic meters and the angular velocity to radians per sec-
1
ond: Energy: E sys = E J EJ = J Ω 22g
2
 15gal   3.79 liters   1m   1min 
3
Reduction: Input T(t), Output p4g
  
min  1gal   1, 000 liters   60sec 
 
T (t ) = Tα + Tγ → T (t ) = − f Tβ + − g Tδ
m3
= 9.48 × 10 −4
sec
6.8  Example Problems 397

d Ω 2g
T (t ) = f TJ + g TP → T (t ) = f J − h g p4 g T(t) 900
dt N •m
d Ω1g 600
T (t ) = f 2 J − h g p4 g
dt 300
f J d Ω3 g
2
T (t ) = − h g p4 g
g dt
-1 0 1 2 3 4 5 6 7 8
f 2 J dQP t, sec
T (t ) = − − h g p4 g
g h dt
Fig. 6.159   Input torque
f J dp4 g
2
T (t ) = − − h g p4 g
g h R dt 300 N• m us (t+7τ) 600 N• m us (t-2)
1 f 2 J dp4 g 600
− T (t ) = 2 2 + p4 g
gh g h R dt 300

Check consistency of the units in terms of the power vari-


ables and time: -7τ 0 1 2 3 4 5 6 7 8
t, sec

dΩ Tt  p T(t) -300
T=J
dt
→ [ J ] =  Ω  p = RQ → [ R] =  Q  N•m
    -600
-900 N• m u s(t-6)
Q T 
[ h] =  Ω  =  p  -900
   
Fig. 6.160   Input torque function constructed of three superposed,
scaled, and time shifted step functions
Pulley ratios f and g are dimensionless:

 1   f 2 J dp4 g  Substitute in parameter values. Express MPa as Pa:


 − g h T ( t ) =  2 2  +  p4 g 
   g h R dt 
1
 Tt  − T (t )
 6in.   −5 m 
3
1   J p  p   Ω p  4in.  4.54 × 10 rad 
 h T  =  h 2 R t  + [ p ] →  T T  =  2 p t  + [ p ]
  h  2
 Q   6in.
(
 5in.  1kg ⋅ m
2
) dp4 g
 T t  = 2 2
+ p4 g
   6in.   −5 m  
3
6 Pa ⋅ sec 
dt
 p   Ω p  4in.  4.54 × 10 rad  100 × 10 m3 
 T =
   + [ p] → [ p] = [ p] + [ p]
T   Q T p t 
Ω p Q  kPa dp4 g
  −14.7 T (t ) = 3.11 + p4 g System Eq. in τ form
N⋅m dt
The units check:
Example Problem Six, Part B The system is run-
ning in steady-state under a previously applied torque of
1 f 2 J dp4 g
− T (t ) = 2 2 + p4 g 300 N · m , when the applied torque is increased to 900 N ⋅ m
gh g h R dt
at time t = 2, and then decreased to zero at time t = 6, as
2 shown in Fig. 6.159. Solve the system equation for the input
 D1 
1  D  Jdp4 g
torque history given and plot the pump’s discharge pressure
− T (t ) = 2
+ p4 g from time t = 0 to t = 6 + 6 τ sec.
D1  D1 
2
dt Create the input torque function by superposing three
h
 D  ( h ) R
2
D3 scaled and shifted step inputs, Fig. 6.160:
3
398 6  Power Transmission, Transformation, and Conversion

T (t ) = (300 N · m ) us (t + 6τ ) 2

+ ( 600 N ·m ) us (t − 2) − (900 N ·m ) us (t − 6) 0
-2

Calculate the unit step response of the system from its de-
p4g(t)
-4
energized state. From the system equation find the time MPa -6
constant and the steady-state response to a unit step input,
T (t ) = 1 N ⋅ m us (t ). -8
-10
1 f 2 J dp4 g -12
− T (t ) = 2 2 + p4 g -20 -10 0 10 20
gh g h R dt
 t, sec
τ
Fig. 6.161   Response function p4g( t). The transducer signs, which led
f2J to negative pressure, would be “fixed” before these results were used
τ= 2 2
g h R
2 The unit step response is a stable exponential growth from
 6in 
(
 5in  1kg ⋅ m
2
) zero of the form
τ= 2
 6in  
2  −t

−5 m   p4 g (t ) = p4 g ( ss ) 1 − e τ 
3
6 Pa ⋅ sec 
 4in   4.54 × 10 rad  100 × 10 m3   

τ = 3.11sec  t
 g 2 h2 R 
1 − 2
p4 g ( t ) = − T0 1 − e  f J  
g h  
1 f 2 J dp4 g ( ss )
− 1N ⋅ m us ( ss ) = 2 2 + p4 g ( ss )
gh 1 g h R dt 0
Substitute in parameter values and evaluate the constant
1
p4 g ( ss ) = − 1N ⋅ m terms of the unit step response
gh
1 N⋅m   t 
−
 3.11

1 p4 g (t ) = −  1 − e 
p4 g ( ss ) = −  m3 
1N ⋅ m
(1.5)  4.54 × 10−5   
u.s.
 m3 
(1.5)  4.54 × 10−5   rad 
 rad 

p4 g ( ss ) = 1, 200 Pa = 1.2 kPa (


p4 g (t ) = −14.7 kPa 1 − e −0.311t
u.s.
)

Determine the initial value of output variable, p4 g (0+ ) , in Create the response function, Fig. 6.161, by scaling and
response to a unit step input. The system is assumed to de- time shifting the unit step response function with the same
energized at time t = 0 − . The state variable is the angular scaling and time shifting as the input torque function, T( t):
velocity of the rotational inertia Ω2g. The value of the state
variable remains unchanged during the transition of the step ( )
p4 g (t ) = 300 ( −14.7 kPa ) 1 − e − 0.311(t + 6τ ) us (t + 6τ )
input, Ω 2 g (0 − ) = 0 = Ω 2 g (0+ ). Calculate the value of output
variable p4 g (0+ ) using the values of the state variable and (
+ 600 ( −14.7 kPa ) 1 − e − 0.311(t − 2) us (t − 2) )
the unit input torque at time t = 0+ .
− 900 ( −14.7 kPa ) 1 − e ( − 0.311(t − 6)
) u (t − 6)
s

( )
p4 g 0+ = RQR 0+ ( ) → p (0 ) = − RQ (0 ) + +
4g P
(
p4 g (t ) = −4.41MPa 1 − e − 0.311(t + 6τ ) us (t + 6τ ) )
p4 g (0 ) = − R h Ω (0 )
+
3g
+
( )
− 8.82 MPa 1 − e − 0.311(t − 2) us (t − 2)

( )
p4 g 0+ = − R h g Ω1g 0+ ( ) (
+ (13.2 MPa ) 1 − e − 0.311(t − 6) us (t − 6) )
Rhg
( )
p4 g 0 + = −
f
Ω 2 g 0+ ( ) 0
→ ( )
p4 g 0+ = 0
Summary 399

Fig. 6.162   Linear graph


a b N
transformer and transducer sign Ω 1g = - ___
2
Ω 2g
conventions for positive power N1
flow A + B = 0 1 2

PowerA Transformer Power B


or PowerA Power B
Transducer
A = TA Ω1g B = TB Ω 2g TA TB

Summary in the derivation of a system equation, since there is a one-


for-one power variable substitution. The power summation
Transformers and transducers interface subsystems with a equation is valuable but inefficient in a reduction, because
larger dynamic system. Transformers interface like subsys- the power variable substitution is one-for-three.
tems, and transducers interface dissimilar subsystems. Power It is often possible to derive a relationship between the
flows across the interface between the subsystems, but there displacements or velocities on the two sides of the interface,
is no direct connection between the through variables in the using the geometry of the mechanism of the volume dis-
two branches of the transducer or transformer interface. The placement of a pump or hydraulic motor. In other cases, such
power flowing down one branch of the interface, equal to the as with electric motors, generators, or turbines, experimental
product of the through variable and the across variable differ- data is needed to establish a transducer equation. Once a re-
ence of that branch, is equal to the product of the through and lationship is obtained, then substitution into the power sum-
across variables in the other branch. mation yields the second one-to-one equation, for example,
The sign convention is based on (a) assuming that the
power flows on both sides of the interface is positive and (b) N2
Ω 1g = − Ω 2g → PA + PB = 0
orienting the through variables in the two branches of the N1
interface towards ground. Because the positive direction of
power flow is into the transformer or transducer interface, Ω 1g TA + Ω 2 g TB = 0
and the interface cannot store power, at any instant, one of
the two power flows must be negative. Consequently, one
of the two equations, relating a power variable in a branch N2 N2
− Ω 2 g TA + Ω 2 g TB = 0 → TB = TA
on one side of the interface to a power variable on the other N1 N1
side, will have a negative sign. Although a negative sign
can be physically possible, as in the case of a pair of gears, When drawing linear graphs of systems with transformers
often, the negative sign is nonsense. The negative sign is and transducers:
a result of the linear graph sign convention. When there is 1. Find nodes of distinct values of the across variables for
a non-sensical negative sign, it should be tolerated, until each subsystem, and identify those nodes on the sche-
the derivation of the system equation is completed. Then matic.
fix the sign, if necessary, before using the results. Correct- 2. Draw and identify the nodes of the linear graph.
ing the sign prior to completion of the reduction leads to 3. Draw the two interface branches for each transducer or
further errors. Sign conflicts between subsystems are com- transformer, and name the through variable of each inter-
mon, because the subdisciplines of mechanical engineering face branch.
have different histories. 4. Add the remaining elements of the system between their
There are three equations associated with a transformer respective nodes.
or transducer. The first equation is the summation of power 5. There is no branch across a transformer or transducer in-
flow into the interface, Fig. 6.162a, which is the sum of the terface! There is no through variable flow across the inter-
products of the power variables in the two interface branch- face. Power flows across the interface.
es. The second and third equations each relate a single power A transformer constant is dimensionless. A transducer con-
variable on one side of the interface with a single power vari- stant has two sets of units. Generally, both sets must be used
able on the other side. The latter two equations are useful to check the units of a system equation with a transducer.
400 6  Power Transmission, Transformation, and Conversion

Fig. P6.1   DC motor driving an Flywheel Electric Motor


inertial load, JL Rotational Inertia JL Transduction KT
Rotational Inertia J M

Electrical Resistance R
Bearings
Damping b Voltage Source v(t)

Fig. P6.2   Hybrid fluid-transla-


tional mechanical system Fluid Power Unit
Pressure Source p(t)

Breather Cap (vent to atmosphere)

Hydraulic Line
Resistance R 1 Hydraulic Piston-Cylinder
Piston diameter D
Fluid Accumulator Damping b between Piston
Capacitance C and Cylinder
Hydraulic Line
Resistance R2 Press Die
Mass M

Hydraulic Line
Resistance R 3
Inertance I

Problems Problem 6.6 Draw the linear graph for the system shown in
Fig. P6.6 and derive the transformer or transducer constants
Problem 6.1 Draw the linear graph for the system shown in equations.
Fig. P6.1 and derive the transformer or transducer constants Problem 6.7 Draw the linear graph for the system shown in
equations. Fig. P6.7 and derive the transformer or transducer constants
Problem 6.2 Draw the linear graph for the system shown in equations.
Fig. P6.2 and derive the transformer or transducer constants Problem 6.8 Draw the linear graph for the system shown in
equations. Fig. P6.8 and derive the transformer or transducer constants
Problem 6.3 Draw the linear graph for the system shown in equations.
Fig. P6.3 and derive the transformer or transducer constants Problem 6.9 A rotational mechanical system is shown in
equations. the Fig. P6.9. An angular velocity source, Ω( t), acts on the
Problem 6.4 Draw the linear graph for the system shown in input shaft of a fluid coupling. The output shaft of the fluid
Fig. P6.4 and derive the transformer or transducer constants coupling drives a pinion with N1 teeth, which is engaged in
equations. a rack with mass M and N2 teeth per meter. The gear has
Problem 6.5 Draw the linear graph for the system shown in N3 teeth and inertia J2. The bearings on the input shaft col-
Fig. P6.5 and derive the transformer or transducer constants lectively have damping b1. The bearings on the output shaft
equations. collectively have damping b2.
Problems 401

Fig. P6.3   Rotational mechanical Flywheel


system Pulley Compliant Shaft
Rotational Inertia J1
Spring Constant K
Diameter D1
Pulleys Diameters Flywheel
D2 and D3 Rotational Inertia J 2

Angular Velocity
Input Ω(t)

Pulley Fluid Coupling


Diameter D4 Damping b

Fig. P6.4   Hybrid fluid-transla-


Fluid Power Unit
tional mechanical system modeled as a
Pressure Source Dashpot
Damping b
Hydraulic Line
Fluid Resistance R
L1

L2
Tank vented to
the atmosphere
Pivot
Hydraulic Cylinder
Diameter D
Linkage Spring K
Rotational Inertia J

Fig. P6.5   Rotational mechanical Bearing


system Damping b 3
Rotational
Inertia J

Compliant Shaft
Torsional Spring K 2

N4
N2

Compliant Shaft N3
Torsional Spring K 1 Bearing
N1 Damping b 2
Fluid Coupling
Damping b1

Angular Velocity
Source, Ω(t)
402 6  Power Transmission, Transformation, and Conversion

Fig. P6.6   Electromechanical Flywheel Electric Motor


system Rotational Inertia J L Gears 2 and 3 are keyed
Transduction K T
to the same shaft
Rotational Inertia J M
N4 N2

N1 Voltage Source v(t)

N3

Bearings
Damping b
Electrical Resistance R

Fig. P6.7   Translational me- Lubricating Film Mass M


chanical system Damping b 2
Spring K2

Spring K 1
Pivot
Dashpot b1 L3

L2

Force Source F(t)


L1

Fig. P6.8   Electromechanical L1 L2


system. The cross on the top Top Linkage
linkage is the centerline of the Solenoid Rotational Inertia J
solenoid’s magnetic force Coil Resistance R
Motor Constant α
Voltage Source
Dashpot
Damping b

Spring K

Frame

6.9.a Draw the linear graph of the existing system and Problem 6.10 A mechanical system consisting of a torque
determine the transformer equations. source and a rack and pinion is shown in Fig. P6.10. The
6.9.b Draw the equivalent linear graph, if the gear set is pinion has N1 = 8 teeth. The rack has N 2 = 4 teeth/in. The
eliminated, and inertias J1 and J2, the bearings b2, and torque source acts on a pinion with mass moment of inertia
the torsional shaft with spring constant K are replaced J = 0.05 kg · m 2 . The pinion is engaged with a rack of mass
by equivalent elements attached to the input shaft. M = 1 kg. The rack slides on a lubricating film with damping
6.9.c Calculate the equivalent elemental parameters for the b = 6 N ⋅ sec/m .
equivalent system of part b.
Problems 403

Fig. P6.9   Hybrid rotational-


Pinion
translational mechanical system
N1 teeth
Inertia J

Angular Velocity
Source, Ω(t)
Gear
N3 teeth
Fluid Coupling, Damper b1
Rack, Mass M, N2 teeth per meter
Viscous Lubricating Film, Damper b2

Fig. P.6.10   Hybrid rotational- Pinion


translational mechanical system Inertia J
N1 teeth
Rack
Mass M
N2 teeth per inch

Base Plate

Torque Source, T(t)

Viscous Lubricating Film


between Rack and
Base Plate, Damping b

6.10.a Derive the system equations for the following: motor torque is TM = KT iM , where KT = 8 N · m/A . The mo-
i. The velocity of the rack. tor’s mass moment of inertia is J M = 0.3 kg · m 2 . The motor
ii. The force acting to accelerate the rack. turns a flywheel with mass moment of inertia J L = 2 kg · m 2
iii. The force acting to shear the lubricating fluid and damping b = 0.1 N · m · sec/rad .
film. 6.11.a Derive the system equations for
Check their units. i. The current through resistance R.
The system is at rest at time t = 0 − , when a torque ii The voltage drop across resistance R.
T (t ) = 60N · m us (t ) is applied. iii. The back EMF.
6.10.b Solve the system equations. iv. The motor’s torque TM.
6.10.c Plot the responses from t = 0 to t = 6 τ using Mathcad v. The torque acting through damping b.
or MATLAB. vi. The torque acting to accelerate mass moment of
inertia JM + JL.
Problem 6.11 An electromechanical schematic of a DC vii. The angular velocity of mass moment of inertia
motor is shown in Fig. P6.11a. The motor’s resistance is JL .
R = 4 Ω . The relationship between the motor current and the and check their units.
404 6  Power Transmission, Transformation, and Conversion

Fig. P6.11  a Schematic of a DC


motor. b Voltage v( t) applied to
a Electric Motor b
the motor Transduction K T
100
Flywheel Inertia J M
Inertia JL v (t)
75
Resistance R VDC
50
Voltage
Source
v(t) t, sec
Bearings
Damping b 0.5 1.0
-25

-50

Fig. P6.12  a Translational


mechanical system. b Force input
a K b
F( t)
Lever

Pivot LA 40
F(t),N
30

20
LB 10
F(t)
0.5 1.0 1.5 2.0 2.5
LC b -10 t, sec

The motor was running in steady-state at time t = 0 under Problem 6.13 The rotational mechanical system shown in
the previously applied step input, when the pulse shown in Fig. P6.13a consists of a velocity source acting on the input
Fig. P6.11b was applied. shaft of rotational damper b1 = 100 N · m · sec (a drag cup).
6.11.b Solve the system equations. The output shaft of damper b1 drives a pinion, N1 = 12 teeth
6.11.c Plot the responses from time t = 0 to t = 6 τ sec meshed with two gears with N 2 = 24 and N 3 = 36 teeth.
using Mathcad or MATLAB. The gears, in turn, drive the input shaft of rotational damper
b2 = 100 N · m · sec and rotational inertia J = 25 kg · m 2 .
Problem 6.12 A mechanical system is shown in Fig. P6.12a. 6.13.a Derive the system equations for the following:
A force source acts on a lever with lengths LA = 1 m, i. The torque from the velocity source.
LB = 2 m, and LC = 1 m. Connected to the lever are a spring, ii. The torque acting to accelerate mass moment of
K = 80 N/m, and a damper, b = 4 N · sec/m . inertia J.
6.12.a Derive the system equations for the following: iii. The angular velocity of mass moment of inertia J.
i. The force acting through the spring. iv. The torque acting through damper b2.
ii. The force acting through the damper. Check their units.
iii. The velocity of the point of application of the The system was in steady-state under a previous step
input force. input of Ω( t) = 5 rad/sec before the angular velocity input,
Check their units. plotted in Fig. P6.13a, was applied.
The system is in steady-state under the action of a previ- 6.13.b Solve the system equations.
ously applied step input of 20 N, before the force input, F( t), 6.13.c Plot the responses from t = 0 to t = 1 + 6 τ sec using
plotted in Fig. P6.12b, is applied. Mathcad or MATLAB.
6.12.b Solve the system equations.
6.12.c Plot the responses from time t = 0 to t = 2 sec + 6τ Problem 6.14 A translational mechanical system is shown
using Mathcad or ­MATLAB. schematically in Fig. P6.14. The system consists of a
Problems 405

Fig. P6.13  a Rotational system.


b Angular velocity input
a b

Shaft welded to plate Ω(t)


Damper b2 rad 10
___
sec
Shaft rotates in Gear 5
ideal bearing N2 teeth

Pinion 0.5 1.0


N1 teeth t, sec
-5
Ω(t)
Inertia J Damper b1
Gear
N3 teeth Velocity source.
Positive rotation direction
defined by the right-hand rule

Fig. P6.14   Translational me- Lubricating Film


chanical system Mass M
Damping b2

Pivot Spring K
Dashpot b1 L3

L2

Force Source F(t)


L1

force source F (t ) acting on a pivoted beam with lengths The system was de-energized for time t < 0, before the
L1 = 36in., L 2 = 48in., and L 2 = 36 in. Attached to the input force F(t) = 1,000 N us (t) acted on the system.
beam are damper b1 = 200 N · sec/m , spring K = 4, 000 N/m, 6.14.b Solve the system equations.
and mass M = 20 kg . There is a lubricating film under mass 6.14.c Plot the responses from time t = 0 to t = 10 + 6 τ
M with damping b2 = 10 N · sec/m . using Mathcad or MATLAB.
6.14.a Derive the system equations for the following:
i. The force acting to accelerate mass M. Problem 6.15 A model of an engine lathe is shown in
ii. The force acting through the damper b1. Fig. P6.15a. The motor and gear box is modeled as a torque
iii. The force acting through the damper b2. source driving a rotational inertia J = 2.5kg · m 2. The motor
iv. The velocity of mass M. drives a lead (or power) screw, which is supported by bear-
v. The velocity of the point of application of the ings in the tail stock. The lead screw is modeled as having
input force. negligible torsional compliance. The lead screw bearings
Check their units. have damping b1 = 10 N · m · sec/rad . The lead screw pitch
is 2 threads per inch. The lead screw engages a lead nut
406 6  Power Transmission, Transformation, and Conversion

Fig. P6.15  a Schematic model


of an engine lathe. b Torque
a Load Carriage Lead Screw b
Lead Bearings with
input T( t) Screw
Lead Damping b1
Motor Nut
Inertia M 200
T(t)
150
T(t) J N •m
100

50

0
0 1 2
Lathe Way Lubricating Film t, sec
Damping b2

Fig. P.6.16   Hybrid rotational-


translational mechanical system
a Angular Velocity Source, Ω(t)
b
Compliant Shaft 80
Torsional Spring K Ω(t)
rad
___ 60
Pinion, N1 teeth sec 40

20
Rack, N2 teeth
per inch
Mass M 0.5 1.0 1.5 2.0 2.5
-20 t, sec

Lubricant Film
Damping b

attached to the load carriage. The lead nut moves the load spring constant K = 100 N · m · sec/rad . The shaft drives a
carriage with mass M = 150 kg along the lubricated ways of pinion with N1 = 10 teeth. The pinion is engaged in a rack.
the lathe. The lubricating film between the load carriage and The rack has N 2 = 4 teeth/in . The rack is attached to mass
the lathe’s ways has a damping coefficient b2 = 60 N · sec/m M = 10 kg. The rack and mass slide on a lubricating film with
6.15.a Derive the system equations for the following: damping b = 20 N · sec/m.
i. The torque acting to accelerate the motor’s mass 6.16.a Derive the system equations for the following:
moment of inertia J. i. The velocity of the rack.
ii. The angular velocity of the lead screw. ii. The force acting to accelerate the rack.
iii. The force acting to accelerate the load carriage iii. The force acting to shear the lubricating fluid
mass M. film.
iv. The velocity of the load carriage mass M. Check their units.
Check their units. The system is in steady-state under a step input of angular
The system was at rest before the input torque T( t), plot- velocity of − 40 rad/sec and time t = 0 − , when the angular
ted in Fig. P6.15b, was applied. velocity pulse shown in Fig. P6.16b is ­applied.
6.15.b Solve the system equations. 6.16.b Solve the system equations.
6.15.c Plot the responses from t = 0 to t = 3 + 6 τ using 6.16.c Plot the responses from t = 0 to t = 2 + 6τ using
Mathcad or MATLAB. Mathcad or MATLAB.

Problem 6.16 A mechanical system is shown in Fig. P6.16a. Problem 6.17 A fluid translational mechanical sys-
The angular velocity source drives a compliant shaft with tem is shown in Fig. P6.17a. The system consists of a
pump modeled as a pressure source, two fluid resistances
Problems 407

Fig. P6.17  a Fluid translational


mechanical system. b Pressure
a Hydraulic b
input Piston
Area A 5,000
C p(t) 4,000
psi
R1 3,000
p(t) M
R2
1,000

0 1.0 2.0
t, sec

Fig. P6.18  a Translational


mechanical system. b Input
a Force Source F(t)
b
force F( t) 400
Dashpot b 1
300

Pivot 200

Spring K1 100
L1 F(t)
0
N 0.5 1.0 1.5 2.0
Dashpot b 2
-100 t, sec
L2
-200

-300
L3
Spring K 2 -400

R1 = 10 MPa · sec/m3 and R2 = 100 MPa · sec/m3, an accu- 6.18.a Derive the system equations for the following:
mulator (fluid capacitor) C = 5.5 L/MPa, an hydraulic cylin- i. The force acting through spring K1.
der with a piston of diameter D = 6 in. and mass M = 2 kg . ii. The force acting through spring K2.
6.17.a Derive the system equations for the following: iii. The force acting through the damper b1.
i. The volume flow rate from the pump. iv. The force acting through the damper b2.
ii. The force acting to accelerate the piston. v. The velocity of the point of application of the
iii. The velocity of the piston. input force.
Check their units. Check their units.
The system was in steady-state under a previous step The system is at de-energized at time t = 0 when the sys-
input of 2,000 psi, when the input plotted in Fig. P6.17b was tem is subjected to input force F( t), plotted in Fig. P6.18b.
applied at time t = 0 . 6.18.b Solve the system equations.
6.17.b Solve the system equations. 6.18.c Plot the responses from time t = 0 to t = 2 + 6 τ
6.17.c Plot the responses from t = 0 to t = 6 τ using using Mathcad or MATLAB.
Mathcad or MATLAB.
Problem 6.19 A rotational mechanical system is shown
Problem 6.18 A translational system consisting of a force schematically in Fig. P6.19. The system is driven by an an-
source F( t), a lever, two springs, and two dashpots is shown gular velocity source which is connected to the input shaft,
in Fig. P6.18a. The parameter values are tabulated. by a fluid coupling with damping b1. The output shaft of the
fluid coupling connects to a compliant shaft with spring con-
Table P6.18   Parameter values stant K, which turns the pinion of a gear set. The pinion has
L1 L2 L3 b1 b2 K1 K2 N1 teeth. Gear N2 drives a shaft which rotates mass moment
N · sec N · sec N N of inertia J and is supported by a bearing with damping b2.
48 in. 60 in. 36 in. 100 125 175 200
m m m m
408 6  Power Transmission, Transformation, and Conversion

Fig. P6.19   Rotational mechani-


cal system Bearing Damping b 2
Pinion Inertia J
N1 teeth

Compliant Shaft
Spring Constant K

Fluid Coupling
Damping b1
Gear
N 2 teeth

Ω(t)
Angular velocity source

Fig. P6.20   DC motor driving an DC Motor


inertial load through a compliant Transduction KT
shaft

Bearings Damping b1
Voltage Source
Compliant Shaft v(t)
Spring Constant K

Resistance R

Rotational Inertia J
Bearings Damping b2

Table P6.19   Parameter values


b1 K N1 N2 J b2 bearings and brushes have damping b1 = 0.4 N · m · rad/sec.
N · m · sec N·m N · m · sec The relationship between the motor current and the motor
80 1,000 12 36 5 kg · m 2 0.2 torque is TM = KT iM , where KT = 8 N · m/A . The motor turns
rad rad rad
a compliant shaft with spring constant K = 500 N · m/rad.
The shaft turns a flywheel with mass moment of inertia
6.19.a Derive the system equations for the following: J = 2 kg · m 2 and damping b2 = 0.1 N · m · sec/rad .
i. The torque acting through spring K. 6.20.a Derive the system equations for the following:
ii. The angular velocity of inertia J. i. The current through resistance R.
iii. The velocity difference across spring K. ii. The back EMF.
Check their units. iii. The motor’s torque TM.
The system is at de-energized at time t = 0 when the sys- iv. The torque acting through the shaft.
tem is subjected to the step input torque of 800 N-m. v. The angular velocity of mass moment of inertia J.
6.19.b Solve the system equations. Check their units.
6.19.c Plot the responses from using Mathcad or MATLAB. The motor de-energized when a step input
v(t ) = 48VDC us (t ) was applied.
Problem 6.20 An electromechanical schematic of a DC motor 6.20.b Solve the system equations.
is shown in Fig. P6.20. The motor is powered by a voltage 6.20.c Plot the responses from time t = 0 to t = 6 τ sec
source v( t). The motor’s resistance is R = 4Ω . The motor’s using Mathcad or MATLAB.
References and Suggested Reading 409

References and Suggested Reading Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs
Rowell D, Wormley DN (1997) System dynamics: an Introduction.
Budynas RG, Nisbett KJ (2011) Shigley’s mechanical engineering Prentice- Hall, Upper Saddle River
Design, 9th edn. McGraw-Hill, New York Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
Fitzgerald AE et al (2003) Electric machinery, 6th edn. McGraw-Hill, dynamics. Addison-Wesley, Reading
New York
Vector-Matrix Algebra and the State-Space
Representation of Dynamic Systems 7

Abstract
An alternative to deriving a high-order differential system equation is to represent the sys-
tem as a set of simultaneous first-order differential equations of the system’s state variables,
which are its energy storage variables. This is known as state-space. State equations are
solved numerically, using finite difference approximations. They can also be solved using
Laplace transformations to yield transfer functions. State-space has three advantages. It is
much easier to derive a set of state equations than the corresponding high-order differential
equation. It is also easy to include non-linear properties, since the solution uses numerical
methods. Finally, state-space is the basis of Modern Control theory, which enables the con-
trol designer to arbitrarily tailor the response of a system, within the physical limitations of
the hardware.

7.1 Overview “state equations.” The set of simultaneous, first-order differ-


ential state equations must be solved numerically, to yield
Up to this point, we have derived first- and second-order sys- the energy storage (or state) variables. If our output variable
tem equations and solved them, using the method of undeter- is one of the energy storage variables, then we are done, once
mined coefficients and Laplace transformations. However, we have solved the state equations. Often, though, we are
many machines and processes have more than two indepen- interested in the force in a damper or other non-energy stor-
dent energy storage variables and, consequently, are higher- age variable. In those cases, a two-step process is used, in
order systems. The effort to reduce the equation list to a dif- which a set of simultaneous algebraic “output” equations,
ferential system equation increases, roughly with the square formulated in terms of the state variables, is solved, once the
of the number of the branches in the linear graph. The reduc- state variables have been determined.
tion of a third-order system is roughly nine times the effort of We will use Mathcad to solve this set of simultaneous dif-
the reduction of a first-order system. We do have an alterna- ferential (state) equations for the state variables, from which
tive. The equation list does not need to be reduced to a single, we can calculate every power variable in the system. In ad-
higher-order differential equation. We can formulate a set of dition to easing the reduction of higher-order systems, the
simultaneous, first-order differential equations containing numerical solution of the system, formulated using the state-
the information needed to describe the system. The infor- variable method, frees us from the restriction of using linear
mation needed to describe the “state” of the system at any elemental equations, imposed by our analytical method.
moment in time is the same information we use to determine
the initial value of the step response, an output variable in a
first-order system; the input variable; and the energy storage 7.2  Vector-Matrix Algebra
variable. The energy storage variables of a system are called
the “state variables,” since they allow us, with knowledge of Matrix notation was invented for the manipulation of sys-
the input, to calculate all the power variables in the system. tems of simultaneous equations. We will use matrices to ex-
The set of simultaneous, first-order differential equations, press the state equations and the output equations in a simple
written in terms of the state variables and time, are called the and compact form. The term “algebra” means a set of rules

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_7, © Springer Science+Business Media New York 2014 411
412 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

to manipulate mathematical symbols. There are many differ- should retain its properties as it is “slid” around in space,
ent types of algebra. The collection of rules for manipulat- allowing us to perform vector addition using the trapezoidal
ing matrices is vector-matrix algebra, or, more simply, just construction. Does a position vector retain its meaning, when
matrix algebra, because a vector is a matrix with only one its tail is no longer at the origin? It does, if you are willing to
row or column. Linear algebra is a set of rules that defines interpret “position” as “displacement.” We will call a posi-
the allowable operations on systems of linear simultaneous tion vector a vector. We will also refer to complex numbers
equations represented as matrices. The easiest way to under- as vectors because they obey most of the rules of vector ma-
stand the manipulations of matrix algebra or linear algebra nipulation, although they do not multiply as vectors. Com-
is to work with a system of equations, consisting of either plex numbers are clearly a special case. They do not exist in
two or three simultaneous equations. The rules that apply to three dimensions. Fourth-dimensional complex numbers are
these systems of equations can be extended to any system of “quaternions” invented by Hamilton.
equations. Unfortunately, the term “vector” has been extended from
Before we begin, we must address a bit of confusing ter- an ordered list of coordinates to any ordered list, primarily
minology. As you know, a “vector” is a physical quantity by the use of “vector” in computer science. For example,
with magnitude and direction, such as a force vector. Pos- we now speak of “data vectors,” which are, in fact, time se-
sessing magnitude and direction are necessary—but not suf- ries, or sequences of scalar variables, which have none of
ficient characteristics. To be a vector, the physical quantity the physical or mathematical properties of true vectors, such
must also obey the rules of vector algebra. For example, a as force vectors. It is important to be aware that there are
force vector can be decomposed into its components in or- multiple meanings. The mathematics which can be applied
thogonal directions to a vector of one meaning cannot necessarily be applied to a
vector of another meaning. We will strive to reduce possible
confusion between “vector,” meaning a physical quantity
F = Fx + Fy + Fz = Fx 
u x + Fy 
u y + Fz 
uz with magnitude and direction, and other quantities which are
only ordered lists, referring to the latter as a “column matrix”
which can be expressed as the product of two vectors or a “row matrix,” rather than the alternatives, “column vec-
tor” and “row vector.”
 ux 
 
Fx 
u x + Fy 
u y + Fz 
u z =  Fx Fy Fz  uy  7.2.1  Matrix Addition
 
 u z 
Matrix addition and subtraction are defined only for vectors
and matrices with the same number of rows and columns.
where the magnitudes of both the force components and the Addition and subtraction are performed element by element.
unit vectors in the principal directions are represented as or- For example, given the matrices
dered lists. We often use a further simplified form by omit-
ting the unit vectors in the principal directions, and present-
 a 1,1 a 1,2   h1,1 h1,2 
ing the magnitudes of the force components, as Cartesian A≡  and H≡ 
coordinates in an ordered list.  a 2,1 a 2,2   h 2,1 h 2,2 

then
 Fx Fy Fz 

 a 1,1 a 1,2   h1,1 h1,2   a 1,1 + h1,1 a 1,2 + h1,2 
It is this form which creates the confusion in terminology. A+H= + =
 a 2,1 a 2,2  
 h 2,1 h 2,2   a 2,1 + h 2,1 a 2,2 + h 2,2 
We read this ordered list as the coordinates of the head of a
vector in a three-dimensional “force space,” with its tail at
the origin. Hence, we can correctly refer to this ordered list (7.1)
as a vector, as it represents a physical quantity with magni- and
tude and direction, but only with reference to the force space.
We refer to a “position” vector r( t) as a vector extending 
from the origin to a point described by coordinates in a two-  a 1,1 a 1,2   h1,1 h1,2   a 1,1 − h1,1 a 1,2 − h1,2 
A−H= − h =  
or three-dimensional space. Most of us base our conception  a 2,1 a 2,2   2,1 h 2,2   a 2,1 − h 2,1 a 2,2 − h 2,2 
of a vector on position vectors, but are they truly vectors?
Do they obey the rules of vector manipulation? They do, de- (7.2)
pending on the meaning you apply to the “position.” A vector
7.2  Vector-Matrix Algebra 413

7.2.2  Matrix Multiplication Generalizing, the result of matrix multiplication of matrices


P and Q will have as many rows as P and as many columns
The matrix representation of simultaneous equations and as Q. It is important since matrix multiplication is evalu-
the procedure for matrix multiplication are both illustrated ated as row times column, first element times first element,
by expressing a system of simultaneous equations in matrix second element times second element, etc., the number of
form. Consider the system of equations columns in P must equal the number of rows in Q or the mul-
tiplication PQ cannot be evaluated.
 Having defined matrix multiplication in “expanded”
y1 = a1,1 x1 + a1,2 x2
(7.3) form, we can now simplify the notation and express the vec-
y2 = a 2,1 x1 + a 2,2 x2
tor-matrix equation vectors, y and x, and the matrix, A, as
variables
where xc and yr are variables, ar,c are constant coefficients,
and subscripts r and c indicate row and column, respectively.   y1 
This system of equations can be written in matrix notation as y≡  (7.5)
 y2 
  a1,1 a1,2 
 y1   a1,1 a1,2   x1   A≡ (7.6)
 y  = a a 2,2   x2 
(7.4)
 a 2,1 a 2,2 
 2   2,1

y   x1   a1,1 a1,2    x1 
where  1  and   are column matrixes and   x≡  (7.7)
 y2   x2   a 2,1 a 2,2   x2 
is a “two by two” matrix, since it has two rows and two col-
umns. where bold is used to denote a vector or matrix. Conven-
Matrix multiplication is performed by multiplying a row tionally, vectors are denoted by lower case, and matrices by
by a column, element by element, and then summing the upper case, where possible. This allows us to write the sys-
products. The mnemonic is R C. tem of equations, Eqs. 7.1 and 7.2, in a more compact but
For example, to multiply abstract form

 y = Ax (7.8)
 a1,1 a1,2   x1 
a a 2,2   x2 
 2,1 Although compact and convenient to manipulate, a matrix
equation written using single variables, such as y = Ax, re-
begin by multiplying the first row of the coefficient matrix A veals little about the system of equations it represents. For
and the column matrix x example, until vectors y and x, and matrix A are defined, the
equation y = Ax could represent a system of three simultane-
ous equations
 x1 
 a1,1 a1,2    = a1,1 x1 + a1,2 x2
 x2    y1   a1,1 a1,2 a1,3   x1 
 y  = a 
a 2,3   x2  (7.9)
next, multiply the second row of the matrix A and the col-  2   2,1 a 2,2
 y3   a 3,1 a 3,2 a 3,3   x3 
umn matrix x

or 300,000 as in the case of large finite element models


 x1 
 a 2,1 a 2,2    = a 2,1 x1 + a 2,2 x2
 x2    y1   a1,1 a1,2  a1, n   x1 
y  a a 2,2  a 2, n  x 
 2  =  2,1  2 (7.10)
The result will have as many rows as there are in the matrix        
A, and as many columns as there are in the vector x.      
y
 n a
 n ,1 an ,2  an , n   xn 
It is frequently necessary to expand a matrix equation writ-
 a1,1 a1, 2   x1   a1,1 x1 + a1, 2 x2 
a = ten in the form of Eq. 7.6 to that of Eq. 7.3, in order to con-
 2,1 a 2, 2   x2   a 2,1 x1 + a 2, 2 x2 
firm the matrix algebra. Whenever you are in doubt as to the
validity of your algebra or the meaning of a matrix-algebra
414 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

equation, expand the equation to a system of simultaneous 7.3 Operating on a Vector-Matrix Expression


equations to evaluate it. with a Linear Operator
Matrix multiplication is not commutative!
The most common errors in matrix algebra are (1) at- An operator is a linear operator, if doubling the input to the
tempting to perform an operation using two terms, which operator doubles the output. Differentiation with respect to
have incompatible numbers of rows and columns and (2) time is a linear operator. For example,
forgetting that matrix multiplication is not commutative.
For example, given the row matrix u and the column ma-  d 2x dx
=2 (7.15)
trix v dt dt
 v1 
u = u1 u 2  and v=  Operating on a matrix or a matrix equation is analogous to
v2  multiplying by a scalar. Differentiating the system of equa-
then tions of Eq. 7.1 with respect to time

 v1 
uv = u1 u 2    = u1v1 + u 2 v2 (7.11)
v2  dy1 d dy1 d d
= (a1,1 x1 + a1, 2 x2 ) = (a1,1 x1 ) + (a1, 2 x2 )
dt dt dt dt dt

which is a scalar. But dy2 d dy2 d d
= (a 2,1 x1 + a 2, 2 x2 ) = (a 2,1 x1 ) + (a 2, 2 x2 )
dt dt dt dt dt
  v1   v1u1 v1u 2 
vu =   u1 u 2  =  (7.12)
v2  v2 u1 v2 u 2  
dy1 dx dx
= a1,1 1 + a1, 2 2
Clearly, uv ≠ vu. These two products demonstrate that the dt dt dt (7.16)
number of rows and columns in the product of a matrix mul- dy2 dx1 dx2
= a 2,1 + a 2, 2
tiplication equals the number of rows in the first term of the dt dt dt
product, and the number of columns in the second term.
This is expressed in matrix notation as
7.2.2.1  Pre- and Post-Multiplication
The fact that vector-matrix is not commutative makes the 
d  y1  d   a1,1 a1,2   x1 
position of a factor in a product important. The terms “pre- =  (7.17)
multiplication” and “post-multiplication” are used to distin- dt  y2  dt   a 2,1 a 2,2   x2 
guish product uv from vu. In the former, u pre-multiplies v.
In the latter, v pre-multiplies u. Alternatively, in the former, v A “constant” matrix is a matrix whose elements are con-
post-multiplies u and in the latter, u post-multiplies v. stants, such as the constants ar,c in Eq. 7.15. Constant ma-
trices are factored out of the derivative of matrix products,
7.2.2.2  Multiplication of a Matrix by a Scalar just as scalar constants are factored out of the derivatives of
The product of a vector or matrix and a scalar (or a single scalar products. Hence
valued function) is evaluated by multiplying each term in the
vector or matrix by the scalar. For example, given a vector 
d  y1   a1,1 a1,2  d  x1 
= (7.18)
 b1  dt  y2   a 2,1 a 2,2  dt  x2 
B= 
b 2 
The same rule applies in the vector-matrix notation
and scalar variable c, the product is
 dy dAx
= (7.19)
  b1   b1c  dt dt
Bc =   c =   (7.13)
b
 2 b 2 c 
where vectors x and y are variables and matrix A is con-
The multiplication of a vector or matrix by a scalar is com- stant, resulting in
mutative
 dy dx
=A (7.20)
  b1   b1c   cb1   b1  dt dt
Bc =   c =   =   = c   = cB (7.14)
b 2  b2 c  cb2  b 2 
7.4  Transpose of a Matrix 415

7.3.1 Laplace Transformation of Matrix or a yields


Vector-Matrix Expression 
sY( s ) = AsX( s ) (7.26)
An important linear operator is the Laplace Transform. and

 a 1,2   x1  

L { f (t )} = ∫ f (t ) e − st
dt ≡ F ( s ) (7.21) 
L  dtd  yy  = L  dtd  aa
1 1,1

a 2,2   x2  
(7.27)
0   2    2,1

The functions f( t) and F( s) are functions of time t and the yields
complex variable s, respectively. The complex variable
s = σ + jw in the exponential within the Laplace integral is   s y1 ( s )   a1,1 a1, 2   s x1 ( s )   a1,1 a1, 2 
= (7.28)
the same complex variable s in the trial solution, x = Ae st ,

 s y2 ( s )  a 2,1 a 2, 2   s x2 ( s )  a 2,1 a 2, 2 
for the homogenous equation formed for a differential equa-
tion. We will not belabor the definition of Laplace transform, The convention is to express transformed variables by capi-
since it can be challenging to evaluate the integral, except tal letters, when possible.
in simple cases. In the future, we will use transform tables, The Laplace variable s can be factored out of either form
when needed. However, we can apply our test for linearity,
i.e., doubling the input must double the output, to the trans-  s y ( s )   a1,1 a1,2   s x1 ( s ) 
sY( s ) = AsX( s ) and  1  = 
form. The input to the Laplace transform is a function of  s y2 ( s )  a 2,1 a 2,2   s x2 ( s )
time f( t). Doubling the input is expressed as 2f( t).
∞ ∞
to yield

L {2 f (t )} = ∫ 2 f (t ) e − st
dt = 2∫ f (t ) e dt = 2 F ( s )
− st
(7.22)
0 0  y (s)  a1,1 a1,2   x1 ( s ) 
sY( s ) = sAX( s ) and s  1  = s 
 y 2 ( s )  a 2,1 a 2,2   x2 ( s )
The resulting transform is doubled.
Our use of the Laplace transform when working with ei- The Laplace transformation of integration with respect to
ther higher-order differential equations, or with systems of 1
time introduces the factor, , which is, accordingly, the
first-order differential equations, will be to transform dif- s
ferential equations into algebraic equations, neglecting the inverse of the Laplace transformation of differentiation with
initial condition terms. Recall the Laplace transformation respect to time.
derivatives with respect to time, when the initial conditions 
are neglected. L {∫ f (t )dt} = 1s F (s) (7.29)

 L  dtd f (t ) = sF (s) (7.23)


  7.4  Transpose of a Matrix

and The transpose of a vector or a matrix is created by inter-


changing the rows and columns. The operation is indicated
L  ddt 
n
 f (t ) = s n F ( s ) by a superscript “T.” For example,
(7.24)
 
  a1 
d A =  a1 a 2  → A T =   (7.30)
The operator
dt
transforms to the complex variable s a 2 
which operates on the transformed function of time,
L { f (t )} = F ( s), as a multiplicative factor. The transfor- 
mation of differential equations to polynomials in the com-  b1 
B =   → B T = b1 b2  (7.31)
plex variable s allows us to work with algebraic equations, b 2 
rather than differential equations.
We will apply the Laplace transform to systems of simul-   c1,1 c 2,1 
taneous, first-order differential equations, where,  c1,1 c1,2 c1,3   
T
C=  → C = c1,2 c 2,2  (7.32)
 c 2,1 c 2,2 c 2,3 
L  ddty  = L  ddtAx  (7.25)  c1,3
 c 2,3 
   
416 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

and A fractional inverse can be written in two complementary


forms.
  d 1,1 d 1, 2 d 1,3   d 1,1 d 2,1 d 3,1   1
    = α −1 (7.35)
 (7.33)
T
D =  d 2,1 d 2, 2 d 2,3  → D =  d 1, 2 d 2, 2 d 3, 2 α
 d 3,1 d 3, 2 d 3,3   
  d 1,3 d 2,3 d 3,3  Using the laws of exponents, and recalling that any base
raised to the zeroth power equals 1, a unity ratio can be ex-
Note that the elements along the main diagonal of square pressed as the product of a scalar and its fractional inverse.
matrix C, which have repeated indices, do not move when
the transpose is created, Eq. 7.32.  α
= α ⋅ α −1 = α 1 ⋅ α −1 = α 1−1 = α 0 = 1 (7.36)
α

7.5  Matrix Inversion The number one is the “identity” quantity for scalar multipli-
cation. Eq. 7.34 is true for any scalar.
Consider the following scalar equation of a straight line
 α ·1 = 1 · α = α (7.37)
y = mx + b
There is a corresponding identity quantity in vector-matrix
We solve for x by first subtracting b from both sides, and algebra, known as the “identity matrix” and given the sym-
then dividing both sides by m bol, I, which, when multiplied by a vector or matrix, yields
the original vector or matrix. Although the phrase “identity
y −b matrix” is used, the identity matrix for a specific vector or,
y = mx + b → y − b = mx → =x
m if matrix is a square matrix, with as many rows and columns
as the largest number of rows or columns in the vector or
This simple bit of scalar algebra is impossible in vector- matrix, with ones along the main diagonal (top left corner to
matrix algebra. Although we can perform the subtraction, bottom right corner) and zeros for every other element.
matrix division is not defined!
Although there is no matrix division, there is an analo-  1 0  0
gous operation called matrix inversion. We will develop 0 1  0
the concept with scalar algebra, and then apply it to vector- I= (7.38)
0 0  
matrix algebra. Consider the fractional inversion of a scalar  
variable. The product of a non-zero scalar and its fractional 0 0  1
inverse is a unity ratio.
The simplest case is a square matrix M. A square matrix can
 1 α be either pre- or post-multiplied by its identity matrix.
α · = =1 for α ≠ 0 (7.34)
α α 
MI = IM = M (7.39)

For example,

 m1,1 m1, 2 
M= 
 m 2,1 m 2, 2 
 m1,1 m1, 2  1 0  m1,1 ⋅1 + m1, 2 ⋅ 0 m1,1 ⋅ 0 + m1, 2 ⋅1   m1,1 m1, 2 
MI =   = = =M
 m 2,1 m 2, 2  0 1  m 2,1 ⋅ 0 + m 2, 2 ⋅1 m 2,1 ⋅1 + m 2, 2 ⋅ 0  m 2,1 m 2, 2 

and

1 0  m1,1 m1, 2   1⋅ m1,1 + 0 ⋅ m1,2 0 ⋅ m1,1 + 1⋅ m1,2   m1,1 m1, 2 


IM =   = = =M
0 1  m 2,1 m 2, 2  0 ⋅ m 2,1 + 1⋅ m 2,2 1⋅ m 2,1 + 0 ⋅ m 2,2   m 2,1 m 2, 2 
7.5  Matrix Inversion 417

The multiplication of a vector or a rectangular matrix by 1 0 0  d 1,1 d 1,2 


its corresponding identity matrix is only possible for the  
ID = 0 1 0  d 2,1 d 2,2
arrangement, which yields the number of rows of the pre-
multiplying quantity equal to the number of columns of the 0 0 1   d 3,1 d 3,2
post-multiplying quantity. For example, row vector A must
be post-multiplied by its identity matrix. 1⋅ d 1,1 + 0 ⋅ d 2,1 + 0 ⋅ d 3,1 1⋅ d 1,2 + 0 ⋅ d 2,2 + 0 ⋅ d 3,2
 
ID = 0 ⋅ d 1,1 + 1⋅ d 2,1 + 0 ⋅ d 3,1 0 ⋅ d 1,2 + 1⋅ d 2,2 + 0 ⋅ d 3,2
A =  a1 a 2  0 ⋅ d 1,1 + 0 ⋅ d 2,1 + 1⋅ d 3,1 0 ⋅ d 1,2 + 0 ⋅ d 2,2 + 1⋅ d 3,2
 
1 0
AI =  a1 a 2    =  a1 ·1 + a 2 · 0 a1 · 0 + a 2 ·1
0 1  d 1,1 d 1,2 
 
=  a1 a 2  D =  d 2,1 d 2,2
 d 3,1 d 3,2
 
but

1 0 but
IA =    a1 a 2  Cannot be evaluated
0 1
 d 1,1 d 1,2  1 0 0 
 
Whereas, column vector B must be pre-multiplied by its DI =  d 2,1 d 2,2  0 1 0  Cannot be evaluated
identity matrix  d 3,1
 d 3,2  0 0 1 
 b1 
B=  The inverse of matrix M, if one exists, is defined as the ma-
b 2 
trix, which when pre- or post-multiplying M, yields the iden-
1 0  b1  1⋅ b1 + 0 ⋅ b2   b1  tity matrix for M. The symbol used for an inverse matrix is
IB =    =  = 
0 1 b2  0 ⋅ b1 + 1⋅ b2  b2  the exponent, − 1, e.g., the inverse of M is M−1.

but  MM −1 = M −1M = I (7.40)

 b1  1 0 Matrix inversion is used in vector-matrix algebra to solve


BI =     Cannot be evaluated
b2  0 1 matrix equations. For example, to solve y = Ax for x, we
multiply both sides of the equation by the inverse of A.
Rectangular matrix C must be post-multiplied by its identity
matrix

1 0 0
 c1,1 c1,2 c1,3  
CI =   0 1 0
c 2,1 c 2,2 c 2,3  0 0 1

 c1, 1 ⋅1 + c1, 2 ⋅ 0 + c1, 3 ⋅ 0 c1, 1 ⋅ 0 + c1, 2 ⋅1 + c1, 3 ⋅ 0 c1, 1 ⋅ 0 + c1, 2 ⋅ 0 + c1, 3 ⋅1 
=
c 2, 1 ⋅1 + c 2, 2 ⋅ 0 + c 2, 3 ⋅ 0 c 2, 1 ⋅ 0 + c 2, 2 ⋅1 + c 2, 3 ⋅ 0 c 2, 1 ⋅ 0 + c 2, 2 ⋅ 0 + c 2, 3 ⋅ 0
 c1, 1 c1, 2 c1, 3 
= =C
c 2, 1 c 2, 2 c 2, 3 

but
y = Ax → A −1y = A −1 Ax
1 0 0 
 c1,1 c1,2 c1,3 
IC = 0 1 0    Cannot be evaluated
c 2,1 c 2,2 c 2,3  A −1y = Ix → A −1y = x
0 0 1 
Whereas rectangular matrix D must be pre-multiplied by its
identity matrix
418 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Note that an identity matrix I factor can be removed from or


inserted into a vector-matrix equation at will, with the restric-  a 1,1 a 1,2  a 1,1 a 1,2 a 1,1 • a 1,2
a = • −
a 2,2  a 2,1 a 2,2
tion that the resulting matrix product can be evaluated. As a
second example, we will solve the vector-matrix equation,  2,1 a 2,1 a 2,2
y = Mx + b for x
y = Mx + b → y − b = Mx → M −1 (y − b) = M −1Mx Fig. 7.1   Graphical procedure for calculating a two by two determinant

M −1 (y − b) = Ix → M −1 (y − b) = x
It is easier to remember the pattern of the elements of the
matrix, shown in Fig. 7.1, rather than the order of the indices.
7.5.1  Calculation of an Inverse Matrix The meaning of a determinant is clearest, when we con-
sider the simplest equation, a two by two matrix A, y = Ax
We will preface this discussion of calculating inverse matri- which expands into a set of two simultaneous equations.
ces with the reminder, that the inverse of a matrix may not
exist, due to the fact that the system of equations represented  y1   a 1,1 a 1,2   x1 
y = Ax →   = 
by the matrix equation has no solution. We will also note that
 y2   a 2,1 a 2,2   x2 
applied mathematicians have put great effort into developing
efficient methods of solving large sets of simultaneous equa-
tions. For computational efficiency and the resulting speed, y1 = a 1,1 ⋅ x1 + a 1,2 ⋅ x2
efficient techniques specifically avoid the task of inverting y2 = a 2,1 ⋅ x1 + a 2,2 ⋅ x2
a large matrix. Although computer code will avoid invert-
ing matrices, the theoretical equations on which the code is For a solution x1 and x2 to exist, the two equations must be
based must use matrix inversion, since matrix division is not independent. Geometrically, in two-dimensional space, the
defined. equations must represent two lines which intersect at one
Engineering software, including Mathcad and MATLAB, point. Two cases without any solution occur when the two
plus an increasing number of handheld calculators include lines are identical, and intersect at every point, or are parallel
functions to calculate the inverse matrices. Although we will and have no intersection. The determinant of a set of simulta-
not compute inverse matrices manually, it is important to un- neous equations equals zero in either of these cases.
derstand the process of inverting a matrix, because interme- Example 1 Two identical equations (or lines)
diate results in the process are themselves important mathe-
matical entities. The mathematical terminology is somewhat y1 = 3 ⋅ x1 + 4 ⋅ x2  y1  3 4  x1 
opaque, in part, because the processes can only be defined →  =  
y2 = 3 ⋅ x1 + 4 ⋅ x2  y2  3 4  x2 
using the names of intermediate quantities. We will define
the terms and describe their calculation, after we use them to
define the inverse of a matrix.  a1,1 a1,2  3 4
A =  = = 3 ⋅ 4 − 3 ⋅ 4 = 12 − 12 = 0
The inverse of matrix A is calculated as the transpose of  a2,1 a2,2  3 4
the cofactors of A divided by the determinant of A.
Example 2 Two dependent equations (or two parallel lines)

adjoint ( A) cofactors ( A)T cof ( A) T
A −1 = = ≡ y1 = 3 ⋅ x1 + 4 ⋅ x2 y1 = 3 ⋅ x1 + 4 ⋅ x2
determinant ( A) determinant ( A) A →
y2 = 2 (3 ⋅ x1 + 4 ⋅ x2 ) y2 = 6 ⋅ x1 + 8 ⋅ x2
(7.41)
 y1   3 4  x1 
 y  = 6 8   x 
 2   2
7.5.2  Determinant of a Matrix
 a1,1 a1,2   3 4
The determinant of a matrix is a scalar. The determinant of A =  = = 3 ⋅ 8 − 6 ⋅ 4 = 24 − 24 = 0
a two by two matrix is easy to remember, and, as the funda-  a 2,1 a 2,2  6 8 
mental definition, forms the basis of a recursive algorithm to
calculate the determinant of larger matrices. The determinant We have stated without proof that the inverse of the matrix
of a two by two matrix is A is calculated as

  a1, 1 a1, 2   adjoint ( A) cofactors ( A)T cof ( A)T


A =  = a1, 1 · a 2, 2 − a1, 2 · a 2, 1 A −1 = = ≡
(7.42)
 a 2, 1 a 2, 2  determinant ( A) determinant ( A) A

 (7.41)
7.5  Matrix Inversion 419

Fig. 7.2   Expansion of the


first minor of a three by three  a 1,1 a 1,2 a 1,3 
determinant    a 2,2 a 2,3   a 2,1 a 2,3   a 2,1 a 2,2 
A =  a 2,1 a 2,2 a 2,3  = a 1,1 a − a 1,2 + a 1,3
 3,2 a 3,3  a
 3,1 a 3,3  a
 3,1 a 3,2 
 a 3,1 a 3,2 a 3,3 

Fig. 7.3   Expansion of the


second minor of a three by three
determinant  a 1,1 a 1,2 a 1,3 
   a 2,2 a 2,3   a 2,1 a 2,3   a 2,1 a 2,2 
A =  a 2,1 a 2,2 a 2,3  = a 1,1 a − a 1,2 + a 1,3
 3,2 a 3,3  a
 3,1 a 3,3  a
 3,1 a 3,2 
 a 3,1 a 3,2 a 3,3 

Fig. 7.4   Expansion of the


third minor of a three by three
determinant  a 1,1 a 1,2 a 1,3 
   a 2,2 a 2,3   a 2,1 a 2,3   a 2,1 a 2,2 
A =  a 2,1 a 2,2 a 2,3  = a 1,1 a − a 1,2 + a 1,3
 3,2 a 3,3  a
 3,1 a 3,3  a
 3,1 a 3,2 
 a 3,1 a 3,2 a 3,3 

If there is no solution to the system of simultaneous equa- The process of expansion by minors is easier to remember
tions, the determinant of A is zero, and the calculation for by the patterns, shown in Figs. 7.2, 7.3 and 7.4 than by a
the inverse of matrix A fails, since it is impossible to divide formula.
by zero or, if you prefer, division by zero yields infinity. A The expansion terms are multiplied by a sign factor calcu-
matrix which cannot be inverted is termed “singular,” a word lated by raising minus one to the power equal to the sum of
used in Victorian England for “unique” or “special” and used the row and column of the element, which is the coefficient
in mathematics to mean “division-by-zero.” of the determinant
The fundamental definition of a determinant is that for
a two by two, as given in Eq. 7.42. The determinant of a  + − + 
larger matrix is calculated by “expanding the matrix by mi- sign factor of
= ( −1) (i + j )
= ( −1) (r + c)
=  − + −  (7.43)
nors,” repeatedly if necessary, until the determinants are two expansion term
 + − + 
by two. The process consists of removing a row or column
from the matrix, and multiplying the elements of that row
or column by the determinant of the submatrix, created by where i and j are the row and column indices. Note that
dropping out both the row and column of the element. The the sign factors alternate from positive to negative in a
determinant of a three by three matrix is checkerboard pattern. (Although conventional mathemati-
cal notation uses i and j for indices, r and c are useful mne-
 a 1,1 monics and cannot be confused with the imaginary number
a 1,2 a 1,3 
  i ≡ j ≡ −1.)
A =  a 2,1 a 2,2 a 2,3  We will illustrate the calculation of the determinant of a
 a 3,1 a 3,2 a 3,3  three by three matrix using two cases. The first case will be

three lines, which form orthogonal axes intersecting at the
 a 2,2 a 2,3   a 2,1 a 2,3   a 2,1 a 2,2  origin. Orthogonality, or mutually normal axes, is an impor-
A = a 1,1  − a 1,2 + a 1,3 tant property of engineering systems described by matrix
 a 3,2 a 3,3  a
 3,1 a 3,3  a
 3,1 a 3,2 
equations, characterized by a square matrix with non-zero
420 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

element on the main diagonal (top left to bottom right) and b1,1 b1,2 b1,3 b1,4
zeros elsewhere. b 2,1 b 2,2 b 2,3 b 2,4
B =
b 3,1 b 3,2 b 3,3 b 3,4
 y1   2 0 0  x1 
b 4,1 b 4,2 b 4,3 b 4,4
y = Px →  y2  =  0 3 0  x2 
 y3   0 0 4  x3 
b 2,2 b 2,3 b 2,4 b 2,1 b 2,3 b 2,4
B = b1,1 b 3,2 b 3,3 b 3,4 − b1,2 b 3,1 b 3,3 b 3,4
This is a “diagonal” matrix. Diagonal matrices are very im-
b 4,2 b 4,3 b 4,4 b 4,1 b 4,3 b 4,4
portant in engineering mathematics because the values of the
elements on the main diagonal are the “eigen” or characteris- b 2,1 b 2,2 b 2,4 b 2,1 b 2,2 b 2,3
tic values of the system, described by the set of simultaneous + b1,3 b 3,1 b 3,2 b 3,4 − b1,4 b 3,1 b 3,2 b 3,3
equations. We are free to expand a matrix by minors, using
b 4,1 b 4,2 b 4,4 b 4,1 b 4,2 b 4,3
a row or column of our choice. Consequently, we always
choose a row or column with as many zeros as possible.
There is one special case which must be considered. What
 2 0 0 is the result, if one takes the determinant of a scalar? This
3 0 0 0 0 3 3 0 special case is handled by definition, since the calculation
P =  0 3 0 = 2 ⋅ − 0⋅ + 0⋅ = 2⋅
0 4 0 4 0 0 0 4 presented above cannot be performed on a single scalar num-
 0 0 4 ber. The determinant of a scalar is defined to be the number
itself, i.e.,
P = 2 ⋅ (3 ⋅ 4 − 0 ⋅ 0) = 24 
determinant ( a ) ≡ a = a where a is a scalar (7.44)
The determinant of P is not zero. Therefore, matrix P can be
inverted to solve the equation. 7.5.3  Cofactor of a Matrix
The second example is a system of three simultaneous
equations, which represents two lines that lie in the same The cofactors of a matrix A are defined in terms of the deter-
plane; are parallel; and a third line which is normal to the minant of A. The cofactors are the values of the determinants
plane of the two parallel lines. This system of equations has of the minors (submatrices), created by deleting the row and
no solution. column of each element of the matrix. The cofactor of a ma-
trix A is the matrix of its cofactors. This seemingly circular

 y1   2 3 0  x1 
y = Qx     
→  y2  =  4 6 0  x2 
 y3   0 0 5  x3 

 2 3 0
6 0 4 0 1 6 6 0 4 0 4 6
Q =  4 6 0 = 2 ⋅ − 3⋅ + 0⋅ = 2⋅ − 3⋅ +0
0 5 0 5 0 0 0 5 0 5 0 0
 0 0 5

Q = 2 ( 6 ⋅ 5 − 0 ⋅ 0) − 3 ( 4 ⋅ 5 − 0 ⋅ 0) − 0 ( 4 ⋅ 5 − 0 ⋅ 0) = 0

The determinant is zero. Therefore, matrix Q is singular and definition requires clarifying illustrations. We will work with
cannot be inverted to solve the equations. a three by three matrix.
Expanding a matrix larger than three by three is a recur- The elements of the cofactor are multiplied by the same
sion of this algorithm. A four by four matrix is first expanded sign factor as the expansion term in a determinant.
by a row or column of choice into four terms, each including
the determinant of a three by three matrix, which are then sign factor of cofactor element = ( −1) (i + j )
each expanded as illustrated above. Using the first row to + − + 
expand the determinant of the matrix B yields
= ( −1) (r +c)
=  − + − 
 + − + 
7.5  Matrix Inversion 421

  a 2,2 a 2,3  a a 2,3  a a 2,2  


( +1)  ( −1)  a 2,1 ( +1)  a 2,1 
  a 3,2 a 3,3   3,1 a 3,3   3,1 a 3,2  
  a 1,1 a 1,2 a 1,3   
    a 1,2 a 1,3  a a 1,3  a a 1,2  
cofactor ( A ) = cofactor   a 2,1 a 2,2 a 2,3  = ( −1)  ( +1)  a 1,1 ( −1)  a 1,1
 a 3,2 a 3,3   a 3,3   a 3,2  
  a 3,1 a 3,2 a 3,3   3,1 3,1 
  
 +1  a 1,2 a 1,3  a a 1,3  a a 1,2  
( )  a 2,2 a 2,3 
( −1)  a 1,1 a 2,3 
( +1)  a 1,1 a 2,2  
   2,1  2,1

7.5.4  Adjoint of a Matrix


 2 0 0
3 0 0 0 0 3
The adjoint of the matrix A is the transpose of cofactor( A): A =  0 3 0 = 2 ⋅ − 0⋅ + 0⋅
0 4 0 4 0 0
 0 0 4
adjoint ( A ) = cofactor ( A )
T

T
  a 2,2 a 2,3  a a 2,3  a a 2,2   3 0
( +1)  ( −1)  a 2,1 ( +1)  a 2,1  A = 2⋅ = 2 ⋅ (3 ⋅ 4 − 0 ⋅ 0) = 24
  a 3,2 a 3,3   3,1 a 3,3   3,1 a 3,2   0 4
 
  a 1,2 a 1,3  a a 1,3  a a 1,2  
= ( −1)  ( +1)  a 1,1 ( −1)  a 1,1
 a 3,2 a 3,3   a 3,3   a 3,2   It remains to calculate cofactor(A).
 3,1 3,1

 a a 1,3  a a 1,3  a a 1,2  
( +1)  1,2 ( −1)  a 1,1 ( +1)  a 1,1   2 0 0
  a 2,2 a 2,3   a 2,3   a 2,2   cof ( A ) = cof  0 3 0
 2 ,1 2,1 
 0 0 4

  a 2,2 a 2,3  a a 1,3  a a 1,3    3 0 0 0  0 3 


( +1)  ( −1)  a 1,2 ( +1)  a 1,2 ( +1)  ( −1)  ( +1)  
 a 3,3  a 2,3  

0 4 
0 4 0 0 
  a 3,2 a 3,3   3,2  2,2 
   
0 0  2 0  2 0 
  a 2,1 a 2,3  a a 1,3  a a 1,3   cof ( A ) = ( −1)  ( +1)  ( −1)   
=  ( −1)   ( +1)  a 1,1 ( −1)  a 1,1  4 
  a 3,1 a 3,3   3,1 a 3,3   2,1 a 2,3   
0  0 4  0 0

   0 0  2 0  2 0 
  a 2,1 a 2,2  a a  a a 1,2   ( +1)  ( −1)  ( +1)   
 ( +1)  a ( −1)  a 1,1 a 1,2  ( +1)  a 1,1 3 0 
 0 0  0 3 
 3,1 a 3,2 

 3,1 3,2   2,1 a 2,2   
 
 3 0 
( +1)   0 0 
cof ( A)T  0 4 
We will now demonstrate that the expression A −1 = ,  
A  2 0
cof ( A ) =  0 ( +1)   0 
does indeed yield the inverse of A. To limit the manual com-   0 4 
 
putation, we will use a diagonal matrix   2 0 
 0 0 ( +1)   
 2 0 0   0 3 
[ A] =  0 3 0
(3 ⋅ 4 − 0 ⋅ 0) 0 0 
 0 0 4  
cof ( A ) =  0 ( 2 ⋅ 4 − 0 ⋅ 0) 0 
We have calculated the determinant of this matrix and know  0 0 ( 2 ⋅ 3 − 0 ⋅ 0)
that this matrix can be inverted.
12 0 0
cof ( A ) =  0 8 0
 0 0 6
422 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

We now transpose cofactor(A) to yield adjoint(A). 7.6 State-Space Representation of Dynamic


Cofactor(A) is a diagonal matrix and, since the transpose of Systems
a diagonal matrix equals the diagonal matrix, the transpose
of cofactor(A) equals cofactor(A). 7.6.1  State Variables

T One meaning of the word “state” is “condition.” A physician


12 0 0 12 0 0
describes the state of a patient’s health by his or her vital sta-
cofactor( A) =  0 8 0  =  0 8 0 
T  
tistics which are variables, such as temperature, blood pres-
 0 0 6   0 0 6  sure, and white blood cell count. One could refer to the vital
statistics as state variables, because these are the variables
The final step is to divide adjoint(A) by determinant(A). needed to determine the state of a person’s health. Likewise,
the word state also means a condition when used in systems
12 0 0  12 dynamics. The condition of a dynamic system is defined by

 0 8 0  0 0 the values of the power variables. For example, the pressure
   24  in a hydraulic system is analogous to blood pressure in a
cof ( A )
T
 0 0 6  8
−1
A = = = 0 0 person.
A 24  24 
  We have referred to state variables in our study of the
0 6 dynamic response of first-order systems. The energy stor-
0
 24  age variable is the state variable of a first-order system. We
calculated the initial values of power variables in first-order
0.500 0 0  systems, using the values of the input variable and the energy
−1
A = 0 0.333 0  storage variable at time, t = 0+ . If we know values of the
 0 0 0.250 input, and the energy storage variables over any duration of
time t, we are able to calculate any other power variable in
the system over that duration. Consequently, the energy stor-
We will now test our result by evaluating the product A −1 A, age variables are the state variables for the dynamic system.
which should yield a matrix approximately equal to the iden- State variables define the state of the system, because if we
tity matrix.

0.500 0 0   2 0 0
−1 
A A= 0 0.333 0   0 3 0
 0 0 0.250  0 0 4

( 0.500 ⋅ 2 + 0 ⋅ 0 + 0 ⋅ 0) (0.500 ⋅ 0 + 0 ⋅ 3 + 0 ⋅ 0) (0.500 ⋅ 0 + 0 ⋅ 0 + 0 ⋅ 4)



A A = ( 0 ⋅ 2 + 0.333 ⋅ 0 + 0 ⋅ 0)
−1
(0 ⋅ 0 + 0.333 ⋅ 3 + 0 ⋅ 0) (0 ⋅ 0 + 0.333 ⋅ 0 + 0 ⋅ 4)
( 0 ⋅ 2 + 0 ⋅ 0 + 0.250 ⋅ 0) (0 ⋅ 0 + 0 ⋅ 3 + 0.250 ⋅ 0) (0 ⋅ 0 + 0 ⋅ 0 + 0.250 ⋅ 4)
1 0 0
A A = 0 0.999 0 ≈ I
−1 
0 0 1 

Success. In practice, it is unlikely that you will ever have know these variables as functions of time, we can solve for
need to manually calculate the inverse of a matrix. Although any other power variable for any time.
it is possible to program the algorithm we just executed, it There is an adage that “state variables are not unique.”
is not used in engineering software, because there are much This refers to the ease with which we can use vector-ma-
more computationally efficient methods available, based on trix algebra to manipulate a state-space description of a
Gaussian elimination. dynamic system. The fact that we can perform algebra on
7.6  State-Space Representation of Dynamic Systems 423

R L Table 7.1   Power variables in a series RLC circuit


1 2 3
Across Variable Through Variable
Voltage Source Known i
iR iL Resistor v12 iR
v(t) isource iC C Inductor v23 iL (state variable)
Capacitor v3g (state variable) iC

g
There are many advantages of representing a system as
Fig. 7.5   Series RLC circuit showing nodes of distinct voltages between
a higher-order differential system equation, introduced in
the elements and positive direction of currents through the elements
Chap. 2 and to be investigated further in Chap. 9. Recall the
algebraic characteristic equation of a system can be written
the representation of a dynamic system does not change its by inspection of the higher-order differential system equa-
underlying properties. We will investigate two very differ- tion. The roots of the characteristic equation are the char-
ent methods of formulating state equations, and find that acteristic values, or eigenvalues, of the system, which do
although the state equations produced are very different, indeed characterize the homogeneous response of the sys-
and both describe the same dynamic system, they use dif- tem.
ferent sets of state variables. However, we derive a set of In this chapter, we will develop a technique to derive the
state equations from the schematic of the energetic model of transfer functions of the state variables of a system from its
a system. The set of state variables will be the energy storage state-space representation. Since transfer functions can also
variables of the system. be formed from the Laplace transformation of a differential
system equation, we will be able to form the correspond-
ing system equation from a transfer function. We will find
7.6.2 Example Second-Order Dynamic System the technique of transforming state equations to transfer
RLC Circuit function to system equation to be significantly easier, than
the direct reduction of an equation list to a higher-order dif-
There are two power variables in every dynamic element. ferential system equation.
The power variables are unknowns, except for one of the two
power variables in a source which we control. Consider the 7.6.2.2 State-Space Representation of the Series
RLC circuit shown in Fig. 7.5. The power variables in the RLC Circuit
circuit are shown in Table 7.1. The state-space representation can be thought of as a par-
tial reduction of the equation list to a set of simultaneous
7.6.2.1 Second-Order Differential System differential equations, rather than to a single higher-order
Equation of the Series RLC Circuit differential equation. We will use the energy storage vari-
We reduced this circuit as an example in Sect. 1.5.2.2 to ables of a system as its state variable. The state variables of a
yield a second-order differential equation relating input v1g system are not unique, because any legitimate linear algebra
and voltage across capacitor v3g. operation performed on the set of state equations can yield
new state variables as weighted sums of the original state
 variables. Some definitions of non-physical state variables
1 d 2 v3 g R dv3 g 1
v1g = + + v3 g (1.48) are important, because of their mathematical convenience in
LC dt 2 L dt LC Modern Control theory, which is based on the state-space
representation.
If we were interested in another output variable, say the cur- There are two independent energy stores in RLC circuit:
rent through the inductor iL, we would need to start from the (1) the capacitor, which stores energy in an electric field, and
equation list and perform a second complete reduction. The (2) the inductor, which stores energy in a magnetic field. The
effort of reducing of an equation list to a single differential state variables are the energy storage variables of these two
equation with the same order as the system increases roughly elements, v3g and iL. The energy storage elements of a system
as the square of the number of elements in the system. Un- make the system dynamic. The flow of energy into or out of
fortunately, as reductions become more involved, they also a storage element occurs at a finite rate, and is described by
become more error prone. a differential equation, relating the derivative of the energy
storage variable (a state variable) to the other power variable
424 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

R L 3. Proceed to eliminate by substitution all power variables,


1 2 3 except for the input variable and the state variables. The
resulting first-order differential equation is the state equa-
tion.
Eliminate the output variables, unless they are also state
v(t) C variables. We will calculate the output variable after cal-
culating the state variables.
4. Express the state equation in standard form.
g g The left side:
• Only the derivative of the state variable without a coef-
Fig. 7.6   Linear graph of the series RLC circuit of Fig. 7.5 ficient may be on the left side of a state equation.
• The state equations must be listed, so that the derivatives
of the element. The elemental equations of energy dissipat- of the state variables are in the same order as the state
ers; transducer and transformer equations; and the continuity variables in the state vector.
and compatibility equations are all algebraic equations. As The right side:
we know from solving for the initial condition of the output • The coefficients of each state variable are collected, so
variable in a first-order system, if the values of the state vari- that there is only one term for each state variable.
ables are known at any time t, then it is possible to determine • The state variables appear in the same order as the state
the values of all of the other unknown power variables at vector x.
that time. • Constants multiplying state variables and the input vari-
The state-space representation of a dynamic system sepa- able are distributed.
rates calculation of the state variables, and calculation of all • The input term is last.
of the other power variables into two discrete steps. We will List the input variable and the state variables before begin-
first formulate the state equations to find state variables v3g ning the reduction. The order of the state variables in the list
and iL. We will then formulate the output equations to cal- will be the order of the state variables in state vector x.
culate all of the remaining unknown power variables, using
input v, and state variables v3g and iL.
 iL 
x= 
Energetic Equations v3 g 

Compatibility:  v(t ) = v12 + v23 + v3 g Refer the list of state variables, so that you don’t lose track
of the variables you wish to keep, as well as those you must
Continuity:  i − iR = 0 iR − iL = 0 iL − iC = 0 eliminate.
dv3 g State-Equation Reduction
diL
Elemental:  v12 = RiR v23 = L iC = C Input Variable v, State Variables iL , v3 g
dt dt First State Equation Begin the reduction with the state
1 2 1
Energy:  E sys = EC + E L EC = Cv3 g E L = LiL2 equation for the first state variable, iL. Start with the elemen-
2 2 tal equation for the inductor.

7.6.3  State Equations diL


v23 = L
dt
State equations are derived by partial reductions of the equa-
tion list to a set of first-order differential equations written in
terms of the input variable, the state variables, the elemental Rearrange to put the derivative of state variable iL on the left
parameters, and time. side by itself without a coefficient.
The reduction procedure is always the same:
1. Start the reduction for a state equation with the elemental diL v23
=
equation of an energy storage element of the system. You dt L
will have as many state equations, as there are indepen-
dent energy storages in the system. The variable on the right side, v23, is neither the input nor a
2. Rearrange the energy storage elemental equation to place state variable. Eliminate it by substitution.
the derivative of the state variable on the left side by itself
diL 1
without a coefficient.
dt
(
= v − v12 − v3 g
L
)
7.6  State-Space Representation of Dynamic Systems 425

Eliminate v12 by substitution because it is neither the input variables. The two equations are coupled to one another.
nor a state variable. Both must be satisfied, if either is to be true. We will use the
Runge–Kutta numerical method to solve sets of simultane-
diL 1
dt
(
= v − RiR − v3 g
L
) ous differential state equations. The Runge–Kutta method is
available in Mathcad and MATLAB as built-in functions. We
will also program the algorithm in MATLAB.
Eliminate iR by substitution because it is neither the input nor The state equations must be solved before the output
a state variable. equations are solved. Solution of state equations requires
a numerical method for solving simultaneous differential
diL 1
dt
(
= v − RiL − v3 g
L
) equations, such as the Runge–Kutta method. Solution of the
output equations will be purely algebraic, in terms of the
state variables and the input variable.
This is a state equation with v as the input and iL and v3g as
state variables. Express the state equation as a summation,
by distributing the factor 1/L. Order the terms in the order 7.6.4  Output Equations
that the state variables appear in the state vector, followed by
the term with the input variable. If we know input v and the state variables, iL and v3g, over a
period of time, then we can calculate every unknown power
diL 1 R 1 di R 1 1 variable in the system over the same period. “Output” equa-
= v − iL − v3 g → L = − iL − v3 g + v
dt L L L dt L L L tions are formulated in terms of the input and the state vari-
ables. The calculation sequence is fixed by necessity. The
Second State Equation  Begin with the elemental equation state equations must be solved before the output equations.
of the capacitor. Solution of the state equations requires a technique for solv-
ing a set of simultaneous differential equations. The solution
dv3 g of the output equations is purely algebraic. Consequently,
iC = C only use the algebraic equations of the system’s energetic
dt
equations. Never use the elemental equations for the energy
Rearrange to put the derivative of state variable v3g on the storage elements, since these are differential equations.
left side.
Example Output Equation  Derive the output equation for
dv3 g i
= C the voltage across inductor, v23. Identify the input, output,
dt C and state variables, before beginning a reduction.
Input v, Output v23, State Variables iL, v3g
Eliminate iC since it is neither the input nor a state variable. We will start the reduction with a compatibility equation.

dv3 g iL
= v = v12 + v
23 + v3g
dt C Input

Desired State
Known Unknown Variable
Known
This is a state equation in standard form with v3 g and iL as
state variables.
Rearrange to express the desired unknown on the left. Place
Summary of the State Equations  List the state equations the power variables to be eliminated, the state variables (in
so that the derivatives of the state variables are in the same the order of the state vector), and input on the right.
order as the state variables in state vector x. This is the form
we will use for the vector-matrix notation. v = −v12 − v3g + v
23
 Input
Desired State
Unknown Known
Variable
 diL R 1 1 Known
= − iL − v3 g + v
dt L L L
(7.45)
dv3 g 1 The voltage drop across resistor v12 is neither the desired
= iL
dt C output nor a state variable. Eliminate v12 by substitution.
There is only one equation which contains v12, other than the
The state equations are a set of simultaneous first-order dif- compatibility equation. That is the elemental equation for the
ferential equations, because both equations share the state resistor
426 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

v = − RiR − v3 g + v 7.6.5  Vector-Matrix Form of the State Equations


23
 Input
Desired State
Unknown Known
Variable
Known The general vector-matrix form of the state equations is

The current through the resistor is neither the desired output  dx


= Ax + Bu (7.46)
nor a state variable. Eliminate it by substitution. Use the con- dt
tinuity equation for node 2
x is the vector of state variables, a.k.a. the state vector,
v = − R iL − v3 g + v
23
 Input  iL 
Desired State State
Unknown Variable Variable Known
x= 
v3 g 
Known Known

We have eliminated every variable except the input, the state Matrix A consists of expressions of the elemental parame-
variables, and the desired output variable. Consequently, this ters, known as the “coefficient” matrix, because its elements
is the output equation. are the coefficients of the state variables in the state equa-
The standard form for output equations is similar to that tions. Vector (or matrix) B is the “input” matrix, since it is
of the state equations, with one crucial difference. Output multiplied by the vector of inputs, u. In this example, we
equations must be purely algebraic. There cannot be a de- have just one input, so u is a single function, the input v(t ),
rivative in an output equation! The state variables must be not a vector of functions.
in the same order in the output equation, as they are in state
vector x. The input variable, if present, must be the last term u = v (t )
on the right side.
Make sure that state variables are in the same order in State equations for the RLC system are
the output equations, as they are in the state equations. If the
state variables are not in the same order, then the two vectors diL R 1 1
= − iL − v3 g + v
are not the same state vector. The order of the output equa- dt L L L
tions themselves is arbitrary, unless an output vector y was dv3 g 1
specified. = iL
dt C
We will now formulate the output equations for the re-
maining unknown power variables in the system, iC , iR , the which are written in the matrix form by reversing the pro-
current i from the source, and v12. cess of vector-matrix multiplication. Coefficient matrix A is
Input Variable v, State Variables iL , v3 g always square, with as many rows and columns, as there are
The output equations for iC , iR , and i are trivial, since all state variables.
of the currents are equal to the current through the inductor, Check the order of state equations, and the order of terms
which is a state variable. in the state equations are the same as the order of state vari-
ables in the state vector, before transcribing the state equa-
iC = iL tions into vector-matrix form. It reduces errors to write state
iR = iL vector x on both sides of the equation first
i = iL
dx d  iL 
= Ax + Bu ⇔  
The output equation for the voltage drop across the resistor dt dt v3 g 
starts with the elemental equation for the resistor. One sub-
   iL   
stitution is needed to eliminate iR. =   +  v
v
   3 g   
v12 = RiR → v12 = RiL
before you enter the elements of coefficient matrix A and
Summary of the output equations for the unknown power input matrix B. Missing terms correspond to elements with a
variables coefficient of zero.
iC = iL dx d  iL 
= Ax + Bu ⇔  
iR = iL dt dt v3 g 
i = iL
 R 1
v12 = RiL − L −  i 1
L  L   
=  + L v
v23 = − RiL − v3 g + v  1 v  
0   3g   0 
 C 
7.6  State-Space Representation of Dynamic Systems 427

7.6.6 Vector-Matrix Form of the Output dx


= Ax + Bu State Equations
Equations dt
y = Cx + Du Output Equations
The general matrix form of the output equations is
State equations are simultaneous first-order differential
(7.47)
y = Cx + Du equations. Output equations are simultaneous algebraic
equations, not differential equations.
where y is the vector of output variables, x is the vector of What can be lost in the convenience of vector-matrix
state variables, and u is the vector of inputs. Matrix C is shorthand notation is the “shape” of each of the matrices,
the output matrix, which has as many rows as there are out- A, B, C, and D. There are as many state equations as there
put variables (or output equations), and as many columns as are state variables. Consequently, coefficient matrix A is a
there are state variables. Matrix D has the peculiar name of square matrix with as many rows as columns. It is possible
“direct pass-through” matrix. The effect of D is to scale the to formulate state equations to calculate the response of sys-
value of the input and add the scaled input value to the output tems without an input. Such a system would respond to en-
variables without “passing through” the state equations. ergy in the system at time, t = 0+ . Alternatively, it is possible
The output equations for the RLC circuit are to have as many or more inputs than there are state variables.
In the case of no inputs, input matrix B is either a null matrix,
iC = iL a matrix filled with elements equal to zero, or simply absent.
iR = iL If there are as many inputs as state variables, then matrix B
i = iL is square. If there are more inputs than state variables, then
there are at least two input matrices, B1 and B2.
v12 = RiL
Similarly, there may be only one output equation, or more
v23 = − RiL − v3 g + v output equations than there are state variables. Consequently,
output matrix C may be a row matrix or a rectangular matrix.
Again, write state vector x and output vector y first. The pass-through matrix D may be absent or, if it exists, it
may have as many rows as there are output variables yi.
    State variables allow for the calculation of a physical state
 iC     
i      system at any moment in time. State equations are first-order
 R    iL    differential equations, since state vector x is differentiated
y = Cx + Du ⇔  i =  v  +  v
   once with respect to time. In the general case, a state equa-
  3g   
 v12      tion relates the rate of change of a state variable, xi, to the
 v23      current values the state variables, x, and the inputs u. Since,
    in the general case, the rate of change of a state variable is
dependent upon the values of all state variables, all state
Then, enter the elements of output matrix C and the direct equations must be solved simultaneously, to determine the
pass-through matrix D to reduce error. future state of the system.
We will solve state equations for the values of state vari-
 iC   1 0 0 able vector x at discrete instances in time, using a numerical
i   1 0 0 algorithm, the Runge–Kutta method. The syntax of engineer-
 R   iL   
 i = 1 ing mathematical software, e.g., Mathcad and MATLAB, re-
y = Cx + Du ⇔ 0    + 0 v
    v3 g   quires that state and output equations be expressed in vector-
 v12   R 0    0 matrix form. The numerical solution of the state equations
v23   − R −1 1  produces the state variable’s history over the duration of the
solution. The result of the Runge–Kutta method is presented
Matrix C is the “output” matrix. Matrix D is the “pass- as an array, where the first column contains the instances in
through” matrix, so named because the product Du contrib- time at which the solution was calculated, and the remaining
utes to the output vector, independent of the state equations. columns are the values of the state variables for each time
step in the solution. The values of the state variables are then
used to solve the output equations at the same instants of
7.6.7 Numerical Solution of the State time. The “temporal” sequences of the state variables, that
and Output Equations is, their values over time, are then used to evaluate the output
variables. If we were analyzing a mechanical system, and we
Note that state equations and output equations are different were interested in either displacement or acceleration (not
types of vector-matrix equations. the state variable, velocity), we can easily integrate or differ-
entiate the temporal sequences numerically. Numerical inte-
428 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. 7.7   Schematic of a third-


Bearing Damping b 2
order rotational system driven by
a velocity source Pinion Inertia J2
N1 teeth

Compliant Shaft
Spring Constant K

Inertia J1

Fluid Coupling Gear


Damping b1 N 2 teeth

Ω(t)
Angular velocity source

gration is both easy to program in MATLAB and accurate. Table 7.2   Parameter values
Differentiation, although easy to program, is less accurate, as 30 N ⋅ m ⋅ sec N⋅m
Damping b1 b1 = Spring Constant K K =5
we shall see in the next chapter. rad rad
4 N ⋅ m ⋅ sec Number of Pinion
Damping b2 b2 = N1 = 16
rad Teeth, N1
7.7 Example Derivations of State and Output Number of Gear
Equations Inertia J1 J1 = 1kg ⋅ m 2 N 2 = 32
Teeth, N2

Inertia J2 J 2 = 6 kg ⋅ m 2
7.7.1 Third-Order Dynamic System Example:
A Rotational Mechanical System

A rotational mechanical system is shown schematically in The power flows through the two branches of the transform-
Fig. 7.7. The system consists of an angular velocity source er interface sum to zero.
acting on a fluid clutch with damping b1. The clutch drives
a rotational inertia J1, which is attached to a compliant shaft Ppinion + PGear = 0 → TP Ω 3 g + TG Ω 4 g = 0
with spring constant K. The shaft drives the pinion of a gear  N1  N1
set. The output shaft of the gear drives inertia J2, and is sup- Ω3 g + TG  − Ω3 g  = 0 → TP = TG
 N2  N2
ported on bearings with damping b2. Derive the state equa-
tions and output equations for this system, where the output Energetic Equations
variables are acting to accelerate rotational inertias J1 and J2,
using the parameter values in Table 7.2. Continuity:
First, find the nodes of distinct values of the across vari- 
T = Tb1 Tb1 = TJ1 + TK TK = TP − TG − TJ 2 − Tb2 = 0
able, angular velocity, and draw the linear graph.
Derive the transducer equations for the pinion and gear. Compatibility:
Use the geometry of the motion to derive the first transform-
er equation, then sum power flows into the transformer to Ω1g = Ω12 + Ω 2 g Ω 2 g = Ω 23 + Ω 3 g Ω 4 g = Ω 4 g
derive the second transformer equation. The pinion and gear Elements:
must have the same translational velocity where they mesh.
Use the tooth counts as a measure of the circumferences. Re- dTK
Tb1 = b1Ω 12 Tb 2 = b 2 Ω 4 g = K Ω 23
member, the pinion and gear counter-rotate. dt
d Ω 2g d Ω 4g
rpinion Ω 3 g = vmesh = rgear Ω 4 g TJ1 = J1 TJ 2 = J 2
dt dt
where
Energy:
rpinion ∝ N1 and rgear ∝ N 2 E sys = E J + E K + E J
1 2

N 1 1 TK2
N1Ω 3 g = − N 2 Ω4 g → Ω4 g = − 1 Ω 3 g E J1 = J1Ω 22 g EJ = J 2 Ω 22 g EK =
N2 2 2
2 2K
7.7  Example Derivations of State and Output Equations 429

Fig. 7.8   Schematic of a third- Bearing Damping b 2


order rotational system showing
nodes of distinct values of the Pinion Inertia J2
across variable rotational velocity N1 teeth

Compliant Shaft
3
Spring Constant K g
4
Inertia J1

Fluid Coupling Gear


Damping b1 N 2 teeth
2
1
Ω(t)
Angular velocity source
g

Reduce the equation list to state and output equations. De- Second State Equation
fine the state variable before beginning the reduction to limit
dTK dTK
errors:
dt
= K Ω 23 →
dt
(
= K Ω 2 g − Ω3 g )
Ω 2 g 
  dTK  N 
x =  TK  = K  Ω 2 g + 2 Ω4 g 
Ω 4 g  dt  N1 
 
dTK N
Derive the state equations by starting with the elemental = K Ω 2g + K 2 Ω 4g State Equation 2
equation with the derivative of corresponding state variable. dt N1
Move the derivative to the left side, and clear its coefficient.
Proceed to eliminate all variables from the right side, which Third State Equation
are neither state variables nor the input variable. The right
side must be purely algebraic. No derivatives appear on the d Ω 4g d Ω 4g 1
TJ 2 = J 2 → = TJ
right side. dt dt J2 2

d Ω 4g 1 d Ω 4g 1
State Equations:
Input Ω 1g ≡ Ω   State variables Ω 2 g , T K , Ω 4 g dt
=
J2
(
−TG − Tb2 ) →
dt
=
J2
(
−TK − b 2 Ω 4 g )
d Ω 4g 1 b2
First State Equation =− TK − Ω 4 g State Equation 3
dt J2 J2
d Ω 2g d Ω 2g 1
TJ1 = J1 → = TJ Summary of State Equations
dt dt J1 1

d Ω 2g 1 d Ω 2g 1 d Ω 2g b1 b1
dt
= (
Tb − TK
J1 1
) →
dt
=
J1
(
b1Ω 12 − TK ) dt
=−
J1
Ω 2g −
1
J1
TK + Ω
J1
d Ω 2g dTK N
dt
=
1
J1
( (
b1 Ω − Ω 2 g − TK ) ) dt
= K Ω 2 g + K 2 Ω 3g
N1
d Ω 4g 1 b2
d Ω 2g b1 1 b1 =− TK − Ω 4 g
=− Ω 2g − TK + Ω State Equation 1 dt J2 J2
dt J1 J1 J1
430 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. 7.9   Linear graph of the


-N1
Ω 4g= ___ Ω 3g
third-order rotational system of
Fig. 7.8 b1 K N2
1 2 3 4

Ω(t) J1 TP TG J2 b2

g g g g

Write the state equations in vector-matrix form. Write the Output Equation for TJ 2
state vectors first.
−TG − TJ 2 − Tb 2 = 0 → TJ 2 = −TG − Tb2
dx
= Ax + Bu N2
dt TJ 2 = − TP − b 2 Ω 4 g
N1
Ω 2 g    Ω 2 g    N2
d        TJ 2 = − TK − b2 Ω4 g Output Equation 2
TK  =    TK  +  Ω N1

dt     Ω   
 Ω 4 g     4 g    Summary of output equations

Then write the elements of coefficient matrix A and input


TJ1 = −b1Ω2 g − TK + b1Ω
matrix B.
N2
TJ 2 = − TK − b2 Ω4 g
 b1 1  N1
− − 0 
Ω 2 g   1
J J1  Ω   b1 
 2g   Write the output equations in vector-matrix form. First, write
d   N2    J
 TK  =  K 0 K   TK  +  1  Ω output vector y and state vector x.
dt  N1    
   0
Ω 4 g   b2  
Ω 4 g  0
 0 1   
− − Ω 
 J2 
J2  TJ1     2g   
y = Cx + Du →  =   TK  +   Ω
TJ 2    Ω   
Output Equations:  4g 

Input Ω 1g ≡ Ω , State Variables Ω 2 g , TK , Ω 4 g, Then enter the components of output matrix C and the pass-
Output Variables TJ1, TJ 2. through matrix D.

Output Equation for TJ1  −b −1 b1  Ω 2 g 


TJ1   1   T  +  b1  Ω t
Tb1 = TJ1 + TK → TJ1 = Tb1 − TK
 = N
− 2 −b 2   K   0  1g
()
TJ 2   0 N1  Ω 4 g 

TJ1 = b1Ω12 − TK (
→ TJ1 = b1 Ω − Ω 2 g − TK )
7.7.2 Fourth-Order Dynamic System Example:
TJ1 = −b1Ω 2 g − TK + b1Ω Output Equation 1 A Spring-Mass-Damper System

The translational mechanical system shown in the schematic


Fig. 7.10 is fourth order, since the two springs and two mass-
es can have an arbitrary amount of energy stored in them,
as an initial condition for the system. There are four state
variables: (1) the forces acting through springs K1 and K2
and (2) the velocities of masses M1 and M2, Table 7.3. From
7.7  Example Derivations of State and Output Equations 431

x,v x,v
K1 b K1 b
2 g
1
M1 M2 M1 M2
F(t) K2 F(t) K2
g
g

Fig. 7.10   Schematic of a fourth-order translational system driven by Fig. 7.11   Schematic of the translational system, annotated with nodes
a force source of distinct values of the across variable, translational velocity

the Newtonian perspective, this mechanical system is classi- 2. The state variable terms appear in the order of the
fied as having two degrees of freedom, since each mass can state vector,
displace independently in the x direction. Note that the posi- 3. Constants multiplying state variables are distributed,
tive x direction is indicated on mass M2 to correspond to the 4. The input term is last.
direction of input force F(t). We will decide the order of the state variables before begin-
The first step of our analysis is independent of the size or ning the reduction, in order to express each state equation in
type of the system; whether we will represent the system as the form we can transcribe directly into vector-matrix nota-
a single higher-order differential equation; or in state-space tion. The order of the state variables in the state vector is
as a set of simultaneous first-order differential equations. We arbitrary but once defined must be respected. We will define
must find the nodes of distinct values of the across variable, the state vector as:
translational velocity, Fig. 7.11, so that we can draw the lin-
ear graph, Fig. 7.12.  v1g 
v 
x= 
2g
State Vector:
Energetic Equations  FK1 
 
 FK2 
Continuity: − FM1 − FK1 = 0 F + FK1 − FM 2 − FK2 − Fb = 0

Compatibility: v1g = v12 + v2 g v2 g = v2 g Input F, State Variables v1g , v2 g , FK1, FK2


dv1g dv2 g
Element: FM1 = M 1 FM 2 = M 2 dv1g dv1g 1
dt dt (1) FM1 = M 1 → = FM
dt dt M1 1
dFK1 dFK2 dv1g 1
= K1v12 = K 2 v2 g Fb = b v2 g =− FK State Equation1
dt dt dt M1 1
Energy: E Sys = E M1 + E M 2 + E K1 + E K2
dv2 g dv2 g 1
(2) FM 2 = M 2 → = ( F + FK1 − FK2 − Fb )
1 1 FK21 FK22 dt dt M2
EM = M 1v12g E M 2 = M 2 v22g E K1 = E K2 =
1
2 2 2 K1 2K2 dv2 g 1
= ( F + FK1 − FK2 − bv2 g )
dt M2
Reduction to State Equations dv2 g b 1 1 1
=− v2 g + FK − FK + F
Reminders: dt M2 M2 1 M2 2 M2
• The derivation of a state equation begins with the energy State Equation 2
storage element equation with the derivative of that state
variable. dFK1 dFK1
(3) = K1v12 → = K1 (v1g − v2 g )
• Only the derivative of the state variable without a coef- dt dt
ficient may be on the left side of a state equation. dFK1
• Eliminate by substitution all power variables on the right = K1v1g − K1v2 g State Equation 3
dt
side, except the state variables and the input variable.
• Once elimination by substitution is complete, then write dFK2
the state equation in standard form on the right side where: (4) = K 2 v2 g State Equation 4
dt
1. The state variables are collected,
432 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Table 7.3   Power variables in the fourth-order mechanical system


K1
1 2 Across variable Through variable
Force Source v 1g (state variable) Known
Mass M1 v 1g (state variable) FM1
M1 F(t) b K2 M2 Mass M2 v 2 g (state variable) FM 2
Spring K1 v12 FK1 (state variable)
Spring K2 v 2 g (state variable) FK 2 (state variable)
g g Damper b v 2 g (state variable) Fb
Fig. 7.12   Linear graph of the fourth-order translational system

Summary of the State Equations ate the mass M1 and M2, FM1 and FM 2 ;, the force acting through
the damper, Fb ,; and the velocity drop across the spring, v12 .
dv1g 1
=− FK The output variable equations are always purely algebraic.
dt M1 1 There can be no time derivatives in the output equations.
dv2 g b 1 1 1 We will use the state variable vector and input variable to
=− v2 g + FK1 − FK2 + F calculate the output variables. Therefore, the standard form
dt M2 M2 M2 M2
of the output equation places the output variable on the left
dFK1
= K1v1g − K1v2 g side without a coefficient, and these conditions apply to the
dt right side.
dFK2 • The state variables are collected,
= K 2 v2 g
dt • The state variable terms appear in the order of the state
vector,
Express the state equations in matrix form • Constants multiplying state variables are distributed, and
• The input term is last.
dx Derive the output equations for FM1, FM 2, v12, Fb
= Ax + Bu
dt (1) Output Equation for FM1
− FM1 − FK1 = 0 → FM1 = − FK1
 v1g     v1g   
v       (2) Output Equation for FM 2
d  2g     v2 g   
=  F  +  F
dt  FK1     K1   
  F + FK1 − FM 2 − FK2 − Fb = 0
  F   
 FK2     K2   
FM 2 = F + FK1 − FK2 − Fb

 1  FM 2 = F + FK1 − FK2 − bv2 g


 0 0 − 0   0 
v
 1g  M1 v 
     1g   1  FM 2 = −bv2 g + FK1 − FK2 + F
d  v2 g   b 1 1   v2 g   
=0 − −  + M2  F
dt  FK1   M2 M2 M 2   FK1   
    0  (3) Output Equation for v12

 FK2   K1 − K1 0 0   FK2  
  0 
K2 v1g = v12 + v2 g → v12 = v1g − v2 g
0 0 0 
(4) Output Equation for Fb Fb = b v2 g
Output Equations
The state equations are solved, yielding the values of state Summary of the output equations for FM1, FM 2, v12, Fb
variables over the period of interest, before the output vari-
FM1 = − FK1
ables are solved. The state variables and input variable are
then used to formulate the output variables equations. In this FM 2 = −bv2 g + FK1 − FK2 + F
system, after the state variables are known, the only remain- v12 = v1g − v2 g
ing unknown power variables are the forces acting to acceler-
Fb = bv2 g
7.7  Example Derivations of State and Output Equations 433

Fig. 7.13   Schematic of a hybrid Hydraulic


rotational mechanical-fluid- Piston
translational mechanical system Area A
driven by a torque source
C
T(t) R2 I
J Pump M
b R1

Express the output equations in matrix form. We relate volume Vol displaced per revolution to the volume
flow rate from the pump QP, and the angular velocity of the
 FM1     v1g    shaft driving the pump Ω 1g .
      
 FM 2     v2 g   
y = Cx + Du →  =  + F Vol
  FK1   QP = Ω 1g
 v12    2π
 Fb    F   
   K2   
The second transducer equation for the gear pump was de-
rived by setting the sum of the two power flows into the gear
 FM1  0 0 −1 0   v1g  0 pump as equal to zero. These power flows are the products of
   
 FM 2  0 −b 1 −1  v2 g  1  the through and across variables of the two branches, which
y = Cx + Du → =   + F
 v  1 −1 0 0   FK1  0 form the interface of the linear graph transducer symbol.
 12      
 Fb  0 b 0 0   FK2  0 Ppump + Ppump = 0 → TP Ω 1g + QP p2 g = 0
mech fluid

7.7.3 Fourth-Order Dynamic System Example: We then eliminated the volume flow rate of the pump QP by
A Fluid-Mechanical System substitution

A fourth-order fluid-mechanical system of Sect. 6.4.4.3 is Vol Vol


TP Ω 1g + Ω 1 g p2 g = 0 → TP Ω 1g + Ω 1g p2 g = 0
shown schematically in Fig. 7.13. This is a hybrid system 2π 2π
with three different energy “domains”; rotational mechani-
cal, fluid, and translational mechanical. The gear pump and to yield the second transducer equation of the pump
the hydraulic piston are transducers which interface the dis-
similar energy domains. Vol
TP = − p2 g
The only difference in the process of expressing a system 2π
with transducers or transformers in state-space from that of
a system with a single energy domain is that the transducer The transducer equation for the hydraulic piston cylinder
or transformer equations must be derived and included in the was derived by calculating the net pressure force acting on
equation list. The first step, as always, is to find the nodes of the fluid-side face of the piston, and equating that to the
distinct values of the across variables. force acting on the mechanical-side face.
Draw the linear graph by first drawing the interfaces for
the two transducers, and then adding the remaining elements FA = Ap4 g
between their respective nodes. The two branches of an in-
terface represent only the transformation or transduction of The second transducer equation for the hydraulic piston cyl-
power. Do not assign them properties of energy storage or inder was derived, by summing the power flows into the hy-
dissipation. draulic piston
Recall we derive the transducer equations for a shaft-driv-
en positive displacement pumps, using the fact that volume Ppiston + Ppiston = 0 → QA p4 g + FA v5 g = 0
Vol of fluid was displaced with each revolution of the shaft. fluid mech
434 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. 7.14   Schematic of a hybrid Hydraulic


rotational mechanical-fluid- Piston
translational mechanical system, Area A
annotated to show distinct C
1 4
2 R2 3 I
locations of the across variables: 5
rotational velocity, pressure, and T(t)
translational velocity J Pump M
g
b R1
g g g

and then substituting in the relationship for FA: Identify state variables of the system. The state variables are
the energy storage variables. Define the order of state vari-
QA p4 g + A p4 g v5 g = 0 → QA p4 g + A p4 g v5 g = 0 ables in the state vector.

QA = − Av5 g .
 Ω 1g 
Q 
The linear graph sign convention should be respected during
x=
I 

the derivation of the state equations. The signs can then be  p4 g 


fixed, so that a positive input yields a positive output.  
 v5 g 
Energetic Equations
Reduction to State Equations
Continuity:   T − TJ − TP = 0 − QP − QR 1 − QR 2 = 0 Input T State Variables Ω1g , QI , p4 g , and v5 g

QR 2 − QI = 0 QI − Q C − QA = 0 − FA − FM = 0 d Ω 1g d Ω 1g 1
(1) TJ = J → = TJ
dt dt J
Compatibility:   Ω1g = Ω1g p2 g = p23 + p34 + p4 g
d Ω 1g 1 d Ω 1g 1 1  Vol 
dt
=
J
(T − TP ) →
dt
= T+ 
J J  2π 
 p2 g
p4 g = p4 g v5 g = v5 g

d Ω 1g 1  Vol  1
Element:   QP =
Vol
Ω 1g TP = −
Vol
p2 g
=   R1QR 1 + T
dt J 2π J
2π 2π

( )
d Ω 1g d Ω 1g 1  Vol  1
FA = Ap4 g QA = − Av5 g TJ = J =   R1 −QP − QR 2 + T
dt dt J  2π  J
dQI
p2 g = R1QR1 p23 = R 2 QR 2 p34 = I d Ω 1g 1  Vol  1  Vol  1
dt =−   R1QP −   R1QR 2 + T
dt J  2π  J 2π  J
dp4 g dv5 g
QC = C FM = M
dt dt d Ω 1g 1  Vol   Vol  1  Vol  1
=   R1   Ω1g −   R1QI + T
dt J 2π 2π J 2π J
Energy:   E sys = E J + E I + EC + E M
d Ω 1g R1  Vol  2 R1  Vol  1
1 1 2 1 1 =   Ω1g −   QI + T
EJ = J Ω12g EI = IQI EC = C p42g EM = Mv52g dt J 2π J  2π  J
2 2 2 2
State Equation 1
7.7  Example Derivations of State and Output Equations 435

Fig. 7.15   Linear graph of the Vol


hybrid rotational mechanical- Q P = ___ Ω 1g R2 I FA= Ap4g
fluid-translational mechanical 2π
system shown in Fig. 7.14 1 2 3 4 5

T(t) J TP QP R1 C QA FA M

g g

dQI dQI 1 dv5 g dv5 g 1 dv5 g 1


(2) p34 = I → = p34 (4) FM = M → = FM → =− FA
dt dt I dt dt M dt M
dv5 g A
dQI 1
dt
=
I
(
p2 g − p23 − p4 g ) dt
=−
M
p4 g State Equation 4

dQI 1
dt
(
= R1QR 1 − R 2 QR 2 − p4 g
I
) Summary of the State Equations

R1 R2 d Ω 1g R1  Vol  2 R1  Vol  1
dQI 1 =   Ω 1g −   QI + T
= QR 1 − QR 2 − p4 g dt J 2π J 2π J
dt I I I
dQI R1 Vol  R1 + R 2  1
( )
dQI R1 R2 1 =− Ω 1g −   QI − p4 g
= −QP − QR 2 − QI − p4 g dt I 2π  I  I
dt I I I
dp4 g1 A
dQI R1  Vol  R2 1 QI + v5 g
=
=  − Ω 1g − QI  − QI − p4 g dt C C
dt I 2π  I I dv5 g A
=− p4 g
dQI R1 Vol  R1 + R 2  1 dt M
=− Ω 1g −   QI − p4 g
dt I 2π  I  I
dx
State Equation 2 Express the state equations in matrix form, = Ax + Bu
dt

dp4 g dp4 g 1
(3) QC = C → = QC
dt dt C

dp4 g 1 dp4 g 1
dt
=
C
( QI − QA ) →
dt
=
C
QI + Av5 g ( )
dp4 g 1 A
= QI + v5 g State Equation 3
dt C C

 R1  Vol  2 R1  Vol  
   −   0 0
 J  2π  J  2π   1

 1g   R Vol   Ω 1g   
 R + R2  1
 
d  QI   − I 2π
1
− 1 − 0  Q  J 
=  I  I   I  +  0 T
dt  p4 g    
   1 A   p4 g   0 
0 0  v  
 v5 g   C C   5 g   0 
 
 A 
 0 0 − 0
 M 
436 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Table 7.4   Power variables in Across variable Through variable


the fourth-order fluid-mechanical
System Torque source Ω
1g (state variable) T (input)
Inertia J Ω
1g (state variable) TJ
Mechanical side of gear pump Ω
1g (state variable) TP
Fluid side of gear pump p2 g QP
Fluid resistance R1 p2 g QR1
Fluid resistance R2 p23 QR 2
Fluid inertance I p34 Q
I (state variable)
Fluid capacitance C p4 g (state variable) QC
Fluid side of piston cylinder p4 g (state variable) QA
Mechanical side of piston cylinder v 5g (state variable) FA
Mass M v5g (state variable) FM

(3) Output Equation  for p2 g:


Output Equations
Vol 2π
The power variables of this system are shown in Table 7.4. TP = − p2 g → p2 g = −TP
Remember, the output equations are algebraic equations, in 2π Vol
terms of the input and state variables. Do not include deriva-   Vol  2  Vol   2π
p2 g = −    R 1Ω 1 g +   R1QR 2 
tives on either side of any output equation.   2π   2π   Vol
(1)  Output Equation  for TJ :
T − TJ − TP = 0 → TJ = −TP + T  Vol 
p2 g = −  R Ω − R1QR 2
 2π  1 1g
Vol Vol
TJ = p2 g + T → TJ = R1QR1 + T
2π 2π   (4) Output Equation for QR1:
p2 g 1   Vol  
 Vol   Vol  QR1 = → QR1 =  −  R1Ω 1g − R1QI 
TJ =  R Q +T → TJ =  R ( −QP − QR2 ) + T R1 R1   2π 
 2π  1 R1  2π  1
 Vol 
 Vol   Vol  QR1 = −  Ω − QI
TJ = −  RQ − R Q +T  2π  1g
 2π  1 P  2π  1 R 2
  (5) Output Equation for QR2: QR2 = QI
 Vol   Vol   Vol    (6) Output Equation for p23: p23 = R 2 QR2 → p23 = R 2 QI
TJ = −  R Ω − R Q +T
 2π  1  2π  1g  2π  1 R2   (7) Output Equation for p34:
2 p2 g = p23 + p34 + p4 g → p34 = p2 g − p23 − p4 g
 Vol   Vol 
TJ = −  RΩ − R Q +T V 
 2π  1 1g  2π  1 R 2 p34 = −   R1Ω 1g − ( R1 + R2 ) QI − p4 g
 2π 
(2)  Output Equation  for TP:   (8) Output Equation for QC:
T − TJ − TP = 0 → TP = −TJ + T QI − QC − QA = 0 → QC = QI − QA
  Vol  2
 Vol   QC = QI + Av5 g
TP = −  −   R1Ω 1g −   R1QR2 + T  + T
  2π   2π  
2   (9)  Output Equation for FA: FA = Ap4 g
 Vol   Vol  (10)  Output Equation for FM:
TP =  RΩ + RQ
 2π  1 1g  2π  1 R2
− FA − FM = 0 → FM = − FA → FM = − A p4 g
7.8  Why “State-Space” is called “State-Space” 437

Summary of the output equations geometric space, since the force axes can be aligned with the
2
position axes.
 Vol   Vol  A two-dimensional geometric “space,” i.e., a plane, is
TJ = −  RΩ − R Q +T
 2π  1 1g  2π  1 I formed by defining two coordinate axes, a three-dimensional
 Vol 
2
 Vol  space by defining three coordinate axes, a four-dimensional
TP =  RΩ + RQ
 2π  1 1g  2π  1 I space by defining four coordinate axes, etc. Although we
cannot visualize spaces above three dimensions and call such
 Vol  spaces “hyperspace,” which means “beyond space,” hyper-
p2 g = −  R Ω − R1QI
 2π  1 1g space follows the same mathematical rules as does two- and
 Vol  three-dimensional spaces. You can define a space to have as
QR1 = −  Ω − QI
 2π  1g many dimensions as you want. The only restriction is that
QR2 = QI its axes must be independent. Axes are independent, if you
cannot plot a point on one axis using the other axes, except at
p23 = R 2 QI the origin. In the most fundamental sense, a space is defined
 Vol  by its coordinate coordinates. If we define coordinates, then
p34 = −  R Ω − ( R1 + R2 ) QI − p4 g
 2π  1 1g we define a space. The axes of the space do not need to have
QC = QI + Av5 g the same units. We often use two-dimensional spaces with
time or position as the horizontal axis and a different physi-
FA = A p4 g
cal quantity as the vertical axis.
FM = − A p4 g A “state-space” is a space defined by coordinates, which
are the state variables of a system. It is neither a “real”
Output equations in matrix form, y = Cx + Dy

  Vol 
2
 Vol  
 − R − R 0 0 
  2π  1  2π  
 2 
 TJ    Vol   Vol  1 
  R1   R1 0 0 
T   2π  2π   0
 P     
 p2 g    Vol 
   −  2π  R1 − R1 0 0  0
  
QR1     Ω 1g   0
    Vol   
QR2  =  −  2π  −1 0 0   QI  0
 +  T
p     p3 g  0
 23   0 1 0 0    
 p34     v4 g  0
Q   0 R2 0 0
 0
 C    Vol    
 FA   −   R1 − (R + R2 ) −1 0  0
   2π 
1
  0
 FM   0 1 0 A 
 
 0 0 A 0 
 0 0 −A 0 

7.8  Why “State-Space” is called “State-Space” space in a geometric sense, but nor is a force space. The en-
ergy storage variables satisfy the mathematical requirement
“State-Space” is a very impressive term for a simple concept. for axes, because they are independent. The values of the
Mathematically, a space is defined by the coordinate axes, state variables at any moment in time are the coordinates of
which allow you to plot a vector in that space. All planes a point (a vector from the origin) in the state-space of that
and spaces you have worked with in engineering are abstrac- dynamic system. We can, and do, define hyper state-spaces.
tions, with the exception of purely geometric spaces used A dynamic system with four independent energy storage
to draw parts. You used force spaces in statics to perform elements has four state variables. Consequently, its state-
equilibrium analysis. The coordinates of a force space are the space is four dimensional, because its state vector has four
force components, Fx, Fy, and Fz. Although an abstraction, components. Is it easier to work with three-dimensional
it is easy to visualize a force vector in a three-dimensional state-spaces than four-dimensional state-spaces? Not re-
438 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

ally. It makes little difference how many dimensions there State variables are defined recursively as the output variable
are in the state-space in practice, because our visualization and its derivatives on the right side.
will be two-dimensional plots of individual state variables
x1 ≡ y
or output variables vs. time. Our major concern with high- 
er-order systems will be with the accuracy of our model, dy
x2 ≡
which will degrade with increased complexity of higher- dt
order systems. d2y (7.51)
x3 ≡ 2
dt
d3y
7.9 Expression of Systems Equations x4 ≡ 3
dt
in State-Space
All but one state equation is formed from the definitions of
The set of energy storage variables will be our preferred the state variables, by noting the relationships between the
choice as the state variable of a system, when we derive the state variables and the derivatives of the output variable.
state equations starting with a schematic or linear graph. En-
ergy storage variables have physical meaning. The derivation  x1 ≡ y
of state equations from the equation list is straightforward. dy dx1
However, other sets of variables can be chosen as the state x2 ≡ =
dt dt
variables of a system.
We will now develop an algorithm to express higher-or- d y dx
2
(7.52)
x3 ≡ 2 = 2
der system equations and transfer functions as sets of state dt dt
equations. The technique is straightforward if the input is not d y dx
3
x4 ≡ 3 = 3
differentiated. Systems with derivatives of input variables dt dt
require a more complicated algorithm.
The final state equation is the equation

7.9.1 Algorithm to Express a Higher-Order d 4 y  a  d 3 y  a  d 2 y  a  dy  a4  b 


= − 1  3 − 2  2 − 3  − y+ 4u
System Equation Without Differentiation dt 4
 a0  dt  a0  dt  a0  dt  a0   a0 
of the Input as State Equations
(7.49)
If the input is not differentiated, then a set of state equa- noting that
tions is defined as the output variable and its n-1 derivatives,
where n is the order of the differential equation. We will use  dx4 d 4 y
= 4 (7.53)
a fourth-order system equation for illustration, where the dt dt
input variable is u (t ), and the output variable is y (t ).
and using the state variables to eliminate output variable y
 d4y d3y d2y dy and its derivatives on the right side
a0 4 + a1 3 + a2 2 + a3 + a4 y = b4 u (7.48)
dt dt dt dt
 dx4 a  a  a  a  b 
= −  1  x4 −  2  x3 −  3  x2 −  4  x1 +  4  u
Standard form for higher-order system equations is to clear dt  a0   a0   a0   a0   a0 
the highest-order derivative of the output variable of any co-
efficient term. (7.54)
Summarizing the state equations
 d 4 y  a1  d 3 y  a2  d 2 y  a3  dy  a4  b 
+  3 +  2 +  +  y =  4u  dx1
dt 4
 a0  dt  a0  dt  a0  dt  a0   a0  = x2
dt
(7.49) dx2
= x3
dt
Isolate (solve for) the highest-order derivative of the output dx3
= x4
variable. dt
dx4 a  a  a  a  b 
d 4 y  a1  d 3 y  a2  d 2 y  a3  dy  a4  b  = −  1  x4 −  2  x3 −  3  x2 −  4  x1 +  4  u
= − −  2 −  −  y+ 4 u dt  a0   a0   a0   a0   a0 
dt 4  
 a0  dt 3
 a0  dt  a0  dt  a0   a0 

(7.50) (7.55)
7.9  Expression of Systems Equations in State-Space 439

R L Rewriting this equation using x1 for the output variable v3 g , u


1 2 3 for the input variable v1g , and α 1, α 2, and β for the constant
terms

 d  dx1  dx1
v(t) C   = α1 x1 + α 2 + βu (7.59)
dt  dt  dt

We see that the equation can be expressed in the form of a


g g state equation by defining

Fig. 7.6   Linear graph of the series RLC circuit of Fig. 7.5  d x1


= x2 (7.60)
dt

dx
Expressing the state equations in matrix form, = Ax + Bu,
dt

 0 1 0 0   0 
 x1     x1   0 
  0 0 1 0  x  
d  x2   
(7.56)
= 0 0 0 1   2 +  0 u
dt  x3     x3   
    a4  a  a   a1      b4  
 x4   −   − 3  − 2  −     x4    
  a0   a0   a0   a0    a0  

and the output equations in matrix form, y = Cx + Du, and substituting

  dx2
 x1  = α1 x1 + α 2 x2 + βu (7.61)
x  dt
y = [1 0 0 0]  
2

 x3  (7.57)
These two equations are a set of state equations.
 
 x4  Applying this method to the system equation of the RLC
circuit,
7.9.1.1 Example Expression of a Higher-Order
System Equation Without Differentiation
of the Input as State Equations d 2 v3 g 1 R dv3 g 1
2
=− v3 g − + v1g
We again use the series RLC circuit as the example system. dt LC L dt LC
The linear graph, Fig. 7.6, and the system equation for the
output variable is the voltage across the capacitor, v3g, from we define the output variable as the first state variable
Sect. 7.6.2 are repeated below for convenience.
The second-order system equation is: v3 g ≡ x1

1 d 2 v3 g R dv3 g 1 and derivative of the output variable as the second state vari-
v1g = + + v3 g
LC dt 2 L dt LC able. It is essential to recognize that this definition also cre-
ates a state equation.
Rearrange the system equation so that the highest-order de-
rivative of the output variable is on the left side. We create dv3 g dx1
≡ x2 =
a summation of three terms: the output variable, its first de- dt dt
rivative, and the input variable, each multiplied by a factor
comprising the system parameters. This is a “recursion,” meaning the next variable is defined
in terms of the previous. The second derivative of the output
 d 2 v3 g 1 R dv3 g 1 variable equals the derivative of the second state variable.
2
=− v3 g − + v1g (7.58)
dt LC L dt LC
d 2 v3 g dx2
2
=
dt dt
440 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Table 7.5   State and output eqs. Energy storage variables as state variables State variables defined recursively from the output vari-
for the RLC circuit of Fig. 7.6 able of the second-order system equation
 R 1  0 1   0 
 − L − L   iL   1  d  x1    x1 
d  iL 
   +  L  v1g   = 1 R    +  1  v1g
 = 
  dt  x2   − −  x2  
dt v3 g   1  v  LC L   LC 
 C 0   3 g   0 
 x 
 iL  v3 g = [1 0]  1 
v3 g = [ 0 1]    x2 
v3 g 

Substituting into the system equation yields 7.9.2 Algorithm to Express a Higher-Order


System Equation with Differentiation
dx2 1 R 1 of the Input as State Equations
=− x1 − x2 + v1g
dt LC L LC
The algorithm to express higher-order system equations with
We now have the set of state equations derivative input terms is slightly more involved. It is based
on the fact that state variables are not unique. We can define
dx1 any number of sets of state variables with algebra. In this
= x2 algorithm, we define a set of state variables consisting of
dt
“weighted” sums of the input and output variables, where
dx2 1 R 1
=− x1 − x2 + v1g the term weighted means each variable is multiplied by a
dt LC L LC different coefficient.
State equations are sets of simultaneous first-order dif-
The output variable of the system is the voltage across the ferential equations, where the only derivative in an equation
capacitor, v3 g . We use our definition of the first state vari- is the state variable on the left-hand side. There can be no
able as an algebraic output equation to express the output in derivatives on the right-hand side, including derivatives of
terms of the state variables the inputs.

v3 g = x1  dx
= Ax + Bu (7.45)
dt
In this particular case, the output equation is just the defi-
nition of state variable. However, in the case of systems in We will see when we develop the transfer function represen-
which the input is differentiated, then state variable x1 and tation of higher-order systems that it is common for the input
output variable will not be the same. variable to appear as one or more derivatives. Derivatives of
Expressing state equations and the output equation in ma- the inputs are dealt with by defining non-physical state vari-
trix form. ables, which are algebraic expressions, typically including
a state variable and a derivative of the input variable. This
dx sounds intimidating, but there is a reasonable algorithm to
= Ax + Bu
dt handle these cases.
The algorithm assumes that both the input and output
 0 1   0  variables are differentiated to the same degree. If this is not
d  x1     x1  
= 1 R   + 1  v1g
dt  x2   − −   x2  
the case, as it typically isn’t, then eliminate unneeded input
 LC L  LC  derivatives, by setting their respective coefficients to zero.
We will use a fourth-order system equation for illustration,
 x1 
y = Cx + Du → v3 g = [1 0]   where the input variable is u (t ), and the output variable is
 x2  y (t ).

When we compare these state and output equations to those d4y d3y d2y dy
a 0 4 + a1 3 + a 2 2 + a 3 + a4 y
we derived using the system’s energy storage variables as the dt dt dt dt
state variables, we see that they are not the same, as shown d 4u d 3u d 2u du
= b 0 4 + b1 3 + b 2 2 + b 3 + b 4u
in Table 7.5. dt dt dt dt
(7.62)
7.9  Expression of Systems Equations in State-Space 441

Again, the standard form for higher-order system equations Define the first state variable, x1 , as the operand of the high-
is to clear the highest-order derivative of the output variable est-order derivative.
of any coefficient term.
  b0 
x1 = y −   u (7.67)
 d y  a 1  d y  a 2  d y  a 3  dy  a 4 
4 3 2
 a0 
+  + +  +  y =
dt 4  a 0  dt 3  a 0  dt 2  a 0  dt  a 0 
The remaining state variables are defined in terms of the de-
 b 0  d 4 u  b1  d 3u  b 2  d 2 u  b 3  du  b 4  rivative of the output variable y, the input variable u, and
  4 +  3 +  2 +  + u
 a 0  dt  a 0  dt  a 0  dt  a 0  dt  a 0  coefficient terms βi

(7.63)  x1 = y − β0 u
dx
x2 = 1 − β1u
The basis of the technique is to define new state variables, dt
which both incorporate the derivatives of the input and allow dx2
x3 = − β2 u
us to write the state equations in the same form, as the case dt
without derivative of the input, i.e., dx
x4 = 3 − β3u (7.68)
dt
  0 1 0 0 
 x1   0 0 1 0   x1   β1  and
      
d  x2   0
= 0 0 1   x2  + β2  u b0
dt  x3     x3   β3   β0 =
    a4   a3   a2   a1       a0
 x4   −   −   −  −     x4  β4 
  a 0   a0   a0   a 0   b1 a1
β1 = − β0
 x1  a0 a0
x  b2 a1 a2
y = [1 0 0 0]   + β0u
2
(7.64) β2 = − β1 − β0
 x3  a0 a0 a0
 
 x4  b3 a1 a2 a3
β3 = − β2 − β1 − β0
a0 a0 a0 a0
First, solve for the highest-order output variable on the left
b4 a1 a2 a3 a4
side. β4 = − β3 − β2 − β1 − β0 (7.69)
a0 a0 a0 a0 a0
 d4y  a 1  d 3 y  a 2  d 2 y  a 3  dy  a 4 
= −  3 −  2 −  −  y
dt 4
 a 0  dt  a 0  dt  a 0  dt  a 0  7.9.2.1 Example Expression of a Second-Order
System Equation with Differentiation of
 b 0  d 4 u  b1  d 3u  b 2  d 2 u the Input as State Equations
+  4 +  3 +  2
 a 0  dt  a 0  dt  a 0  dt We will demonstrate this method with the translational sys-
tem, consisting of a force source acting on the mass attached
 b 3  du  b 4 
+  + u (7.65) to a spring and a damper of Sect. 3.2. The schematic and
 a 0  dt  a 0  linear graph are Fig. 7.16.

Collect the input and output variable terms by order of dif-


ferentiation to guide our choice of state variable.


d4   b0   d 3   a1   b1   d 2   a2   b2   d   a 3   b3    a 4  b 
 y −   u = 3 −   y +   u + 2 −   y +   u + −   y +   u −   y +  4  u
dt 4   a 0   dt   a 0   a 0   dt   a 0   a 0   dt   a 0   a 0    a 0   a0 
(7.66)
442 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. 7.16  a Schematic and b lin-


ear graph of a spring-mass-damp-
a x,v b 1
er system with an input force b
F(t)
M F(t) b K M

K
g

The second-order system equation, which relates the force Instruments calculators use the convention, which identifies
input to the velocity of the mass, Eq. 3.34, has differentiation the coefficient of the zero-order term with the subscript of
of the input term with respect to time. zero.
You must identify which of the two conventions is used,
2
 1 dF (t ) d v1g b dv1g K before applying an algorithm which identifies coefficients
= + + v1g (3.34)
M dt dt 2 M dt M with subscripts. Authors should identify which convention
they are using, but they do not always. If the convention is
The system is of second order. Hence, it will be represented not identified, then you must not use the algorithm. We will
by a set of two state equations of the form identify the convention used by showing a differential equa-
tion with the coefficients identified, thusly:

 0 1 
d  x1     x  β 
  =   a2   a1    1  +  1  u d 2v dv d 2F dF
dt  2   −   −     x2  β2 
x a0 2
+ a 1 + a 2 v = b 0 2
+ b1 + b2 F
dt dt dt dt
  a0   a0  
 x1 
y = [1 0]   + β0 u (7.70) Note that this differential equation is not in standard form,
 x2  since the standard form would have no coefficient on the
highest-order term of the output variable, v, in this case.
where By convention, the coefficients ai are assigned to the out-
put variable terms and the coefficients bi are assigned to the
 b0 input variable terms.
β0 = First, rewrite the system equation in the required form
a0
with the input variable differentiated to the same order as
x1 = y − β0 u
b1 a1 the output variable, adding terms with coefficients of zero,
β1 = − β0 and dx1
a0 a0 x2 = − β1u if needed
dt
b2 a1 a2
β2 = − β1 − β0
a0 a0 a0 d 2 v1g b dv1g K d 2 F 1 dF
+ + v1g = 0 2 + + 0F
dt 2 M dt M dt M dt
(7.71)
Next, equate the coefficients of the system equation with the
There are two conventions for identifying the coefficients in coefficients in the required form
polynomials and differential equations. We will use the con-
vention, which identifies the coefficient of the highest-order a0 = 1 b0 = 0
term with the subscript of zero, a0. The other convention b 1
identifies the coefficient with a subscript equal to the order a1 = and b1 =
M M
of the term, with the constant or zero-order term a0. Texas
K b2 = 0
a2 =
M

Evaluate Eqs 7.71.


7.10  Eigenvalues and Eigenvectors 443

b0 0
β0 = β0 = =0
a0 1
b1 a1 1 b 1
β1 = − β0 → β1 = − 0 =
a0 a0 M M M
b2 a1 a2 0 b K b 1 b
β2 = − β1 − β0 β2 = − β1 − 0 =− =− 2
a0 a0 a0 1 M M M M M

x1 = y − β0 u x1 = v − β0 F x1 = v
dx → dx → dv 1
x2 = 1 − β1u x2 = 1 − β1 F x2 = − F
dt dt dt M

Substituting into the state and output equations yields In order to perform the needed algebra, the state equation
dx
= Ax + Bu must be converted into algebraic equations by
 1  dt
 0 1   M  use of Laplace transformation.
d  x1    x 
b    + 
1
= K F
dt  x2   − −  x2   − b 
 M M dx  dx 
 M 2  = Ax + Bu → L   = L { Ax + Bu}
dt  dt 
 x1 
y = [1 0]   + 0 F
 x2  L  ddtx  = L {Ax} + L {Bu}
 

7.10  Eigenvalues and Eigenvectors L  ddtx  = AL {x} + BL {u}


 
7.10.1 Eigenvalues
 sX ( s ) = AX ( s ) + BU ( s ) (7.73)
As the algorithm for expressing a system equation in the
state-space illustrated, state variables and their correspond- Perform Laplace transformation on Eq. 7.71. The transfor-
ing state equations are not unique. We can define a new set mation matrix is a constant matrix.
of state equations by performing vector-matrix algebra on an
existing set. When different sets of state equations describe
the same physical system, there are attributes of the different
L {x} = L {Pz} → L {x} = PL {z}
state equations which are the same. Specifically, all of the
different sets of state equations must yield the same char- X( s ) = PZ( s )
(7.74)
acteristic equation, since the characteristic equation, indeed,
characterizes the physical attributes of the system’s homoge- We can now eliminate the existing state vector, X( s), from
neous, or natural, response to a disturbance. The roots of the the state equations 7.72 by substitution.
characteristic equation, the characteristic values, are more
commonly known by their combination German–English 
sPZ ( s ) = APZ ( s ) + BU ( s ) (7.75)
name, “eigenvalues.” “Eigen” translates from German as
“one’s own” or “inherent,” or, in other words, characteristic.
Conventionally, characteristic values are represented by the The left side of Eq. 7.74 is not in the proper form for state
Roman character, s, and eigenvalues by the Greek character, equations. We must remove matrix P from the product,
λ, but, again, they are of the same quantity. sPZ( s). To do so, we premultiply both sides of Eq. 7.74 by
The algebra of creating a new set of state variables and the inverse of matrix P.
state equations is known as a “transformation.” In the ab-
stract, a transformation is straightforward. Define a square
“transformation” matrix P and a new vector of state variables P −1 sPZ ( s ) = P −1 ( APZ ( s ) + BU ( s ))
z, by equating their product with the existing state vector x.
 P −1 sPZ ( s ) = P −1 APZ ( s ) + P −1BU ( s ) (7.76)
 x = Pz (7.72)
444 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Although the Laplace variable s is complex, it is a scalar by expressing the equation in a form which can be expanded.
variable and can be factored out of the product on the left Using a two by two coefficient matrix; for example,
side of Eq.  7.75.
 a11 a12   v1   v1   a11v1 + a12 v2   λ v1 
 a    = λ  → a v + a v  = λv 
sP PZ ( s ) = P APZ ( s ) + P BU ( s )
−1 −1 −1
(7.77)
 21 a22  v2  v2   21 1 22 2   2

The product of the matrix P and its inverse matrix P−1 is the Just as eigenvalues have physical meaning when they are as-
identity matrix I sociated with a physical system, so too their corresponding
eigenvectors. In the case of a lumped parameter mechani-
 cal system, i.e., a spring-mass-damper system, with multiple
sIZ ( s ) = P −1 APZ ( s ) + P −1BU ( s ) (7.78) masses and an oscillatory impulse or step response, the ei-
genvalues represent the different decay rates and frequen-
The identity matrix I can be inserted, if the resulting product cies. Their corresponding eigenvectors represent the shape
can be evaluated, or removed at will. We remove it to yield of each “mode” superimposed to create the vibration.
the transformed state equations in the standard form. In the bad old days, engineering students studied arcane
methods for calculating eigenvectors. In the present day,
 Mathcad, MATLAB, and hand-held engineering calculators
sZ ( s ) = P −1 APZ ( s ) + P −1BU ( s ) (7.79) have built-in routines for finding eigenvectors. Using the
coefficient matrix of the RLC circuit’s state-space represen-
where the products, P −1 AP and P −1B, are the transformations tation, with energy storage variables as state variables, and
of the original coefficient matrix A and input matrix B. letting R = 3, L = 2, and C = 1, for illustration,
State-space based automatic control is known as “Modern
Control,” because it was modern in the 1960s. The transfor-
 R 1  3 1
mation illustrated in Eqs. 7.73–7.79 is a key step in Modern − L −  −
L 2
− 
2
Control design. The objective of the transformation is to use A=  = 
a transformation matrix P, which yields a diagonal matrix  1 0  
1
0 
P −1 AP. The significance is that the elements of the diagonal  C   1 
are the eigenvalues of the system
The eigenvalues are
  z1   λ 1 0 0 0   z1 
z   0 λ 0 0   z2  R  R
2
4 3  3 4
2

s  =    + P −1BU ( s ) (7.80) − ±   − − ±   −
2 2

           L  L LC 2  2 2 ⋅1
     λ1 , λ 2 = =
0 0 λ n   zn  2 2
 zn   0
λ 1 = −0.5 and λ 2 = −1.0
allowing the control designer to change the eigenvalues indi-
vidually, and tailor the response of the dynamic system under The corresponding eigenvectors, calculated with Mathcad,
feedback control. are

 0.447   −0.707 
v1 =   and v2 =  
7.10.2 Eigenvectors  −0.894  0.707 

An eigenvalue and its associated eigenvector are defined by


the scalar-matrix equation. Summary

 Av = λ v (7.81) Linear algebra is a condensed form of the algebra of simulta-


neous equations where the variables are represented as vec-
Note that the left side of the equation is the product of a ma- tors and the constants or coefficients are represented by ma-
trix and a vector, and the right side is the product of a scalar trices. The key differences between scalar and vector-matrix
and a vector. The equality of the two sides is easiest to see, algebra are:
Problems 445

Vector-matrix multiplication is not commutative, except transformation. State variables created by a transformation
in the two special cases below. usually have no physical meaning.
The reduction of the energetic equations of a system to
AB ≠ BA a set of state equations is significantly easier than the cor-
responding reduction to a higher-order differential equation
Scalar operators are applied to each element with a matrix. because more variables are retained, leaving a few that must
Scalar operations on a matrix are commutative. be eliminated by substitution.
The state equations are solved by finite difference meth-
aB = Ba ods, specifically the Runge–Kutta algorithm. The solution
yields the state variable values over the duration of the solu-
There is no vector-matrix division; it is not defined. The tion.
equivalent is to multiply a matrix by its inverse, which is The solution for a power variable other than a state vari-
the quantity that when multiplied by the matrix yields the able requires a second set of purely algebraic equations
identity matrix I. called the output equations. The output equations cannot
be solved until the state equations have been solved and the
AA −1 = A −1 A = I state vector x is known. The output equations in vector-ma-
trix form are:
where I is the identity matrix:
y = Cx + Du
AI = IA
where y is the vector of output variables, C is the output
and matrix, and D is the direct pass-through matrix.
1 0 … 0
0 1  0
I= Problems
   
  Problems 1, 2, and 3 are to be solved “manually,” mean-
0 0 0 1
ing, that you may use your calculator to perform arithmetic
State-space represents a dynamic system as a set of first- operations but not linear algebra. After you have solved the
order simultaneous differential equations: problems manually, you are encouraged to check your man-

 x1   a 1,1 a 1,2  a 1, n   x1   b1,1 b1,2  b1, m   u 1 (t ) 


    
d  x2   a 2,1 a 2,2  a 2, n   x2  b 2,1 b 2,2  b 2, m   u 2 (t ) 
=  +
dt                   
       
 xn   a n ,1 a n ,2  a n , n   xn  b n ,1 b n ,2  b n , m  u m (t )

Expressed in vector-matrix form as: ual linear algebra computations using your calculator so that
you become familiar with its capabilities.
dx
= Ax + Bu
dt Problem 7.1 Evaluate the following matrix expressions
manually
where x is the vector of state variables, known as the state
vector, A is the coefficient matrix, B is the input matrix, and 1 2  5 6 1 2  5 6
u is a vector of inputs.
7.1.a 3 4  +  7 8  = 7.1.b 8  − =
    3 4   7 8 
The state variables used to reduce the energetic equations
of a system, extracted from a schematic or a linear graph, are Problem 7.2 Evaluate the following matrix expressions
the energy storage variables. However, the state variables of manually
a system are not unique since linear algebra performed on
the state equations can be used to redefine the state variables 1 2  5 6  5 6 1 2
7.2.a 3 4   7 8  = 7.2.b   =
as a combination of a previous set, in what is known as a     7 8  3 4 
446 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

x,v Flywheel
Rotational Inertia J

F(t) K Compliant Shaft


M Torsional Spring K

Fluid Coupling
Damping b
Lubricating fluid
Damping b Angular Velocity
Input Ω(t)
Fig. P7.4   Second-order translational mechanical system

Fig. P7.5   Second-order rotational mechanical system

1 2 5  5
7.2.c    = 7.2.d [3 4]   =
3 4   6   6 undamped natural frequency ωn, the damped natural
frequency ωd, and the damping ratio ζ.
5 7.4.f The system is de-energized when it is acted on by
7.2.e   [3 4] =
6 a step input of 10 N. Solve the state equations and
the output equations using Mathcad or MATLAB for
Cases I and II. Plot the responses of the output vari-
Problem 7.3 Calculate the inverses of the following matrices ables.
manually, as the transpose of the cofactors of the matrix di-
vided by the determinant of matrix. No credit will be awarded Problem 7.5 A rotational mechanical system consisting of
for inverting the matrices using a graphical algorithm. Check a spring, an inertia, and a damper acted upon by an applied
by multiplying the matrix by its inverse manually. torque is shown in the schematic Fig. P7.5.

−1
−1  2 0 3 Table P7.5   Parameter values
 2 3
7.3.a  4 5 = 7.3.b  4 5 6 = J K b
  7 8 9  Case I 3 2 1
Case II 1 2 3

Problem 7.4 A translational mechanical system consisting


of a mass M sliding on a lubricating fluid film with damping 7.5.a Derive the second-order system equation for the
b. A spring K is attached between the mass and ground. The torque in the spring and express it in the standard
mass is acted upon by an applied force F( t), Fig. P7.4. form.
7.5.b Derive the state equations for this system.
Table P7.4   Parameter values 7.5.c Derive the output equations for:
M K b
i The angular velocity of the inertia J.
ii The torque acting through the compliant shaft
Case I 3 2 1
spring K.
Case II 1 2 3
iii The angular velocity difference across the compli-
ant shaft spring K.
7.4.a Derive the second-order system equation for the force iv The angular velocity difference across the fluid
acting through the spring and express it in standard coupling b.
form. 7.5.d Express the state and output equations in matrix form.
7.4.b Derive the state equations for this system. 7.5.e For Cases I and II, calculate the system’s eigenvalues.
7.4.c Derive the output equations for: If the system is underdamped, determine the ideal,
i The velocity of the mass M. undamped natural frequency ωn, the damped natural
ii The force acting through spring K. frequency ωd, and the damping ratio ζ.
iii The angular velocity difference across spring K. 7.5.f The system is de-energized when it is acted on by a
7.4.d Express the state and output equations in matrix form. step input of 10 rad/sec. Solve the state equations and
7.4.e For Cases I and II, calculate the system’s eigenvalues. the output equations using Mathcad or MATLAB for
If the system is underdamped, determine the ideal, Cases I and II. Plot the responses of the output vari-
ables.
Problems 447

x,v Problem 7.7 A fluid system consisting of a fluid power unit


K modeled as a pressure source, three fluid resistances, a fluid
b2 accumulator with capacitance C, a fluid inertance I, and a hy-
F(t) draulic piston/cylinder driving a mass M is shown is shown
M1 b1 M2
in the schematic, Fig. P7.7. Note that the mass and the fluid
inertance are dependent energy storage elements.

Fig. P7.6   Third-order mechanical system Table P7.7   Parameter values


R1 R2 R3 C I D M
0.1 0.2 0.3 2 3 6 4
Problem 7.6 A translational mechanical system consisting
of a spring, two masses, and two dampers acted upon by an 7.7.a Derive the state equations for this system and check
applied force is shown in the schematic Fig. P7.6. their units
7.7.b Derive output equations for:
Table P7.6   Parameter values i The pressure in the fluid accumulator.
M1 M2 K b1 b2 ii The velocity of mass M.
Case I 1.0 2.0 1.0 0.1 2.0 iii The volume flow rate through the fluid inertance.
Case II 0.2 0.4 1.0 0.1 0.2 iv The volume flow rate from the pressure source.
v The force acting to accelerate mass M.
vi The pressure drop across the fluid inertance.
7.6.a Derive the state equations for this system. 7.7.c Express the state and output equations in matrix form.
7.6.b Derive the output equations 7.7.d Calculate the system’s eigenvalues. If the system is
i The velocity of mass M1. underdamped, determine the ideal, undamped natural
ii The force acting through the spring K. frequency ωn, the damped natural frequency ωd, and
iii The velocity of mass M2. the damping ratio ζ.
iv The force acting through damper b1. 7.7.e The system was de-energized when a step input of 100
v The force acting through damper b2. was applied. Solve the state equations and the out-
7.6.c Express the state and output equations in matrix form. put equations using Mathcad or MATLAB. Plot the
7.6.d Calculate the system’s eigenvalues for Cases I and responses of the output variables.
II. If the system is underdamped, determine the ideal,
undamped natural frequency ωn, the damped natural Problem 7.8 A hybrid fluid-translational rotational system
frequency ωd, and the damping ratio ζ. is shown in Fig. P7.8. The fluid power unit is modeled as a
7.6.e The system was de-energized when a step input of pressure source, which discharges into a hydraulic line with
10 was applied. For Cases I and II, solve the state fluid resistance R. The hydraulic piston/cylinder has diam-
equations and the output equations using Mathcad or eter D. Its piston rod is attached to the linkage. A dashpot
MATLAB. Plot the responses of the output variables. with damping b and a spring K are also attached to the link-
age. The linkage has rotational inertia J.

Fig. P7.7   Hybrid fluid-transla- Fluid Power Unit


tional mechanical system Pressure Source p(t)

Breather Cap
(vent to atmosphere)

Hydraulic Line
Resistance R1 Fluid Accumulator
Capacitance C
Hydraulic Line
Resistance R 2
Hydraulic Piston-Cylinder
Diameter D

Press Die, Mass M


Hydraulic Line
Resistance R3
Inertance I
448 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. P7.8   Fluid-translational


Fluid Power Unit
mechanical system modeled as a
Pressure Source Dashpot
Damping b
Hydraulic Line
Fluid Resistance R
L1

L2
Tank vented to
the atmosphere
Pivot
Hydraulic Cylinder
Diameter D
Linkage Spring K
Rotational Inertia J

Fig. P7.9   Translational mechani- Mass M


cal system Lubricating Film
Damping b 2
Spring K2

Pivot Spring K1
Dashpot b1 L3

L2

Force Source F(t)


L1

Table P7.8   Parameter values


Problem 7.9 A translational mechanical system is shown
R D L1 L2 b J K
schematically in Fig. P7.9. The system consists of a force
0.6 0.12 1 2 50 20 1,500 source F( t) acting on a lever, to which are attached a damper
b1 and two springs, K1 and K2. Spring K2 is attached to mass
7.8.a Derive the state equations for this system and check M, which slides on a rail with a lubricating film, damping
their units. b2. The system is de-energized before the input step force of
7.8.b Derive output equations for: 1,000 N acts on the system.
i The volume flow rate from the source.
Table P7.9   Parameter values
ii The force applied by the piston rod.
iii The force in spring K. L1 L2 L3 b1 b2 K1 K2 M
iv The angular velocity of the linkage. 1 2 1 100 0.5 1,000 1,500 10
v The velocity difference across the spring.
7.8.c Express the state and output equations in matrix form. 7.9.a Derive the state equations for this system and check
7.8.d Calculate the system’s eigenvalues. If the system is their units.
underdamped, determine the ideal, undamped natural 7.9.b Derive output equations for:
frequency ωn, the damped natural frequency ωd, and i The force acting to accelerate mass M.
the damping ratio ζ. ii The velocity of mass M.
7.8.e The system was de-energized when a step input of iii The force acting through spring K1.
2,000 was applied. Solve the state equations and the iv The force acting through spring K2.
output equations using Mathcad or MATLAB. Plot the v The velocity of the point of application of the
responses of the output variables. input force F( t).
Problems 449

Fig. P7.10   Hybrid electrome- Electric Motor


Rotational Inertia J 1
chanical system Transduction α
Bearings Damping b1

Voltage Source v(t)


Compliant Shaft
Spring Constant K

Electrical Resistance R

Rotational Inertia J2
Bearings Damping b2

7.9.c Express the state and output equations in matrix form. 7.10.e The system was de-energized before a step input of
7.9.d Calculate the system’s eigenvalues. If the system is 48 VDC was applied. Solve the state equations and
underdamped, determine the ideal, undamped natural the output equations using Mathcad or MATLAB.
frequency ωn, the damped natural frequency ωd, and Plot the responses of the output variables.
the damping ratio ζ.
7.9.e Solve the state equations and the output equations Problem 7.11 A rotational system show is shown in Fig. 7.11.
using Mathcad or MATLAB. Plot the responses of the An angular velocity source drives the shaft attached to the
output variables. pulley with diameter D1. The belt over pulley D1 drives the
pulley with diameter D2, which is attached to the same shaft
Problem 7.10 An electromechanical system is shown as the pulley with diameter D3. That shaft is compliant, with
in Fig. P7.10. The input to the system is a voltage which spring constant K, and drives the flywheel with rotational
drives a DC motor through wires with electrical resistance inertia J1 and damping b1. The belt over pulley D3 drives the
R. The relationship between the motor’s current and torque pulley with diameter D4, which is attached to a shaft connect
is TM =  αiM. Flywheel J1 is attached to the motor’s shaft and to the fluid coupling with damping b2. The output shaft of the
is supported by bearings with damping b1. Coupled to the fluid coupling drives the flywheel with rotational inertia J2
motor’s shaft is a compliant shaft modeled as torsion spring and damping b1.
K. The compliant shaft drives flywheel J2,which is supported
by bearings with damping b2. Table P7.11a   Parameter values
D1 D2 D3 D4 b1 J1 J2 b2
Table P7.10   Parameter values
2 4 8 6 1,500 12 7 2
R α J1 b1 K J2 b2
2 8 3 0.6 1,000 4 0.5
Table 7.11b   Torsion spring stiffness K values
7.10.a Derive the state equations for this system and check Case I Case II Case III
their units. 500 1,500 3,000
7.10.b Derive output equations for:
i The current from the source.
ii The torque acting to accelerate angular velocity 7.11.a Derive the state equations for this system and check
of rotational inertia J1. their units.
iii The angular velocity of rotational inertia J1. 7.11.b Derive output equations for:
iv The angular velocity of rotational inertia J2. i The torque acting through spring K.
v The torque acting through spring K. ii The angular velocity of flywheel J1.
7.10.c Express the state and output equations in matrix iii The torque applied by the angular velocity source.
form. iv The torque acting to accelerate flywheel J2.
7.10.d Calculate the system’s eigenvalues. If the system is v The angular velocity of flywheel J2.
underdamped, determine the ideal, undamped natural
frequency ωn, the damped natural frequency ωd, and
the damping ratio ζ.
450 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. P.7.11   Rotational system Flywheel


Rotational Inertia J1
Pulley Compliant Shaft Damping b1
Spring Constant K
Diameter D1 Flywheel
Pulleys Diameters Rotational Inertia J 2
D2 and D3 Damping b1

Angular Velocity
Input Ω(t)

Pulley Fluid Coupling


Diameter D4 Damping b2

Fig. P7.12a   A shaft-driven rota- Rotational Inertia J


tional system with a compound
gear Compliant Shaft Bearing
Torsion Spring K2 Damping b3

N4 N2
Compliant Shaft
Torsion Spring K 1 N3 Bearing
Fluid Coupling N1 Damping b2
Damping b1

Angular Velocity
Source, Ω(t)

7.11.c Express the state and output equations in matrix torsional spring constant K. The compliant shaft drives a pin-
form. ion with N1 teeth. The pinion engages a gear with N2 teeth.
7.11.d Calculate the system’s eigenvalues for Cases I, The gear shaft drives rotational inertia J2. The gear shaft
II, and III using the damping coefficient values of bearings have damping b2.
Table  7.11b. If the system is underdamped, deter-
mine the ideal, undamped natural frequency ωn, Table 7.12a   Parameter values
the damped natural frequency ωd, and the damping b2 b3 K1 K2 J
ratio ζ. N ⋅ m ⋅ sec N ⋅ m ⋅ sec N N
20 10 2,000 4,000 10 kg ⋅ m 2
7.11.e The system was de-energized before a step input of rad rad rad rad
1,500 RPM was applied. Solve the state equations
using Mathcad or MATLAB. Table 7.12b   Fluid coupling damping b1 values
7.11.f Solve and plot the responses of the output variables Case I Case II Case III
using Mathcad or MATLAB. N ⋅ m ⋅ sec N ⋅ m ⋅ sec N ⋅ m ⋅ sec
5 1 0.1
rad rad rad
Problem 7.12 A rotational mechanical system is shown
schematically in Fig. P7.12. The system consists of an an-
gular velocity source Ω( t) acting on the input shaft of a fluid Table 7.12c   Gear teeth numbers
coupling with damping b1. The output shaft of the fluid cou- N1 N2 N3 N4
pling drives inertia J1, which drives a compliant shaft with 12 48 18 54
Problems 451

3,000
200

100
2,000
p(t) T(t)
0
psi N •m 0 1 2 t, sec
1,000 -100

-200
0
0 1 2 3 Fig. P7.13b   Torque input
t, sec

Fig. P7.12b   Angular velocity input tional inertia J. The gear drives a compliant shaft modeled
as a torsional spring with spring constant K. The shaft turns
a power screw with a linear pitch of n threads per inch. The
7.12.a Derive the state equations and check their units. power screw threads through the crosshead. The crosshead
7.12.b Derive output equations for the following variables: translates when the power screw rotates. The viscous friction
i The angular velocity of the pinion. between the rotating power screw and the translating cross-
ii The torque acting through the compliant shaft. head is modeled as rotational damping b1. The crosshead has
iii The torque acting to accelerate inertia J2. mass M and slides on two rails. The combined viscous fric-
iv The torque acting through the fluid coupling b1. tion of the crosshead sliding on the two rails is modeled as
7.12.c Write the state equations and output equations in translational damping b2.
vector-matrix form. 7.13.a Derive the state equations and check their units.
7.12.d Calculate the system’s eigenvalues for Cases I, II, 7.13.b Derive output equations for the following variables:
and III. If the system is underdamped, determine the i The angular velocity of the pinion shaft.
ideal, undamped natural frequency ωn, the damped ii The angular velocity gear N2.
natural frequency ωd, and the damping ratio ζ. iii The torque in the compliant shaft.
The system is de-energized before the input pulse of iv The force acting to accelerate the crosshead.
angular velocity shown in Fig. 7.12b is applied. v The translational velocity of the crosshead.
7.12.e Solve and plot state equations for the velocity pulse 7.13.c Write the state equations and output equations in
shown in Fig. P7.12b. vector-matrix form.
7.12.f Solve and plot the output equations. 7.13.d Calculate the system’s eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
Problem 7.13 A rotational-translational mechanical system frequency ωn, the damped natural frequency ωd, and
is shown schematically in Fig. P7.13. The system consists the damping ratio ζ.
of a torque source acting on the input shaft of a pinion with
N1 teeth. The pinion engages a gear with N2 teeth and rota-

Fig. P7.13   A hybrid rotational- Crosshead


translational system Mass M
Compliant Shaft Rail
Gear Spring Constant K
N 2 teeth
Power Screw
Inertia J 4 threads per inch

T(t)

Pinion
Viscous Friction of
N1 teeth threads in crosshead Viscous Friction of crosshead
Rotational Damping b1 sliding on rail Damping b2
(total for both rails)
452 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. P7.14 a  Hybrid fluid-rota- Flywheel, Inertia J and Damping b


tional mechanical system
Compliant Shaft, Torsion Spring K
Hydraulic Motor
Displacement Vol per revolution

High Pressure Line with


Fluid Inertance I and
Fluid Resistance R

Lower Pressure
Return Line

Breather Cap

Fluid Power Unit


Pressure Source p(t)

Table 7.13   Parameter values 3,000


N1 N2 J b1 K M b2 p(t)
psi
N ⋅ m ⋅ sec N N ⋅ m ⋅ sec 2,000
12 48 1 kg ⋅ m 2 1 2,000 5 kg 0.2
rad rad rad 1,500
1,000
The system is de-energized before the input torque
pulse shown in Fig. P7.13b was applied to the sys-
tem. 0 1 2 3
t, sec
7.13.e Solve the state equations using Mathcad or MAT-
LAB for the input torque pulse shown in Fig. P7.13b Fig. P7.14b   Pressure input
with the parameter values of Table 7.
7.13.f Solve and plot the responses of the output variables
using Mathcad or MATLAB. 7.14.a Derive the state equations and check their units.
7.14.b Derive output equations for the following variables:
Problem 7.14 A fluid-rotational mechanical system is shown i The volume flow rate from the fluid power unit.
in Fig. P7.14. The fluid power unit is modeled as pressure ii The pressure in the fluid accumulator.
source p( t), drawing fluid from a reservoir vented to the iii The angular velocity of the hydraulic motor.
atmosphere. The pump’s high-pressure line has fluid resis- iv The torque acting to accelerate the flywheel.
tance R and inertance I. The rotational hydraulic motor pro- v The angular velocity of the flywheel.
duces 600 ft.-lbs. of torque at a hydraulic fluid pressure of 7.14.c Write the state equations and output equations in
3,000 psi. The fluid discharging from the motor returns to vector-matrix form.
the reservoir. The output shaft of the motor turns a compli- 7.14.d Calculate the system’s eigenvalues. If the system is
ant shaft with torsion spring constant K. The compliant shaft underdamped, determine the ideal, undamped natu-
drives a flywheel with mass moment of inertia J supported ral frequency ωn, the damped natural frequency ωd,
on bearings with damping b. and the damping ratio ζ. If the real components of
the eigenvalues vary by two or more orders of mag-
Table 7.14   Parameter values nitude, then the duration of the components of the
R I b K J response vary by like magnitude. If so, solve the
MPa ⋅ sec kg N ⋅ m ⋅ sec N state equations twice, once using a short duration to
80 24,000 4 10 4,000 5 kg ⋅ m 2
m3 m rad rad see the fast response.
The system is running in steady state at the input
pressure of 1,500 psi before the input torque pulse
shown in Fig. P7.13b was applied to the system at
time t = 0.
Problems 453

Fig. P7.15   A rotational mechani- Shaft rigidly attached


cal system Drag Cup, Damping b 2

Flywheel, Rotational Inertia J2

Compliant Shaft, Torsional Spring K2

Flywheel, Rotational Inertia J1

Compliant Shaft, Torsional Spring K1

Fluid Coupling, Damping b1

Angular Velocity
Input Ω(t)

7.14.e Solve the state equations using Mathcad or MATLAB


L1 L2
for the input pressure pulse shown in Fig. P7.13b
with the parameter values of Table 7.14.
7.14.f Plot the responses of the output variables using
Mathcad or MATLAB. v(t) C1 R C2
Problem 7.15 A rotational system is shown in Fig. 7.15. An
angular velocity source drives the input shaft to a fluid cou-
pling with damping b1. The output shaft of the fluid coupling Fig. P7.16   An RLC circuit
drives shaft is compliant, with spring constant K1, which
drives a flywheel J1 with rotational inertia J1. A compliant
shaft with spring constant K2, from flywheel J1 drives fly- 7.15.e The system was de-energized before a step input of
wheel J2 with rotational inertia J2. A drag cup with damping 1,500 RPM was applied. Solve the state equations
b2 is connected between the hub of flywheel J2 and the ma- and plot their responses using Mathcad or MATLAB
chine frame, which is ground. for cases I, II, and III.
7.15.f Solve and plot the responses of the output variables
Table 7.15a   Parameter values using Mathcad or MATLAB for cases I, II, and III.
K1 J1 K2 J2 b2 7.15.g Calculate and plot the power dissipated in the fluid
N⋅m N⋅m N ⋅ m ⋅ sec coupling b1 and the drag cup b2 for cases I, II, and III.
600 2 kg ⋅ m 2 400 1 kg ⋅ m 2 0.2
rad rad rad
Problem 7.16 The schematic of an electric circuit with a
voltage source, two inductors, two capacitors and a resistor
Table 7.15b   Fluid coupling damping b1 values
is shown in Fig. P7.16.
Case I Case II Case III
N ⋅ m ⋅ sec N ⋅ m ⋅ sec N ⋅ m ⋅ sec Table 7.16a   Parameter values
2 20 40
rad rad rad C1 C2 L1 L2
1 µF 2 µF 1 mH 2 mH
7.15.a Derive the state equations and check their units.
7.15.b Derive output equations for the following variables:
i The angular velocity of the pinion. Table 7.16b  Resistor R values
ii The torque acting through the compliant shaft. Case I Case II Case III
iii The torque acting to accelerate inertia J2. 2Ω 20 Ω 40 Ω
iv The torque acting through the fluid coupling b1.
7.15.c Write the state equations and output equations in 7.16.a Derive the state equations and check their units.
vector-matrix form. 7.16.b Derive output equations for the following variables:
7.15.d Calculate the system’s eigenvalues for Cases I, II, i The current flowing through capacitor C1.
and III. If the system is underdamped, determine the ii The current flowing through capacitor C2.
ideal, undamped natural frequency ωn, the damped
natural frequency ωd, and the damping ratio ζ.
454 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

L1 L2 Table 7.17b  Resistor R values


Case I Case II Case III
2Ω 10 Ω 20 Ω

v(t) C1 R C2
7.17.a Derive the state equations and check their units.
7.17.b Derive output equations for the following variables:
i The current flowing through capacitor C1.
Fig. P7.17   An RLC circuit
ii The current flowing through capacitor C2.
iii The voltage across inductor L1.
iv The current through resistor R.
iii The voltage across inductor L1. 7.17.c Write the state equations and output equations in
iv The current through resistor R. vector-matrix form.
7.16.c Write the state equations and output equations in 7.17.d Calculate the system’s eigenvalues for Cases I, II,
vector-matrix form. and III. If the system is underdamped, determine the
7.16.d Calculate the system’s eigenvalues for Cases I, II, ideal, undamped natural frequency ωn, the damped
and III. If the system is underdamped, determine the natural frequency ωd, and the damping ratio ζ.
ideal, undamped natural frequency ωn, the damped 7.17.e The system was de-energized before a step input of
natural frequency ωd, and the damping ratio ζ. 24 VDC was applied. Solve the state equations and
7.16.e The system was de-energized before a step input of plot their responses using Mathcad or MATLAB for
24 VDC was applied. Solve the state equations and cases I, II, and III.
plot their responses using Mathcad or MATLAB for 7.17.f Solve and plot the responses of the output variables
cases I, II, and III. using Mathcad or MATLAB for cases I, II, and III.
7.16.f Solve and plot the responses of the output variables 7.17.g Calculate and plot the power dissipated in resistor R
using Mathcad or MATLAB for cases I, II, and III. for cases I, II, and III.
7.16.g Calculate and plot the power dissipated in resistor R
for cases I, II, and III. Problem 7.18 An electromechanical system is shown in
Fig. P7.18a. A voltage source drives a DC motor, which has
Problem 7.17 The schematic of an electric circuit with a resistance R and torque constant KT. The DC motor turns a
voltage source, two inductors, two capacitors and a resistor compliant shaft, modeled as a torsion spring with spring con-
is shown in Fig. P7.17. stant K1.
The shaft turns flywheel1 with mass moment of inertia J1.
Table 7.17a   Parameter values Flywheel1 is supported on bearings with damping b1. Fly-
C1 C2 L1 L2
wheel2 has mass moment of inertia J2 bearings with damping
b2, and is belt driven. The belt which runs between flywheels
1 μF 2 μF 1 mH 2 mH
one and two is long enough that the energy stored in the taut

Fig. P7.18a   Electromechanical Voltage Source v(t) Electrical Resistance R


system with a compliant shaft
and a compliant belt Compliant Shaft
Torsional Spring K1

Flywheel 1
Rotational Inertia J 1 DC Electric Motor
Bearings Damping b1 Torque constant K T

Flywheel 2
Rotational Inertia J2
6 in. Diameter Pulley Bearings Damping b 2

Compliant Belt
Translational Spring Constant K2

4 in. Diameter Pulley


Problems 455

Fig. P7.18b   The compliant vsurface vsurface


belt’s taut and slack sides switch
positions during an oscillation Slack Taut
Ω 2g Ω1g Ω 2g
Ω1g Taut Slack
Driving vsurface Driven Driven
Pulley Driving vsurface
Pulley Pulley Pulley

7.18.a Derive the state equations and check their units.


K2 7.18.b Derive output equations for the following variables:
n-1 n n+1 n+2
i The current drawn from the voltage source.
ii The back-EMF of the DC motor.
iii The torque acting to accelerate flywheel J1.
Tα Fα Fβ Tβ iv The torque acting to accelerate flywheel J2.
7.18.c Write the state equations and output equations in
vector-matrix form.
7.18.d Calculate the system’s eigenvalues. If the system is
g
underdamped, determine the ideal, undamped natural
Fig. P7.18c   Linear graph representing a compliant belt as a transla- frequency ωn, the damped natural frequency ωd, and
tional spring between two rotational to translational transducer inter- the damping ratio ζ.
faces 7.18.e The system was de-energized before a step input of
48 VDC was applied. Solve the state equations and
or stretched side of the belt, Fig. P7.18b, must be included plot their responses using Mathcad or MATLAB.
in the dynamic system. The compliant belt is represented 7.18.f Solve and plot the responses of the output variables
by a translational spring between two transducers, which using Mathcad or MATLAB.
interface between the torque on the pulley’s shafts and the
force carried by the belt, Fig. P7.18c. The sign reversal in the Problem 7.19 A hybrid rotational-translational mechanical
translational spring represents the taut and slack sides of the system is shown schematically in Fig. P7.16. The system
belt’s switching positions. consists of a force source F( t) acting on a lever, to which are
attached a damper b1 and two springs, K1 and K2. The lever
Table 7.18   Parameter values has mass moment of inertia J. Spring K2 is attached to mass
R KT K1 J1 b1 M, which slides on a rail with a lubricating film, damping
N⋅m N⋅m N ⋅ m ⋅ sec
b2. The system is de-energized before the input step force of
3Ω 9 800 6 kg ⋅ m 2 1 1,000 N acts on the system.
A rad rad
K2 J2 b2
Table P7.19   Parameter values
N N ⋅ m ⋅ sec
8,000 5 kg ⋅ m 2 1 L1 L2 L3 b1 b2 J K1 K2 M
m rad
1 2 1 20 1 50 1,000 1,500 10

Fig. P7.19   Hybrid rotational- Mass M


translational mechanical system

Lubricating Film Spring K2


Damping b 2

Rotational Inertia J

Pivot Spring K1
Dashpot b1 L3

L2

Force Source F(t)


L1
456 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. 5.52   Thermal system Perfect Insulation


divides into five regions, each No heat flow
with a thermal capacitance
between the center temperature
node of the region and ground RSS RSS RAl RAl RAl RAl RAl RAl RAl RAl
temperature θ1 θ11
q in 2 3 4 5 6 7 8 9 10 qout
Heat Flow Stainless
Rate Steel Aluminum

Surface Area 0.5 in 2 in


A = 48 ft 2

Fig. 5.53  Linear graph of the R ss1 Rss 2 R al 1 R al2 R al 3 Ral 4


stainless steel and aluminum 1 2 3 4 5 6 7
thermal system. The linear graph
is split between nodes 6 and 7 for
this figure Css C al1 Cal 2
θ1(t)

Ral 4 R al 5 Ral 6 R al 7 Ral 8


6 7 8 9 10 11

Cal 2 Cal 3 Cal 4 θ2(t)

7.19.a Derive the state equations for this system and check surfaces. The thermal system model consists of thermal re-
their units. sistances in series and thermal capacitances in parallel. The
7.19.b Derive output equations for linear graph is Fig. 5.53. The thermal resistances of the stain-
i The force acting to accelerate mass M. less steel and aluminum are calculated for 0.25 in. layer. The
ii The velocity of mass M. thermal capacitances are calculated pairing adjacent layers
iii The force acting through spring K1. and using the temperature node between them as the tem-
iv The force acting through spring K2. perature of the 0.5 in. thick capacitance.
v The angular velocity of the lever. The thermal resistances of the 0.25 in. thick stainless steel
vi The velocity of the point of application of the and aluminum layers are:
input force F( t). o
o
K K
7.19.c Express the state and output equations in matrix RSS = 8.23 × 10 −5 and RAl = 6.91 × 10 −6
W W
form.
7.19.d Calculate the system’s eigenvalues. If the system is The thermal capacitances of the stainless steel and aluminum
underdamped, determine the ideal, undamped natural are:
frequency ωn, the damped natural frequency ωd, and
the damping ratio ζ. kJ kJ
C pSS = 204 o
and C pAl = 137 o
7.19.e Solve the state equations and the output equations K K
using Mathcad or MATLAB.
7.19.f Plot the responses of the output variables using 7.20.a Derive the state equations for this system and check
Mathcad or MATLAB. their units.
7.20.b Derive output equations for:
Problem 7.20 The stainless steel and aluminum thermal sys- i The temperatures of nodes 2, 4, 6, 8, and 10.
tem of Sect. 5.8.4 is reproduced as Fig. 5.52. The stainless ii The heat flow rate through thermal resistances
steel is 0.5 in. thick. The aluminum is 2 in. thick. Both metals RSS1, Ral1, Ral3, Ral5, Ral7, Ral9, and Ral10.
are divided into layers 0.25 in. thick, parallel to the external
Chapter 7 Appendix 457

Fig. 3.13  a Translational me-


chanical mass-damper system
a b 1
driven by a force source, b Lin- x,v
ear graph of the mass-damper
system driven by a force source
Mass F(t) b M
F(t) Lubricating
g M fluid film
1 Damping b
g g

7.20.c Express the state and output equations in matrix extracted using a special operator. Typing Ctrl and 6 simul-
form. taneously immediately following a matrix variable produces
7.20.d Calculate the system’s eigenvalues. a pair of exponentiated angle brackets. The column number
The thermal system is at the uniform temperature of to be extracted is inserted into the angle brackets. Remember
20 °C when the temperature of node 1 is given a step that Mathcad’s “origin,” the beginning value of indices, is
increase to 100 °C. The temperature of node 11 is zero, unless it is changed by the user. Hence, the first column
held at 20 °C. of a matrix is column zero. The following command assigns
7.20.e Solve the state equations and the output equations the first column of the array Z to the variable t.
using Mathcad or MATLAB.
7.20.f Plot the responses of the output variables using t:= Z0
Mathcad or MATLAB.
The Runge–Kutta method will fail when simulating second-
order or higher systems if the time step is too large. The cal-
Chapter 7 Appendix culation will become “numerically unstable” and produce a
result with growing oscillations that head to either positive
Mathcad’s Runge–Kutta Solver or negative infinity. Errors formulating the state equations,
particularly sign errors, can also lead to numerical instability.
Mathcad has two versions of the Runge–Kutta algorithm, The physical systems which are not under feedback control
rkfixed(), which has a constant, or fixed, time step Δt and are physically stable. Therefore, if the calculation is unstable
Rkadapt(). Rkadapt() adapts or varies the time step, reducing reduce the time step by an order of magnitude or two. If the
its size if the finite differences become large and increasing calculation is still numerically unstable, then check your
it when they are small. We will use rkfixed(). Rkadapt() is state equations.
used when a calculation is numerically stable only with an The procedure will be illustrated with a first-order mass-
excessive number of time steps. damper system. The first-order system will also be used to
The Runge–Kutta solver rkfixed() is a function. It is used demonstrate the response of the system to inputs other than
in an assignment statement where the result of the Runge– step inputs. Next, the response of a second-order spring-
Kutta calculation is assigned to a variable, say Z. The mass-damper system will be investigated.
Mathcad syntax is:

Z : = rkfixed(a, b, c, d, e) Step Response of a Linear Mass-Damper System

where Figure  3.13 showing the schematic and linear graph of a


a is the state variable vector. force source sliding a mass on a lubricating film and the
b is the beginning time of the simulation. system equation for the velocity of the mass, Eq. 3.26, are
c is the end time of the simulation. reproduced here for reference.
d is the number of time intervals in the simulation.
e is the vector containing the state equations. 
The time step Δt = (c-b)/d. F (t ) M dv1g F (t ) dv1g b
= + v1g → = − v1g (3.26)
The function rkfixed() returns an array, Z in this example. b b dt M dt M
The first column of the array is time. The succeeding col-
umns are the values of the state variables at those times, i.e., Standard form of a first-order system equation requires little
the second column is the first state variable, the third column rearrangement to form into a state equation.
is the second state variable, etc. A column of a matrix can be
458 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

6
F (t ) dv1g b dv1g b F (t ) 6
= + v1g → = − v1g +
M dt M dt M M

4
The Mathcad code begins by defining the input function, pa-
rameter values, and the initial value of the state variable. We V.1g
will first solve the state equation for a step input with the sys-
2
tem initially de-energized using the following parameters:

M : = 3 b : = 2 F : = 10 0
0
0 2 4 6 8 10
Define a state variable vector by assigning the initial value
0 t1 10
to each state variable. Here, the mathematical meaning of
“vector” collides with the computer science meaning. Each Fig. A7.1   Step response of a linear mass-damper system
individual state variable must be a computer science vec-
tor to contain the value of the variable at each time step in
the calculation. Recall that the period indicates a “literal” 100 = number of increments
subscript that Mathcad accepts as part of the variable name D = vector of state equations
[indicates what follows to be a vector or matrix subscript]. We can plot the columns of the array Z directly, but this leads
to errors because it is easy to forget which state variable is
Type x[0 to see x0 when there is more than one. It is best to copy the columns
of the array to vectors.
Mathcad recognizes x as a vector variable because it has vec-
tor a subscript. Define the initial value of the state variable t1: = Z 0 v1g : = Z 1
x to be zero.
The subscripts 1 g are literal subscripts produced by typing a
x0 : = 0 period after “v,” v.1 g
This procedure hardly saves time when solving a linear
The state equation must be defined as a function to two vari- first-order equation subjected to a step input, since there are
ables. The first argument, t, is the independent variable. The only four possible solutions. Fortunately, we can use the
second argument is the name of the state vector. We have a same procedure to solve some non-linear equations and all
single state variable, not a state variable, so we identify the linear equations subjected to arbitrary inputs.
state variable. One can name the function which defines the First, let’s compare the above response for a mass-damper
state equations. We will use D(t, x). with linear damping Fb = bv1g with the response of a system
with non-linear damping FbNL = bv12g .
−b F
D ( t, x ) : = ⋅x +
M M
Step Response of a Mass-Damper System with a
In general, the function D(t, x) contains a vector of the state Non-Linear Damper
equations. In this case, there is only one equation. Notice
that the independent variable t does not need to appear on The Mathcad code is identical, except the state equation.
the right side of the assignment statement. It can be used, as We will again define the parameter values, the magnitude
demonstrated below. of the input force, and the initial value of the state variable
First-order system difference equations cannot be nu- so that the code is complete. There is no need to repeat these
merically unstable. We will use a large number of steps, say Mathcad statements if they are unchanged.
1,000, which is far more than is needed for accuracy.
The rkfixed function assigning its result to Z is: M: = 3 b: = 2 F : = 10

Z := rkfixed(x, 0,10,1000, D) Define the initial value of the state variable to be zero.

The arguments of rkfixed() are: x0 : = 0


x = vector of state variables.
0 = start time Define the state equation. Note that the state variable is
10 = end time squared.
Chapter 7 Appendix 459

6 6
6 6 5

4
4
4 2
v.1g

v.NL
v.1g 2 0 ( )
F t3q

2 −2
0
−4
0 − 1.357 −5
0
0 2 4 6 8 10 −2 −6
0 2 4 6 8 10
0 t2 10
0 t3 , t3q 10

Fig. A7.2   Step response of a linear mass-damper system, solid trace,


Fig. A7.3   Pulse response of a linear mass-damper system, solid trace,
and a non-linear mass-damper system, dotted trace
and the pulse input, dotted trace

−b 2 F
D ( t, x ) : = ⋅x +
M M Copy the columns of the result to vectors.

Assign the result of the Runge–Kutta function to the variable t3 : = Zpulses 0 v pulses : = Zpulses 1
ZNL, where the subscript NL is a literal subscript.
In order to plot the input pulse function, we define a range
Z NL : = rkfixed(x, 0,10,1000, D) variable, q, of the same length as the time vector t3. The vari-
able q must start at zero, so we must subtract one from the
It is best to copy the columns of the matrix Z NL to vectors. length of the vector t3.
We could use the same name for the time vector since they
are identical. We will create a new time vector. q : = 0..length(t3) − 1

t2 : = Z NL 0 v NL : = Z NL 1 The pulse input is plotted on the secondary y-axis using the


range variable as the vector subscript for the time vector,
Fig. A7.3.
Response of a Linear Mass-Damper System
Subjected to a Pulse Train
Response of a Mass-Damper System to Sinusoidal
D(t, x) is a function of t, allowing the input functions of the Inputs
state equations to functions of time. Example: Fpulses( t) is a
pulse train made of step functions. Calculation of the response of the first-order mass-damper
system to sinusoidal inputs is a second example of the use of
F(t) : = 5 ⋅ Φ (t) − 10 ⋅ Φ (t − 2) + 5 ⋅ Φ (t − 4) the independent variable t to define a function of time as the
input to the state equations.
The remaining Mathcad code is unchanged except that the We will again define values even though they have been
input force in the state equations is now a function of time. previously defined and are unchanged so the each block of
Mathcad code is complete.
M: = 3 b: = 2
M: = 3 b: = 2
x0 : = 0
x0 : = 0
−b 1
D ( t, x ) : = ⋅ x + F(t)
M M −b 1
D(t, x) : = ⋅ x + sin(π ⋅ t)
M M
Assign the result of the Runge–Kutta function to the variable Zpi : = rkfixed(x, 0,10,1000, D)
Zpulses.
t4 : = Zpi 0 v pi : = Zpi 1
Zpulses : = rkfixed(x, 0,10,1000, D)
460 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

0.1
0.091
in text as a comment, such as x1 = v1g and x2 = FK, and then
the initial values of the state variables defined as
0.05
x10 : = 0 x 20 : = 0
v.2pi

v.4pi
0
where the subscripts are vector subscripts created by typing.
The state vector is then defined as:
− 0.05
 x10 
x:=
− 0.053  x 20 
− 0.1
0 2 4 6 8 10
0 t1 , t2 10 This syntax is very specific. The state vector x cannot have a
vector subscript. The state variables x1 and x2 must have the
Fig. A7.4   Sinusoidal responses of the first-order mass-damper sys-
vector subscript zero.
tem. Solid trace input frequency ω = 2π. Dashed trace, input frequency
ω = 4π The alternative is to give the individual state variables
meaningful names:
Second block of code:
v1g 0 : = 0 FK 0 : = 0 or v1g : = 0 FK 0 : = 0
M: = 3 b: = 2 0

x0 : = 0 where the latter have literal subscripts for 1 g and K and vec-
−b 1 tor subscripts for zero.
D(t, x) : = ⋅ x + sin(4 ⋅ π ⋅ t) The state vector x is the vector of the initial values of the
M M
Z4pi : = rkfixed(x , 0,10,1000, D) state variables where, again, the state vector x cannot have a
vector subscript and the state variables v1g and FK must have
t4 : = Z4pi 0 v 4pi : = Z 4pi 1 the vector subscript of zero.

Step Response of a Spring-Mass-Damper System  v1g 0 


x:= 
The Mathcad code to solve higher-order state space systems  FK 
0
has an additional assignment statement to create a vector of
the state variables. We will illustrate the calculation using the The state equations are:
spring-mass-damper system, Fig. A7.5.
The first Mathcad instructions are identical to those of the  −b 1  1
first-order mass-damper system, except that there is now a D(t, x) : =  M M ⋅ x +  M ⋅F

   
spring constant K. K 0  0

M: = 3 b : = 0.2 K: =1 F : = 10 Finite difference computations of higher-order systems


can go unstable if the time step Δt is too large. As a rule of
Assume that the system is at rest and relaxed. Define a vector thumb, if the eigenvalues of complex conjugates, 200 steps
of the initial conditions of the state variables. It is essential per period of the (of the smallest period in the system if four
that the state variables are identified in the Mathcad work- order) is sufficient for a stable and accurate calculation. If
sheet. A common error is to not identify which state variable the plots of the state equations blow up (head off to positive
is which and then forget and use the wrong state variable in a or negative infinity) or if Mathcad indicates an error in the
subsequent calculation. The state variables can be identified rkfixed() function, either decrease the duration of the calcu-

Fig. A7. 5  a force source-spring-


mass-damper system. b Linear
a x,v b 1
graph of the force source-spring- b
mass-damper system
F(t)
M F(t) b K M

K
g
MATLAB’s Runge–Kutta Solver ode45() 461

Fig. A7.6   Response of the state 3 15


2.938 11.085
variable v1g and FK to a step input
10 N
2
10

v.1g 1 F.K

− 0.319 0
−1 0
0 10 20
0 t.MbK 25

Fig. A7.7   Response of output 3 15


2.938 11.085
variables v1 g, FK, FM and Fb to a
step input of 10 N
2 10

F .K

v.1g F .M
1 5
F .b

0 0

−0.319 − 2.59
−1 −5
0 10 20
0 t.MbK 25

lation, or increase the number of calculations, to decrease the MATLAB’s Runge–Kutta Solver ode45()
time step. If the calculation is still “numerically unstable,”
check for a sign error in the Mathcad code or the reduction of MATLAB’s general purpose Runge–Kutta solver function
the state equations. A single sign error is sufficient to create is named ode45(). If it helps you remember the name, the
a numerical instability. 45 stands for “fourth-order” overall error, “fifth-order” local
We will increase to duration of the calculation to 25 sec error.
and the number of calculations to 2,000 for the rkfixed() One of the arguments of the function ode45() is a user-
solver function. written MATLAB function which contains the state equa-
tions to be solved. A MATLAB user-written function re-
ZMbK : rkfixed(x, 0, 25, 2000, D) sembles a short program or script except for the syntax of
the first line of the code is an instruction, which “declares”
Extract the time and state variable vectors (in the computer the function. The declaration, or first line, of a MATLAB
science meaning) from the array ZMbK. user-defined function begins with the key word “function”
followed by an assignment statement where the left side is
t MbK : = ZMbK 0 v1g : = ZMbK 1 FK : = ZMbK 2 a variable name and the right side is the function name fol-
lowed by the function’s argument list. The three dots or peri-
The output variables are calculated using values of the state ods “…” are ellipses, the mathematical symbol for continua-
variables from the solution of the state equations. Often, the tion of a series. Ellipses are MATLAB's operator to continue
state variables are included as output variables, as below. a statement on the following line.

 v1g   1 0  0 function variableName = functionName(argument1,…


F   0 1   v   0 argument2, argument3, additionalArguments)
 K:=  ⋅  1g  +   ⋅ F
F
 M  − b − 1  FK   1
 F   b 0   0 The above syntax is generic for any function one wishes
b
to create. The function containing the state equations to be
solved by MATLAB’s Runge–Kutta solver ode45() expects
462 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

just two arguments. The first argument is the independent An end statement completes the function. Collecting the
variable t for time in our application. The second argument lines of code without the interspersed text:
is the variable name of the state vector x. Attempting to pass
other arguments to the function, which one can do with a function dx = StateEq(t,x,) % Function Declaration
normal user-written MATLAB function, will not work with % Create and zero the column vector dx with one row
the ode45() solving function. Parameter values and input dx = zeros(1,1);
functions must be written into to the user-written function. M = 3; % Mass
They cannot be passed to it as arguments. b = 2; % damping
F = 10: % Force
dx(1) = (-b/m)*x(1) + (1/m)*F; %State Eq
Step Response of a Linear Mass-Damper System end

We will illustrate creation of a user-written function and its The user-written MATLAB function is then saved using the
solution, first with a mass-damper and then with a spring- function’s name and followed by the extension “.m.” In the
mass-damper system. The state equation of the mass-damper command window, the following command runs MATLAB’s
system shown in Fig. 3.13 is repeated below for reference: ode45() Runge–Kutta solver and writes the time data to the
vector T and the values of the state variable to vector X.
F (t ) dv1g b dv1g b F (t )
= + v1g → = − v1g + [T,X]=ode45(@functionName,[tstart tend],[vector …
M dt M dt M M of the initial values of state variables])

The state variable is v1g, the velocity of the mass. The square brackets vectors. The elements in a vector can be
The first step is to create a MATLAB user-defined func- separated by a blank space or a comma. Rows are separated
tion containing the state equations in MATLAB’s editor. The by a semicolon. Note that the name of user- written function
function declaration begins with the keyword “function.” The with the state equations is preceded by the at symbol, @.
statement needs a variable name for the output of the function This is essential.
and a name for the function itself. We will name the variable We will solve the state equation of the mass-damper sys-
dx and the function StateEq. There are two arguments of the tem out to 12 sec and use zero as the initial value of the
function. They must be the independent variable t for time in velocity of the mass.
our applications, and the state vector x in that order.
[T,X] = ode45(@StateEq,[0 12],[0])
function dx = StateEq(t,x,) % Function Declaration
The vector T contains the times of the calculations and X
The first line of code within the function is to define the vari- contains the values of the state variable at those times. The
able dx as a vector variable with the number of elements vectors T and X are plotted with the command:
equal to the number of state variables, one in this case. The
instruction used to create (or allocate) a vector or array and plot(T,X)
fill it with zeros is named zeros( r, c), where r is the number
of rows and c is the number of columns. A column vector has The result is shown in Fig. A7.8. Right clicking after having
one column. The instruction to create the column vector dx selected an object brings up a context menu for editing the
with one element with the initial value of zero is: plot. Select axes to label them and change their font size.
Increase the thickness of the output trace.
% Create and zero the column vector dx with one row
dx = zeros(1,1);
Step Response of a Non-Linear Mass-Damper
The state equation is assigned to the element of the only ele- System
ment vector dx, dx(1). The state variable is also an element
of the state vector x. It is the only element in the state vector We can edit the state equations and add some types of non-
x, x(1). linearities, such as a power law viscosity for the damping.
We will square the velocity of the mass, x(1), in the state
M = 3; % Mass equation.
b = 2; % damping
F = 10; % Force function dx = NLStateEq(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
dx(1) = (-b/m)*x(1) + (1/m)*F; % State Eq. dx = zeros(1,1);
MATLAB’s Runge–Kutta Solver ode45() 463

Mass−Damper Step Response Linear and Non−Linear Mass−Damper Step Responses


5 5
4.5
4 4
V1g, m/sec

3.5 Linear

v1g, m/sec
3 3 Non−linear
2.5
2 2
1.5
1 1
0.5

0 0 2 4 6 8 10 12 0
0 2 4 6 8 10 12
t,sec t, sec
Fig. A7. 8   Step response of a linear mass-damper system, M = 3, b = 2, Fig. A7.9   Linear and non-linear step response of a mass-damper sys-
F = 10 tem, M = 3, b = 2, F = 10

M = 3; % Mass if(t<2)
b = 2; % damping F = 5;
F = 10: % Force elseif(t<4)
dx(1) = (-b/m)*x(1)^2 + (1/m)*F; %State Eq F = -5;
end else
F = 0;
Save this function as NLStateEq.m and run the Runge–Kutta end;
solver ode45() with different names for the output time and dx(1) = (-b/m)*x(1)^2 + (1/m)*F; %State Eq.
state variable vector than used previously. end;

[Tnl,Xnl] = ode45(@NLStateEq,[0 12],[0]) Save the edited function as PulseStateEq. m. Run the Runge–
Kutta solver ode45() and plot the result with the commands:
We can now plot the linear and the non-linear step response
of the mass-damper system on the same plot with the com- [T,X] = ode45(@PulseStateEq,[0 12],[0])
mand:
plot(T,X)
plot(T,X,Tnl,Xnl)

Response of a Linear Mass-Damper System Response of a Mass-Damper System to Sinusoidal


Subjected to a Pulse Train Inputs

The independent variable t can be used in the state equations We use the independent variable t to drive the system with
for the input function. We will use an if, elseif, else structure the input F (t ) = 10 sin(2t )
to create the input pulse:
function dx = SinStateEq(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
F (t ) = 5 us (t ) − 10 us (t − 2) + 5 us (t − 4) . dx = zeros(1,1);
M = 3; % Mass
Edit the first-order linear function StateEq(t, x) to add the b = 2; % damping
if, elseif, and else, shown below to calculate the input force dx(1) = (-b/m)*x(1)^2 + (1/m)*10*sin(2*t); %State Eq
F( t). end;

function dx = PulseStateEq(t,x,) % Function Declaration Save the function as SinStateEq.m, solve it with ode45() and
% Create and zero the column vector dx with one row plot it.
dx = zeros(1,1);
M = 3; % Mass [T,X] = ode45(@SinStateEq,[0 12],[0])
b = 2; % damping
% Force Pulse if, elseif, else plot(T,X)
464 7  Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Pulse Response of Linear Mass−Damper System Input F(t)=10sin(2t) Linear Mass−Damper System
2 ode45( ) option ’refine’ set to 8
3
1.5
2
1
v1g, m/sec

v1g, m/sec
0.5 1

0
0
−0.5
−1
−1

−1.5 0 2 4 6 8 10 12 −2
0 2 4 6 8 10 12
t, sec t, sec

Fig. A7.10   Response of the linear mass-damper system, system, Fig. A7.12   Response of the linear mass-damper system, M = 3, b = 2,
M = 3, b = 2, to the input pulse F( t) to the input F(t) = sin (2t). The “refine” option was increased from its
default of four to eight
if(t<2)
F = 5;
elseif(t<4) than decreasing the time step. The refine option instructs
F = -5; the ode solver interpolate within a time step. To the user, it
else appears that the time step has been decreased. The default
F = 0; valve of refine for ode45() is four. We will double the value
end; to eight. The syntax is:

Input sin(2t), Linear Mass−Damper System options = odeset('refine',8)


2.5
2 The variable name, options, is the last argument in the argu-
1.5 ment list of ode45
v1g, m/sec

1
0.5 [T,X] = ode45(@SinStateEq,[0 12],[0],options)
0
−0.5 The “refined” output is plotted in Fig. A7.12. The appearance
−1 of the trace is improved. It no longer has straight segments.
−1.5
−2
0 2 4 6 8 10 12
t, sec
Step Response of a Spring-Mass-Damper System

Fig. A7.11   Response of the linear mass-damper system, M = 3, b = 2, Our final example will be to solve for the step response of
to the input F( t) = sin(2t). Notice the slightly blocky shape of the trace the state variable and a number of output variables of the
second-order linear spring-mass-damper system, shown in
The quality of the response, Fig. A7.11, is disappointing. Fig. 7.5.
What should be pleasingly smooth sine curves are not. There We will illustrate its use with the state equations of the
are straight segments. mass-spring-damper system:
We must delve into the options available for ode45() re-
duce the blockiness. MATLAB has a built in function named  b 1 1
d  v1g   − −  v1g 
odeset() for setting options for their ordinary differential = M M   + M  F
dt  FK    F  
solvers. The function odeset() exists because there are 22 op-  K 0  K  0 
tions for ode solvers and you cannot pass just a single option
value to the solver. It expects all 22 arguments. The func- The MATLAB code requires individual equations. Expand-
tion odeset() will accept a single option and its value. It then ing the matrix form:
creates a file will the remaining option at their default values.
The set of options is assigned to a variable name. dv1g b 1 1
=− v1g − FK + F
The option we will change is titled “refine.” MATLAB dt M M M
has a faster routine for refining the result of an ode solver dFK
= Kv1g
dt
References and Recommended Reading 465

Spring−Mass−Damper Step Response Force FM and Fb, Step Response Spring−Mass−Damper System
20 10

FK
15
FM
5
Fb
V1g, ms/sec

FM, N
Fb, N
10
FK, N

5
V1g
−5
0

−10
−5 0 20 40 60 80 100
0 10 20 30 40 50 60 70 80 90 100 t, sec
t, sec
Fig. A7.14   Step response of the output variables FM and Fb of a linear
Fig. A7.13   Step response of the linear spring-mass-damper system, spring-mass-damper system, M = 3, b = 0.2, K = 1 to the input F = 10
M = 3, b = 0.2, K = 1 to the input F = 10

It is always worth the time to define which state variable is The time values are in the vector variable T. Its length can be
which. In this case, v1g ≡ x1 and FK ≡ x2 . determined by the function length(). The state variables v1g
We will name the variable dx and the function StateEqs. and FK are the columns in the state vector X and should be
The two arguments of the function must be t for time and x extracted and assigned to individual vector variables. MAT-
for the state vector, in that order. LAB’s “colon operator”,:, is a wildcard to force a variable to
The first line of code within the function is to define the take on every value in a range. The colon operator succes-
variable dx as a vector with the number of elements equation sively takes on the values of the row indices, allowing the
to the number of state variables, two in this case. The instruc- columns of the array to be copied to vectors without the use
tion to create the column vector dx with two elements with of a for end loop.
the initial values of zero is:
v1g = X(:,1); % Extracts column 1 from the array X
function dx = StateEqs(t,x,) % Function Declaration FK = X(:,2); % Extracts column 2 from the array X
% Create and zero the column vector dx with one row
dx = zeros(1,1); The output variables are calculated with a for loop.
M = 3; % Mass
b = 0.2; % damping b = 0.2; % Damping
K = 1; % Spring F = 10; % Step input
F = 10; % Force N = length(T); % Determines length of time vector T
dx(1) = (-b/M)*x(1) + (-1/M)*x(2) + (1/M)*F; FM = zeros(N); % Creates and zeros vector FM
dx(2) = K*x(1); Fb = zeros(N); % Creates and zeros vector Fb
end % for loop to calculate output equations
for n=1:N;
The output equations are purely algebraic equations. They FM(n)=-b*v1g(n)-FK(n)+F;
are solved after the state equations have been solved and the Fb(n)=b*v1g(n);
state variables are known. We must expand the output equa- end
tions from vector-matrix form into a set of equations in order
to program them.
References and Recommended Reading
 v1g   1 0 0 v1g = v1g
Ogata, K (2009) Modern control engineering, 5th edn. Prentice-Hall,
F   0 1  v   FK = FK
 K=   1g  + 0 F →
Englewood Cliffs
 FM   − b −1  FK  1  FM = − bv1g − FK − F Rowell D, Wormley DN (1997) System dynamics: an introduction.
      Prentice-Hall, Upper Saddle River
 Fb   b 0  0 Fb = bv1g
Finite Difference Methods and MATLAB
8

Abstract
Finite difference approximations are the foundation of computer-based numerical solutions
of differential equations. Numerical solutions, also known as numerical methods, are es-
sential to solve non-linear differential equations. The state-space representation of dynamic
systems requires numerical solution. The Euler method is the simplest finite difference
method and is used to introduce the concepts. The more accurate Runge–Kutta algorithm is
also more involved. Fortunately, the use of the Runge–Kutta method is straightforward. The
Appendix to the chapter is a brief introduction to computer programming and programming
in MATLAB for those who need it.

8.1 Finite Difference Approximation It is good practice to test the effect of the size of the time
of Differential Equations step on the results of the calculation, by reducing ∆t by a
significant fraction, and rerunning at least part of the calcu-
We will approximate the differential equations, which de- lation. If there is no difference, then the time step is small
scribe the dynamics of energetic systems using a “finite dif- enough. If the two calculations differ, reduce the time until
ference” approximation. Recall from your introduction to there is no significant difference in the results.
calculus, the definition of a derivative of a variable, say v, A finite difference approximation of a differential equa-
with respect to time as tion is created by replacing all of the differentials with fi-
nite differences. For example, the differential equation with
∆v dv force, F( t), is the input and velocity, v( t), is the output
lim =
∆t → 0 ∆t dt
dv(t )
F (t ) = 3 + v(t )
In a finite difference approximation of the derivative, the in- dt
finitesimal ∆t produced by evaluating lim is replaced with
∆t → 0
a small but finite ∆t: which is approximated as

 ∆v dv ∆v(t )
≈ (8.1) F (t ) = 3 + v(t )
∆t dt ∆t

How small a ∆t is small enough for a reasonable approxima- Notice this equation expresses time in two different forms.
tion of the derivative? That depends on the time scale of the The variables, F(t) and v( t), are written as functions of con-
dynamic response of the system. ­Using the time constant, tinuous time, but the derivative is approximated by using a
τ, of a first-order system and the period, T, of an oscillatory finite (or discrete) time, ∆t. We cannot evaluate this equation
­second-order system as time scale of the respective system’s at any arbitrary time, t, but only at times which are multiples
dynamic responses, the upper limits of ∆t should be approxi- of our finite time, ∆t,
mately ∆ t ≤ τ / 20 and ∆ t ≤ T / 200 so that plots that are sup-
posed to be smooth curves are smooth curves.  t = n ∆t (8.2)

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_8, © Springer Science+Business Media New York 2014 467
468 8  Finite Difference Methods and MATLAB

Fig. 8.1   Forward-stepping finite


difference calculation as a func-

{
v((n+1)∆t) v((n+1)∆t)
tion of discrete time
∆v(n∆t)
Velocity v(n∆t) v(n∆t)
Continuous
Variable v((n-1)∆t) v((n-1)∆t)
∆v((n-1)∆t)
{
v((n-2)∆t) v((n-2)∆t)
∆v((n-2)∆t)
{

∆t ∆t ∆t
0
0 (n-2)∆t (n-1)∆t n∆t (n+1)∆t

Time
Discrete Variable

where n is an integer. Time calculated as a multiple of a time Substituting our expression for ∆ v(n∆ t ) , Eq. 8.4, into Eq. 8.5
step, ∆t, is known as “discrete time,” as opposed to continu- yields
ous time. Substituting n ∆ t for t:
 F (n ∆ t ) v(n ∆ t )
v((n + 1) ∆ t ) = ∆t − ∆ t + v(n ∆ t ) (8.6)
 ∆ v(n ∆ t ) 3 3
F (n ∆ t ) = 3 + v(n ∆ t )
∆t (8.3)
The calculation “steps” forward in time, by calculating the
change in velocity, ∆v(n∆ t ), over the time period, ∆t, and
This approximation of the differential equation, expressed then adding it to the velocity at the beginning of the time
using finite differences, is called a “difference” equation. period, to calculate the velocity at the end of the time period.
The time period, ∆t, is known as the “time step.” The for-
ward stepping solution of a difference equation is illustrated
8.2 Euler Method, Forward Stepping, in Fig. 8.1.
Finite Difference Algorithm The time axis is identified as a “discrete variable,” which
means it can only take on certain values. In this example,
A recursive calculation is an algorithm, in which the new time must be an integer product, or multiple, of the time step,
value is calculated from past values. A forward-stepping, fi- ∆t. In contrast, the velocity axis is identified as a “continu-
nite difference calculation is a recursive algorithm, where the ous variable.” There is no restriction on the value the veloc-
new value calculated from the past values is the value of the ity variable can take.
output variable for the next forward step in time. The forward-stepping Euler algorithm will be illustrated
We can rearrange the difference equation, Eq. 8.3, to by calculating two steps of the difference equation
allow us to calculate the change in the output variable, ∆v,
at time, t = n ∆ t F (n ∆ t ) v(n ∆ t )
∆ v(n ∆ t ) = ∆t − ∆t
3 3
 F (n ∆ t ) v(n ∆ t )
∆ v(n ∆ t ) = ∆t − ∆t (8.4) To begin the calculation, we must know the initial value
3 3
of the output variable. We will assume v(0+ ) = 0. We will
impose a step input on the system of a magnitude of ten,
∆ v(n∆ t ) is the change in the velocity, v(n ∆ t ), over the finite F (t ) = 10. Lastly, we will use a time step, ∆ t = 0.5.
time, ∆t, from time, t = n ∆ t, to time, t = (n + 1) ∆ t. Hence, The first calculation is the change of the output variable
adding the current value of v to the “forward” change ∆v at time, t = 0, ∆ v(0.0):
yields the next value of v:
F (0.0) v(0.0)
∆v(0.0) = ∆t − ∆t
v((n + 1) ∆ t ) = ∆ v(n∆ t ) + v(n ∆ t )
(8.5) 3 3

 10   0
∆v(0.0) =  0.5 −   0.5 = 1.67
 3   3
8.2  Euler Method, Forward Stepping, Finite Difference Algorithm 469

Add the change of the output value, ∆ v(0.0), to its existing


v(n+1)

{
v(n+1)
value, v(0.0), to yield the value of the output variable for
∆v(n)
next discrete time, t + ∆ t = 0.0 + 0.5 = 0.5. This is the for-
ward “step” in time to the velocity at time, v(0.5): Velocity v(n) v(n)

v(0.5) = v(0.0) + ∆v(0.0) v(n-1) v(n-1)


∆v(n-1) {
v(0.5) = 0 + 1.67 = 1.67 v(n-2) v(n-2)
∆v(n-2) {
The process now repeats. Loop back to the first calculation.
Repeat the calculation using the value of v(0.5) to calculate
∆ v(0.5) : 0
0 n-2 n-1 n n+1
F (0.5) v(0.5) Step Number
∆v(1.0) = ∆t − ∆t
3 3
Fig. 8.2   Forward-stepping finite difference calculation, referencing a
 10   1.67  value by its step number, n
∆v(1.0) =   0.5 −  0.5 = 1.39
 3  3 
and written using the counter variable, n, as the index for the
Add v(0.5) and ∆ v(0.5) to again “step” forward in time to input vector, F, and the output vector, v:
the velocity at time, v(1.0):
 F (n ) v(n ) (8.7)
v(1.0) = v(0.5) + ∆v(0.5) ∆ v(n ) = ∆t − ∆t
3 3
v(1.0) = 1.67 + 1.39 = 3.06
The equation written as a function of discrete time is how a
Recursive algorithms must retain previous values of the cal- human may think of a step in the calculation, but the calcula-
culation in order to calculate the next value. In our applica- tion is programmed using the vector form of Eq. 8.7, shown
tions, i.e., the solution of a system equation or a set of state in Fig. 8.2. Discrete time is not needed to perform the calcu-
equations, we will have to create a complete record of the lation, only to plot the result on a time axis.
results of the calculation in order to plot them. We will store Since computer code is restricted to standard keyboard
the values of the output variable and time in vectors. This characters, there are no Greek characters. We will rewrite
raises the question how to modify the calculation we just the finite difference approximation of the derivative using
performed to store the values we need in vectors. Although “D” for ∆ in ∆x( n). Likewise, we will use T rather than ∆t.
we wrote the finite difference equation, Eq. 8.4, in discrete Be careful with the variable, “T”. “T” can also represent the
time, t = n∆t, we do use the value of the discrete time in the period of an oscillation. Comments in the MATLAB code
calculation. It is sufficient to merely identify the previous which identify variables are a great help to understanding an
value of the output variable. A step in the calculation can be algorithm.
identified by either the time corresponding to the step, n∆t,
or, more simply, by just value of the integer variable, n. It is
easier to program the calculation as a sequence with the step 8.2.1 MATLAB Programming of the Euler
number, n, as the index of the vector variables. Method, First-Order System
The standard mathematical notation for an element in a
series is a subscript, as is used in Mathcad to index vector The first MATLAB program, or script, will be the solution of
elements. Unfortunately, computer code uses neither sub- the finite difference we just solved manually:
scripts nor superscripts. The characters of computer code are
those found on a standard keyboard. The syntax used to de- dv(t )
F (t ) = 3 + v(t )
note a vector index in MATLAB is to append the subscript in dt
parentheses to the end of the vector variable name, i.e., the
nth element of the vector, x, is x( n). Notice the notation for a To make the program of more general use, we will code it in
vector is identical to the notation used to denote a function the standard notation of a state equation, using u( t) for the
of one variable. input; x( t) for the state variable; a for the coefficient multi-
For example, consider the same first order written as a plying the state variable; and b for the coefficient multiply-
function of discrete time: ing the input:

F (n ∆ t ) v(n ∆ t ) dx(t )
∆ v(n ∆ t ) = ∆t − ∆t bu (t ) = + ax(t )
3 3 dt
470 8  Finite Difference Methods and MATLAB

Rearranging both equations into the standard form of a state


Start
equation

dv(t ) 1 1 dx(t )
= − v(t ) + F (t ) → = ax(t ) + bu (t ) Input values
dt 3 3 dt N, t end , a, b, x(1), u

yields a = –1/3 and b = 1/3.


Calculate time step
We also need to specify the number of iterations, or steps N, t end
T = ___
that the program is to execute and the time step, T. Rather N
than assign a value to the time step, T, it is more convenient
to specify the duration, tend, of the calculation and let the
Initialize index variable n
computer calculate the time step, T: and time vector t(n)
n=1
 tend t(1) = 0
∆t ≡ T = (8.8)
N steps

In order to plot the results, we must create vectors of time


and the output, which, in this case, is the state variable, x. Calculate the change Dx
The vectors must have the same number of elements. MAT- Dx = a x(n) T + b u T
LAB’s vector index, n, begins at one, not zero, but we want
time to begin at zero. Consequently, we must subtract one
from n to calculate the time of a step.
Calculate forward step
x(n+1) = Dx + x(n)
 t (n) = (n − 1) · T (8.9)

Equation 8.9 illustrates an unfortunate consequence of not


being able to use subscripts in computer code. The paren- Calculate time t(n+1)

thetical n in t(n) is the index of the vector, t. The parentheti- t(n+1) = n T

cal ( n − 1) is a factor in the product, (n − 1) · T . Be careful


interpreting parentheses.
A flow chart of the Euler method for solving one-state
equation is Fig. 8.3. The counter variable, n, is compared Increment counter by one No
n=N?
with the number of steps, N, and the execution of the loop is n=n+1
terminated if n=N. The easiest way to program this is to use a
for loop to control the number of iterations through the loop, Yes
as shown in the following code.
Plot results
% Euler Method, First-order system, for loop
N = 2000 % Number of steps
t_end = 20 % End time of calculation
a = -1/3 % Coefficient of state variable Finish

b = 1/3 % Coefficient of input variable


x(1) = 0 % Initial value of the state variable Fig. 8.3   Flowchart of the Euler method solution of a first-order sys-
u = 10 % Input function, a step of 10 tem’s state equation
T = t_end/N % Time step
t(1) = 0 % Initial value of the time vector Note that the change in the state variable is now a scalar
for n = 1:N; % Finite difference loop
variable, Dx, rather than a vector variable, Dx(n). There is
Dx = a*x(n)*T+b*u*T;
no need to store the change in the state variable for each
x(n+1) = Dx+x(n);
time step, unless there is a desire to plot it. The scalar vari-
t(n+1) = n*T;
able, Dx, is assigned the current value for each iteration of
end
plot(t,x) the loop.
An alternative method uses a while loop to control the
number of iterations.
8.2  Euler Method, Forward Stepping, Finite Difference Algorithm 471

Fig. 8.4  a Force source act-


ing on a spring-mass-damper a x,v b 1
system. b Linear graph of the b
spring-mass-damper system
F(t)
M F(t) b K M
K
g
Frictionless rollers

% Euler Method, First-order system, while loop The state equations expressed in vector-matrix form are:
N = 2000 % Number of steps
t_end = 20 % End time of calculation  b 1 1
d  v1g   − −   v1g   
a = -1/3 % Coefficient of state variable = M M   M F +
dt  FK    F  
b = 1/3 % Coefficient of input variable  K 0  K  0 
x(1) = 0 % Initial value of the state variable
u = 10 % Input function, a step of 10 We can derive the finite difference equations for this sys-
T = t_end/N % Time step tem, by either operating on the individual state equations,
t(1) = 0 % Initial value of the time vector or on the state equations in vector-matrix form. However,
while (N+1) > n; we need the individual equations to program the equations.
Dx= a*x(n) T+b*u*T; The first step is to approximate the derivatives as finite dif-
x(n+1) = Dx+x(n);
ferences:
t(n+1) = n*T;
n = n+1; ∆v1g b 1 1
end = − v1g − FK + F
plot(t,x) ∆t M M M
∆ FK
= Kv1g
∆t
8.2.2 Euler Method Solution of Second-Order Multiplying both sides by the time step, ∆t, yields a set of
State Equations simultaneous difference equations:

We will illustrate the solution of a set of state equations using b 1 1


∆v1g = − v1g ∆ t − FK ∆ t + F ∆t
the Euler method for the force source spring mass-damper M M M
system, Fig. 8.4. ∆ FK = Kv1g ∆ t
Energetic Equations
Continuity F − FM − FK − Fb = 0 As we did with the Euler solution for a single-state equa-
tion, we will write the program for a system of two state
Compatibility v1g = v1g equations in standard notation. We will begin in vector-
dv1g dFK matrix form.
Elements FM = M = Kv1g Fb = bv1g
dt dt
d  x1   a11 a12   x1   b1 
1 F2 = + u
Energy E sys = E M + E K E M = Mv12g E K = K dt  x2   a21 a22   x2  b2 
2 2K
The state variables in this formulation are the energy storage Our first step is to approximate the differential operator
variables, the velocity of the mass, v1g, and the force acting with the finite difference operator. In other words, replace
through spring, FK. The reduction of the energetic equations d with ∆.
yields the following state equations:
∆  x1   a11 a12   x1   b1 
= + u
dv1g b
= − v1g −
1
FK +
1
F
∆ t  x2   a21 a22   x2  b2 
dt M M M
dFK
= Kv1g
dt
472 8  Finite Difference Methods and MATLAB

Next, multiply both sides by ∆t and distribute the finite dif-


Start
ference operator, ∆, onto the elements of the state vector.
We have a set of finite difference equations for the two state
variables in vector-matrix form. Input values
N, t end , a11, a12, a21, a22
 ∆ x1   a11 a12   x1  b  b1, b2, x1(1), x2(1), u(n)
 ∆ x  = a    ∆t +  1  u∆t
 2   21 a22   x2  b2 
Calculate time step
Expand the vector-matrix notation into a set of two simulta- t end
T= ___
neous finite difference equations, in order to write the MAT- N
LAB code.

 ∆ x1 = ( a11 x1 + a12 x2 + b1u ) ∆ t Initialize index variable n


(8.10) and time vector t(n)
∆ x2 = ( a21 x1 + a22 x2 + b2 u ) ∆ t n=1 t(1)=0

Change the notation and express Eqs. 8.10 with the char-


acters available on a standard keyboard, without ∆ or sub-
scripts. Rename the variables, x1 ≡ x1, ∆ x1 ≡ Dx1, x2 ≡ x 2,
and ∆x2 ≡ Dx 2. Rewrite the coefficients without sub- Calculate Dx1 and Dx2
scripts, as a11 ≡ a11 and b1 ≡ b1 Use the vector index, n, Dx1 = (a11 x1(n) + a12 x2(n) + b1 u(n)) T
and the parenthetical notation for a vector subscript used Dx2 = (a21 x1(n) + a22 x2(n) + b2 u(n)) T
in MATLAB to rename the variables and coefficients of
Eqs. 8.10:
Calculate forward step
 Dx1 = ( a11 · x1(n) + a12 · x 2(n) + b1 · u (n) ) · T x1(n+1) = Dx1 + x1(n)
(8.11)
Dx 2 = ( a 21 · x1(n) + a 22 · x 2(n) + b 2 · u (n) ) · T x2(n+1) = Dx2 + x2(n)

Equations 8.11 are both evaluated before the vectors, x1( n)


and x2( n), are “updated” by adding the change, Dx1 and Calculate time t(n+1)
Dx2, to the current values, x1( n) and x2( n), to yield the value t(n+1) = nT
for the next, n + 1, calculation.

 x1(n + 1) = Dx1 + x1(n)


(8.12) Increment counter No
x 2(n + 1) = Dx 2 + x 2(n) n=N?
n=n+1

The index, n, is then incremented and the loop repeated, until Yes
n = N. Combine Eqs. 8.11 and 8.12 as:
Plot results
 x1(n + 1) = ( a11 · x1(n) + a12 · x 2(n) + b1 · u (n) ) · T + x1(n)

x 2(n + 1) = ( a 21 · x1(n) + a 22 · x 2(n) + b 2 · u (n) ) ·T + x 2(n)


Finish
(8.13)
Fig. 8.5   Flowchart of the Euler method algorithm solution for a two-
state system
The flowchart of the second-order program is Fig. 8.5.
The following is the MATLAB code of the Euler method
solver for a second-order system with the state equations and t_end = 100.0 % Duration of calculation
parameters of the spring-mass-damper system. M = 3.0 % Mass
K = 1.0 % Spring constant
%Euler Method, Second-order system, B = 0.2 % Damping coefficient
%spring-mass-damper % Coefficient matrix elements
N = 2000 % Number of steps a11 = -B/M
a12 = -1/M
8.2  Euler Method, Forward Stepping, Finite Difference Algorithm 473

100 force, when v1g = 2 m/sec. We will model a thickness of the


lubricating film as constant and independent of the velocity
80 of the mass. Assume a thickness of h = 0.02 mm for the lu-
bricating film. We need an area for the shearing surfaces
µ(γ) 60
and will use A = 10 cm × 10 cm = 0.01 m 2 . We begin by
Viscosity 40 equating the damping forces at v1g = 2 m/sec to calculate
the value of µ 0:
20
3 3
0  v1g  2  v1g  2
0 5 10 15 20 Fb = bv1g = Aµ 0   v1g → b = Aµ 0  
 h  h
Shear Strain Rate, γ
Fig. 8.6   Power law viscosity, Eq. 8.14  N ⋅ sec  −
3

 0.2    2 m   2
3

b  v1g  2
m   sec  
µ0 = =
A  h  0.01 m 2
 
a21 = K  0.00002 m 
a22 = 0.0 1
% Input vector Elements µ 0 = 6.3 × 10 −7 kg ⋅ m 2 ⋅ sec 2
b1 = 1/M
b2 = 0.0
%Initial values of the state variables Rearrange the elemental equation for the power law damper
x1(1) = 0.0 % Velocity v1g to collect the velocity term
x2(1) = 0.0 % Force FK 3

u = 10.0 % Input step magnitude  v1g  2


Fb = bv1g → Fb = µ 0 A   v1g
T = t_end/N % Time step  h
t(1) = 0.0 % Initial value of time vector
for n = 1:N;  52  
v1g Aµ 0  5
Dx1 = (a11*x1(n) + a12*x2(n) + b1*u)*T; Fb = µ 0 A  3  =  3  v1g 2
 
Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;  h 2   h 2 
x1(n+1)= Dx1 + x1(n);
x2(n+1)= Dx2 + x2(n); yielding
t(n+1) = n*T;
end;  Aµ  5 5
Aµ 0
plot(t,x1,t,x2) 0
Fb =  3  v1g 2 → Fb = bN .L.v1g 2 where b= 3
 2 
h h2

8.2.3 Example: Euler Method Solver, The state equations of the spring-mass-damper system are:
Non-Linear State Equation
∆ v1g b 1 1
= − v1g − FK + F
MATLAB and Mathcad both include a “built-in” Runge– ∆t M M M
Kutta solver function. The reason to program a finite dif- ∆ FK
ference solver is for the inclusion of non-linear system ele- = Kv1g
∆t
ments. In this example, rather than assuming that lubricating
film has ideally linear viscous friction, we will use a power 3
law model for the viscosity of the film, Fig. 8.6 and Eq. 8.14: Substitute µ0 / h 2 for b and raise v1g to the 5/2 power:

 3
 Aµ   5  1
µ (γ ) = µ 0 γ 2 (8.14) ∆v1g 1 1
=−  3 0  v1g2  − FK + F
∆t M  2  M M
h
We will derive a power law viscosity, such that the damp- ∆ FK
ing force for the linear damper with damping coefficient = Kv1g
∆t
b = 0.2 N · sec/m and the power law damper have the same
474 8  Finite Difference Methods and MATLAB

Multiply both sides by the time step, ∆t we can rewrite the expression into the product of x and the
absolute value of x raised to a fractional power.
  Aµ   5  1 
1 1
∆v1g =  −  3 0  v1g2  − FK + F  ∆t α α 
 β − 1
 α −β 
 β 
 M  2  M M 
 xβ = x · x =x· x
h
∆ FK = Kv1g ∆t
This allows us to maintain the sign of velocity and take the
square root:
Our second-order state equation solver is written in terms of
coefficients, ar,c, and state variables, x1 and x2. Hence,   Aµ  3
Fb =  3 0  v1g v1g 2
(8.16)
 2 
1  Aµ  h
a11 = −  3 0 
M  2 
h The state equations are now:

and    3

∆ x1 =  a11 x1 x1  
2 + a12 x2 + b1u  ∆t
  (8.17)
   5

 
∆ x1 =  a11 x1 2  + a12 x2 + b1u  ∆t (
∆ x2 = a 21 x1 + a 22 x2 + b2 u ∆t )
  (8.15)
∆ x2 = ( a21 x1 + a22 x2 + b2 u ) ∆t The syntax of the absolute value function is abs( x), where x
can be a real number, complex number, variable, or function.
If we were to code this set of finite difference equations, and In this case, we take the absolute value of an element of a
run it in MATLAB, we would receive an error message vector.
MATLAB code for the Euler method solver for a second-
Warning Imaginary parts of complex X and/or Y order system with the power law viscosity follows:
arguments ignored.
% Second-order system with power law viscosity
The problem is that the velocity of the mass, v1g , changes % Euler Method state equation solver
sign, as it oscillates about zero. The power law viscosity ex- N = 2000 % number of steps
pression yields a real value when the velocity is positive, but t_end = 40.0 % duration of simulation
a complex number when the velocity is negative. Raising a % Parameter values
number to an odd power does not change its sign. Conse- M = 3.0 % Mass
quently, when the velocity is negative, we are attempting to K = 1.0 % Spring constant
take the square root of a negative number. mu = 6.3*10^-9 % Viscosity
h = 0.02*10^-3 % Height
The power law model viscosity model, Eq. 8.14, should
% Coefficient matrix elements
have been written using the absolute value of strain rate to
a11 = (-1/M)*(mu/h^(3/2))
avoid computational error:
a12 = -1/M
3 a21 = K
µ (γ ) = µ 0 γ 2
a22 = 0.0
% Input vector elements
The state equation was rewritten to calculate the fluid strain b1 = 1/M
rate in terms of the mass velocity, v1g . We must maintain the b2 = 0.0
sign of the damping force. We cannot take the absolute value % Initial values of state variables
of velocity in the damping force expression, in this case: x1(1) = 0.0 % mass velocity v1g
x2(1) = 0.0 % Force acting through spring
 Aµ  5 %
Fb ≠  3 0  v1g 2
u = 10.0 % step input F
 2 
h T = t_end/N % time step
t(1) = 0.0 % initial value of time vector
Using the relationships: % Difference equation loop
for n = 1:N;
α 1
dx1 = (a11*x1(n) (abs(x1(n))^(3/2)) + a12*x2(n) …
( )
x β = xα β and x (α +1) = x · xα
+ b1*u)*T;
8.3  User-Written MATLAB Functions 475

Friction
Kinetic Friction Force those variables. A mass velocity of v(n) = 0.001 mm/sec,
FC or a micron per second, is certainly small enough that
we would perceive it to be zero. However, the following
MATLAB code:
Static Friction
% Set value of v(n) to 1 micron/sec = 0.000001 m/sec
Velocity
v(n) = 0.000001
% Set logical variables static and kinetic false
static = false
kinetic = false
% Test velocity and set friction condition true
Fig. 8.7   Static and kinetic Coulomb friction model. FC is the maxi-
if(v(n) ==0)
mum value of static Coulomb friction
static = true
else
kinetic = true
dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T; end
x1(n+1)= dx1 + x1(n);
x2(n+1)= dx2 + x2(n); yields kinetic = true, when the condition  v( n) = = 0 is eval­
t(n+1) = n*T; uated since the velocity is not “exactly,” to the numerical
end; precision of the variable, zero. We eliminate the problem of
plot(t,x1,t,x2) needing a value to be exactly zero by defining a velocity
limit, which is sufficiently small that it is, for practical pur-
poses, zero, and then test whether the velocity in question is
8.3  User-Written MATLAB Functions above or below that limit. We must use the absolute value of
v(n) to prevent negative velocities with magnitudes greater
8.3.1  Static-Kinetic Coulomb Friction Model than the zero limit from satisfying the inequality.

Coulomb friction is a more strongly non-linear phenomenon  v(n) < vZeroLimit (8.18)
than power law viscosity, and significantly more difficult to
incorporate into a dynamic model. We will use the familiar, The following MATLAB code with a limit defined for zero
simplified model of static Coulomb friction and kinetic Cou- velocity yields static = true when the magnitude of the veloc-
lomb friction, shown in Fig. 8.7. ity is less than the zero limit.
Static and kinetic Coulomb friction both manifest as func-
tions of the normal force, acting across two surfaces in con- % Define zero velocity limit to 0.1 mm/sec
tact. Coulomb friction results from asperities (microscopic vZeroLimit = 0.0001
high spots) on one surface making contact with asperities % Set value of v(n) to 1 micron/sec or 0.000001 m/sec
on the other surface. Coulomb friction opposes shear motion v(n) = 0.000001
between solid surfaces. Static friction opposes impending % Set logical variables static and kinetic false
motion. Kinetic friction opposes actual motion. static = false
In the abstract, it seems reasonable to define static fric- kinetic = false
tion as applicable for v(n) = 0, and kinetic friction as ap- % Test velocity and set friction condition true
plicable when v(n) ≠ 0. However, this is not practical. We if(abs(v(n)) < vZeroLimit)
static = true
should not use the criterion that the velocity of the mass
else
equals zero, as the condition to distinguish between ki-
kinetic = true
netic and static frictions. An MATLAB if statement con-
end
tains a “conditional statement” which tests an equality or
an inequality, or evaluates a Boolean variable or condition.
It the conditional statement is true, then the instructions in Kinetic friction is modeled as constant force opposing mo-
the if statement are executed. If the condition is false, then tion. We determine the direction of the motion by the sign
they are not. If we test whether two variables are equal, of the velocity variable. All programming languages have
and the variables are numerical variables (as opposed to a function to determine the sign of a number, generically
“logical” variables, which are Boolean), the two quantities known as the “signum” function, contraction of “sign of
will be compared to the numerical precision inherent in number.” MATLAB’s signum function is sign( x). If the
476 8  Finite Difference Methods and MATLAB

sign of x is positive, the function returns one. If the sign the sine function, sin( x), are input–output ­relationships. The
of x is negative, the function returns negative one. The input is the parenthetical quantity, known as the “argument.”
following MATLAB statement calculates the kinetic cou- The function yields, or “returns,” a value. Software functions
lomb friction. are also input–output relationships. The input for a software
function, such as the absolute value function, abs( x), is also
% if kinetic, calculate kinetic friction the argument. The output of the function is the value it “re-
if(kinetic) turns.”
Fkinetic = mu_k*F_N*sign(v(n)) Computer languages allow users to write their own func-
end tions. A computer function differs little from a computer pro-
gram. There are two advantages to using a function. First,
Static Coulomb force is more difficult to incorporate in a dy- functions have fewer instructions and are easier to debug than
namic model for two reasons. First, because static Coulomb programs. Second, functions allow easy reuse of code in the
force opposes impending motion, the model must predict the same program or a different program.
direction of the impending motion, in order to assign the cor- The syntax of using (or “calling”) a function in a com-
rect direction to the friction force. Second, the magnitude of putation follows mathematical convention. We will name
static friction is variable. The magnitude indicated in the plot our Coulomb friction function, “Coulomb.” We will need to
as FC, as per Fig. 8.7, is the maximum static friction, which
can be “mobilized” to resist impending motion. The static Table 8.1   Function Coulomb arguments
friction mobilized falls in the range of ±FC. Velocity of the mass v(n)
The direction of impending motion is determined by the Sum of the input force and the spring force F + FK ≡ Fsum
sign of the sum of the forces acting on a mass. Suppose an Normal force acting across the surface FN
applied force F and a spring force FK act on a mass, in addi- Static coefficient of friction µs
tion to the friction force. The sign of the direction of impend- Kinetic coefficient of friction µk
ing ­motion is:

% If static, determine sign of impending motion of mass “pass” the function the following arguments:
if(static) Write functions with the same process and logic as a
impending = sign(F+FK) MATLAB script (program). First, make a flow chart of the
end logic, then encode it as MATLAB instructions. The flow
chart is Fig. 8.8.
There are two cases for calculating the magnitude of the A function must begin with a MATLAB statement called
static friction. If the value of “completely mobilized” fric- a “declaration,” which tells MATLAB that the following
tion, FCMax = µ s FN , is less than the sum of the non-friction code is (1) a function which will “return,” or provide, the
forces acting on the mass, then FCMax is the friction force. value of a given variable; (2) has a given function name;
The second case applies, when the sum of the non-friction and (3) uses the given list of variables as arguments in its
forces acting on the mass is less than the completely mobi- algorithm:
lized friction, FCMax. The maximum permissible value of the
static Coulomb friction force is the force, which produces function return_variable
static equilibrium, and, hence, equals the opposite of the = function_name(argument or argument list)
sum of the other forces acting on the mass. If the static
Coulomb friction calculated by the algorithm exceeds the In standard mathematical notation, we would calculate the
sum of the non-friction forces acting on the mass, then the Coulomb friction using our function Coulomb, as
mass would accelerate under the action of friction, which
is nonsense. FC = Coulomb (v(n), Fsum , FN , µ s , µ k )

The declaration of our Coulomb friction function in MAT-


8.3.2  Programming a Function in MATLAB LAB provides the same information, preceded by the key-
word, “function”
Although we could include the instructions we need to deter-
mine the Coulomb friction in our MATLAB program, we will function Fc = Coulomb(V, Fsum, Fn, mu_s, mu_k)
write a MATLAB “function” instead. The term, function, has
the same meaning in computer programming, as it does in Note that function declaration expresses the velocity as V,
conventional mathematics. Mathematical functions, such as not as an element, v( n), of the vector, v. We will call (or use)
8.3  User-Written MATLAB Functions 477

Fig. 8.8   Flowchart of the


Function Coulomb
Coulomb friction model with
indeterminate static friction v(n), Fsum ,FN ,µ s ,µ k

Define zero velocity


v zero = 0.001

Yes
|v(n)| > v zero

No

Calculate kinetic Calculate max limit of


Coulomb friction static Coulomb friction
FC = -µk FN sign(v(n)) FC = µs FN

Yes
FC >= |Fsum|

No
FC = -Fsum
FC = -sign(Fsum) FC

Return Coulomb Friction


FC

function Coulomb, each time we pass through the loop of %


finite difference code. We will pass only the current value % If the absolute value of the velocity is greater than
of the velocity of the mass to the function. We will not % the zero limit then calculate the kinetic Coulomb
pass the entire vector, v, to the function. % friction, else calculate the static Coulomb friction
The following is the same logic, shown in the Fig. 8.8 %
flowchart, coded using an else statement. if (abs(V) > vZeroLimit);
Fc = -mu_k*Fn*sign(V);
% Function Coulomb else;
% Returns static or kinetic friction %
% % Calculate the static Coulomb Friction
function Fc = Coulomb(V,Fsum,Fn,mu_s,mu_k); %
vZeroLimit = 0.001; Fc = mu_s*Fn;
478 8  Finite Difference Methods and MATLAB

if (Fc >= abs(Fsum)); other choice is whether to use an assignment statement to


Fc = -Fsum; define a Coulomb friction variable, say Fc, or to include
else; the function call in the state equation. The following code
Fc = -sign(Fsum)*Fc uses the latter method. The function Coulomb is called
end; and the value returned is assigned to the variable, Fc.
end; The friction force is then summed with the input force,
F. First, express the logic in conventional mathematical
Here is the same logic coded, using nested “if” statements. notation:

% Function Coulomb_ifs Fc = Coulomb(V , Fsum, Fn, mu _ s, mu _ k )


% Returns static or kinetic friction  b 1 1 
% ∆v1g =  − v1g − FK + ( Fc + F ) ∆t
 M M M 
function Fc = Coulomb_ifs(V,Fsum,Fn,mu_s,mu_k);
% ∆ FK = Kv1g ∆t
vZeroLimit = 0.001;
% Now, change notation to express these equations using the
% If the absolute value of the velocity is greater than characters available on a standard keyboard. Express the
% the limit then zero limit then calculate the variables, x1, ∆ x1 , x2, and ∆ x2 , as vectors with the index, n,
% kinetic Coulomb friction.
using the standard notation for state equations.
%
if (abs(V) > vZeroLimit);
% Second-order system Euler method state equation
Fc = -mu_k*Fn*sign(V);
% solver with linear damping and static and kinetic
end;
% friction
%
%
% If the absolute value of the velocity is less than or
N = 2000 % number of steps
% equal to the zero limit then calculate the
t_end = 100.0 % duration of simulation
% static Coulomb friction
%
%
% System parameter values
if (V <= vZeroLimit);
M = 3.0 % mass
Fc = mu_s*Fn;
K = 1.0 % spring constant
if (Fc >= abs(Fsum));
B = 0.2 % damping coefficient
Fc = -Fsum;
mu_s = 0.03 % coefficient static of friction
end;
mu_k = 0.02 % coefficient kinetic of friction
if (Fc < abs(Fsum));
%
Fc = -sign(Fsum)*Fc
Fn = M * 9.81 % Normal force
end;
%
end;
% Coefficient matrix elements
a11 = -B/M
The function is used, or called, in our state equation solver a12 = -1/M
program in the same manner, that functions are used in con- a21 = K
ventional mathematical notation. Return to the state equations a22 = 0.0
of the spring-mass-damper system expressed in terms of finite % Input matrix elements
differences: b1 = 1/M
b2 = 0.0
∆v1g b 1 1 % Initial values of state variables
= − v1g − FK + F
∆t M M M %
∆ FK x1(1) = 0.0 % mass velocity v1g
= Kv1g x2(1) = 0.0 % Force acting through spring FK
∆t
%
We see that the Coulomb friction force, FC, can be includ- u = 10.0 % step input F
ed in a state equation for the velocity of the mass, either as T = t_end/N % time step
an additional term with the coefficient, 1/M, or summed t(1) = 0.0 % initial element of time vector
with either the spring force, FK, or the input force, F. The %
% Difference equation loop
8.4  Runge–Kutta Method 479

for n = 1:N; 8.4.1 Two-State, Fourth-Order Runge–Kutta


V = x1(n); %Current mass velocity Algorithm and Code
Fsum = x2(n) + u; %Net force w/o Coulomb friction
% Function call The fourth-order Runge–Kutta algorithm is presented below
Fc = Coulomb(V,Fsum,Fn,mu_s,mu_k); for a system of two-state equations, first in conventional
% Euler Algorithm mathematical notation, then as MATLAB code. The Runge–
Dx1 = (a11*x1(n) + a12*x2(n) + b1*(Fc + u))*T; Kutta algorithm replaces the Euler method in the for loop of
Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
the preceding MATLAB examples:
x1(n+1)= Dx1 + x1(n);
x2(n+1)= Dx2 + x2(n);
t(n+1) = n*T;
f1 = (a x (n) + a 12 x2 (n) + b1u T
11 1 )
end;
plot(t,x1,t,x2)
g1 = (a 21 1 x (n) + a 22 x ( n) + b u ) T
2 2

  f   g  
f 2 =  a 11 ⋅  x1 (n) + 1  + a 12 ⋅  x2 (n) + 1  + b1u  T
  2  2 

8.4  Runge–Kutta Method   f   g  


g 2 =  a 21 ⋅  x1 (n) + 1  + a 22 ⋅  x2 (n) + 1  + b 2 u  T
  2  2 
The Euler method steps forward, by the product of a finite   f   g  
difference approximation of the first derivative of the state f 3 =  a 11 ⋅  x1 (n) + 2  + a 12 ⋅  x2 (n) + 2  + b1u  T
  2  2 
variable and the time:
  f   g  
g3 =  a 21 ⋅  x1 (n) + 2  + a 22 ⋅  x2 (n) + 2  + b 2 u  T
 ∆vn (8.19)   2  2 
vn +1 = vn + ∆t
∆t f4 = ( a ⋅ ( x ( n) + f ) + a ⋅ ( x ( n) + g ) + b u ) T
11 1 3 12 2 3 1

Although the Euler method is simple, it is inaccurate and g4 = ( a ⋅ ( x ( n) + f ) + a ⋅ ( x ( n) + g ) + b u ) T


21 1 3 22 2 3 2
can be “numerically unstable,” meaning, the accumulated
x1 (n + 1) = x1 (n) +
( f1 + 2 f 2 + 2 f3 + f 4 )
errors in the response will eventually lead to the result of
6
positive or negative infinity. The accuracy of the Euler
method can be improved, by reducing the time step to im- ( g + 2 g 2 + 2 g3 + g 4 )
x (n + 1) = x (n) + 1
2 2
prove the approximation of the derivatives. This has the 6
unintended effect, however, of increasing the round-off
error. A reduction in the number of time steps increases the The corresponding MATLAB code follows.
number of computations.
A more accurate and stable finite difference method is %
% Runge-Kutta Difference Equation loop
the Runge–Kutta method, which resembles a Taylor series
% Two State Equations
expansion. The Runge–Kutta method is also known as a
%
“quadrature” method, a name which refers to the use of
for n = 1:N;
graph paper with axes at 90° or “at quadrature.” The meth-
%
od was developed for manual evaluation, when no other f1 = (a11*x1(n) + a12*x2(n) + b1*u)*T;
means of integration was possible. We will use a fixed- g1 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
time step, “fourth-order” Runge–Kutta scheme. The term %
“fourth-order” refers to the successive approximations f2 = (a11*(x1(n)+f1/2) + a12*(x2(n)+g1/2) + b1*u)*T;
of each step that reduce the error to the fourth order. The g2 = (a21*(x1(n)+f1/2) + a22*(x2(n)+g1/2) + b2*u)*T;
Runge–Kutta algorithm improves the accuracy of the for- %
ward step, by estimating half a step first, then using those f3 = (a11*(x1(n)+f2/2) + a12*(x2(n)+g2/2) + b1*u)*T;
results to estimate a full step. The method uses the current g3 = (a21*(x1(n)+f2/2) + a22*(x2(n)+g2/2) + b2*u)*T;
value of the state vector, x(n), as the operating point. An %
intermediate value, x(n + T /2), is then calculated to project f4 = (a11*(x1(n)+f3) + a12*(x2(n)+g3) + b1*u)*T;
forward by half a time step, T/2. This value is then used to g4 = (a21*(x1(n)+f3) + a22*(x2(n)+g3) + b2*u)*T;
create the next intermediate value in a recursive formula. %
The process is repeated. The intermediate values are then x1(n+1)= x1(n)+(f1+2*f2+2*f3+f4)/6;
weighted and summed to calculate a value for the state x2(n+1)= x2(n)+(g1+2*g2+2*g3+g4)/6;
vector, x(n +1). %
t(n+1) = n*T;
end;
480 8  Finite Difference Methods and MATLAB

f1 = (a x (n) + a 12 x2 (n) + a 13 x3 (n) + b1u T


11 1 )
g1 = (a 21 1 x (n) + a 22 x2 (n) + a 23 x3 (n) + b 2 u T )
h1 = (a 31 1 x (n) + a 32 x2 (n) + a 33 x ( n) + b u ) T
3 3

  f   g   h1  
f 2 =  a 11 ⋅  x1 (n) + 1  + a 12 ⋅  x2 (n) + 1  + a 13 ⋅  x3 (n) +  + b1u  T
  2   2   2 
  f   g   h1  
g 2 =  a 21 ⋅  x1 (n) + 1  + a 22 ⋅  x2 (n) + 1  + a 23 ⋅  x3 (n) +  + b 2 u  T
  2  2  2 
  f   g   h1  
h 2 =  a 31 ⋅  x1 (n) + 1  + a 32 ⋅  x2 (n) + 1  + a 33 ⋅  x3 (n) +  + b 3u  T
  2  2  2 
  f   g   h2  
f 3 =  a 11 ⋅  x1 (n) + 2  + a 12 ⋅  x2 (n) + 2  + a 13 ⋅  x3 (n) +  + b1u  T
  2  2  2 
  f   g   h2  
g3 =  a 21 ⋅  x1 (n) + 2  + a 22 ⋅  x2 (n) + 2  + a 23 ⋅  x3 (n) +  + b 2 u  T
  2  2  2 
  f   g   h2  
h 3 =  a 31 ⋅  x1 (n) + 2  + a 32 ⋅  x2 (n) + 2  + a 33 ⋅  x3 (n) +  + b 3u  T
  2   2   2 

f4 = (a 11 (
⋅ ( x1 (n) + f 3 ) + a 12 ⋅ ( x2 (n) + g3 ) + a 13 ⋅ x3 (n) + h 3 + b1u T ) )
g4 = (a 21 ⋅ ( x1 (n) + f 3 ) + a 22 ⋅ ( x2 (n) + g3 ) + a 23 ⋅ ( x ( n) + h ) + b u ) T
3 3 2

h4 = (a 31 ⋅ ( x1 (n) + f 3 ) + a 32 ⋅ ( x2 (n) + g3 ) + a 33 ⋅ ( x ( n) + h ) + b u ) T
3 3 3

x1 (n + 1) = x1 (n) +
(f 1 + 2 f 2 + 2 f3 + f 4 )
6

x2 (n + 1) = x2 (n) +
( g1 + 2 g 2 + 2 g3 + g 4 )
6

x3 (n + 1) = x3 (n) +
(h 1 + 2h 2 + 2h 3 + h 4 )
6

%
8.4.2 Three-State, Fourth-Order Runge–Kutta
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T;
Algorithm and Code g1 = (a21*x1(n) + a22*x2(n) + a23*x3(n) + b2*u)*T;
h1 = (a31*x1(n) + a32*x2(n) + a33*x3(n) + b3*u)*T;
The Runge–Kutta method can be extended to any number
of state equations, by adding an additional set of factors for %
each additional state variable. Below is the algorithm in con- f2 = (a11*(x1(n)+f1/2) + a12*(x2(n)+g1/2) …
ventional mathematical notation for a set of three-state equa- + a13*(x3(n)+h1/2) + b1*u)*T;
tions and the corresponding MATLAB code: Note the use of g2 = (a21*(x1(n)+f1/2) + a22*(x2(n)+g1/2) …
ellipses (the three dots or periods) to continue a MATLAB + a23*(x3(n)+h1/2) + b2*u)*T;
statement to the next line. h2 = (a31*(x1(n)+f1/2) + a32*(x2(n)+g1/2) …
+ a33*(x3(n)+h1/2) + b3*u)*T
%
% Runge-Kutta Difference Equation loop %
% Three State Equations f3 = (a11*(x1(n)+f2/2) + a12*(x2(n)+g2/2) …
% + a13*(x3(n)+h2/2) + b1*u)*T;
for n = 1:N; g3 = (a21*(x1(n)+f2/2) + a22*(x2(n)+g2/2) …
+ a23*(x3(n)+h2/2) + b2*u)*T;
8.5  Programming Non-Linearities and Input Functions 481

20 The system equation relating the input force to the veloc-


Euler
Runge-Kutta ity of the mass is
15
1 d F d 2 v b dv K
10 = 2 + + v
v1g M dt dt M dt M
FK
5 Euler
Runge-Kutta and the system equation relating the input force to the force
0 carried by the spring is
N=2000
-5 K d 2 FK b d FK K
0 20 40 60 80 100 F= 2
+ + FK
M dt M dt M
t, sec
Fig. 8.9   Comparison of the Euler, Runge–Kutta methods and analyti- The results produced by the Runge–Kutta and the Euler
cal calculations. The Runge–Kutta results were indistinguishable from methods, and the analytical solutions for the responses of the
the analytical results whereas the error between the Euler and analytical state variables of a spring-mass-damper system are plotted in
results increased as the calculation progressed
Fig. 8.9. The same time step was used for both Runge–Kutta
and Euler methods. The Runge–Kutta results and the ana-
h3 = (a31*(x1(n)+f2/2) + a32*(x2(n)+g2/2) …
lytical solutions are plotted on top of one another. The Euler
+ a33*(x3(n)+h2/2) + b3*u)*T;
method’s error increases with the number of steps.
%
f4 = (a11*(x1(n)+f3) + a12*(x2(n)+g3) …
+ a13*(x3(n)+h3) + b1*u)*T;
g4 = (a21*(x1(n)+f3) + a22*(x2(n)+g3) … 8.5 Programming Non-Linearities and Input
+ a23*(x3(n)+h3) + b2*u)*T; Functions
h4 = (a31*(x1(n)+f3) + a32*(x2(n)+g3) …
+ a33*(x3(n)+h3) + b3*u)*T; The Runge–Kutta functions in Mathcad and MATLAB can
solve non-linear state equations if the non-linearities can be
% expressed in a single, conventional mathematical expression.
x1(n+1)= x1(n)+(f1+2*f2+2*f3+f4)/6; The term “conventional” in this context means an expression
x2(n+1)= x2(n)+(g1+2*g2+2*g3+g4)/6; without a conditional or “logical” function, such as an if()
x3(n+1)= x3(n)+(h1+2*h2+2*h3+h4)/6; statement. For example, the power-law damping force of
Eq. 8.16 and Fig. 8.10a is a non-linearity which can be de-
% scribed without a conditional statement. However, the ap-
t(n+1) = n*T; proximation of the function created from three straight lines,
end; Fig. 8.10b, requires conditional statements, Eq. 8.20:

Compare the results of the Euler and Runge–Kutta finite  8 N · sec


difference method solutions with the analytical result for  v1g if − 6.4 ≤ v1g ≤ 6.4
 6.4 m
the state-space representation of a spring-mass-damper   27 − 8  N · sec (8.20)
system, where K = 1, M = 3, and b = 0.2. The input is Fb =  8 N +   v1g if 6.4 < v1g
  10 − 6.4 m
F (t ) = 10us (t ).
  27 − 8  N · sec
 −8 N −   v1g if v1g < −6.4
  10 − 6.4  m

Fig. 8.10  a Power law a b 27


damping, Eq. 8.16.,
 3
 3 20 20
Fb =  Aµ0 /h 2  v1g v1g 2.
  8
b Piece-wise continuous Fb Fb
0 0
approximation of the power N N
law damping superposed on the -8
trace of a with inflection points
at v1g = −6.4 m/sec, Fb = −8 N -20 -20
-27
and v1g = 6.4 m/sec, Fb = 8 N
-10 -5 0 5 10 -10 -5 0 5 10
-6.4 m 6.4
m
v1g, ___ v1g, ___
sec sec
482 8  Finite Difference Methods and MATLAB

Fig. 8.11   a Saturation limits the


damper force, Eq. 8.21. b Clip-
a b 6
ping truncates voltage measure- 20 4
ment, Eq. 8.21
2
Fb v(t) 0
0
N VDC
-2
-20 -4
-6
-10 -5 0 5 10 0 1 2 3 4 5
m
v1g, ___ t,sec
sec

Fig. 8.12   a Bingham fluid with


a threshold pressure. b Dead- a b
zone of insufficient angular
20 20
velocity to transmit torque
p12 T 0
0
MPa N •m

-20 -20

-0.01 -0.005 0 0.005 0.01 -2 -1 0 1 2


m3
Q, ___ rad
Ω1g , ___
sec sec

8.5.1  Common Non-Linearities a sinusoidal signal with an amplitude of 5.25 volts acquired
on an oscilloscope set to acquire up to five volts is “clipped”
A non-linearity due to inflections points may be deliberately at five volts, Fig. 8.11b.
created to meet design requirements. For example, when a
force exceeds a certain level, the resulting deflection may 
 N · sec m
bring the component into contact with a second “helper”
3.6 m v1g if v1g ≤ 7 sec
spring. 
Fb =  v (8.21)
A non-linearity due to inflections points can also be the m
 25 N 1g if v1g > 7
result of exceeding the intended operating limits of a device.  v1g sec
For example, rather than a deflecting component contacting
another spring, it contacts the machine frame, changing its
bending mechanism.  5.25 VDC sin(2t ) if v(t ) ≤ 5 VDC

A second type or class of common non-linearities is “sat- v(t ) =  v(t ) (8.22)
uration” or “clipping.” As the term “clipping” connotes, this  5.0 VDC v(t ) if v(t ) > 5 VDC

type of non-linearity is a result of a limit affecting the am-
plitude of a function, Fig. 8.11a. An energetic device is satu- Notice that the clipping described by Eq. 8.22 is a function
rated when additional power input does not yield additional of the output magnitude. The second condition in Eq. 8.22
power output, or additional power output is negligible. For is sensible for a computer program but not for conventional
example, an electronic amplifier is saturated when increas- mathematics. It is an instruction to overwrite a variable’s
ing the input signal does not increase the output signal. A fer- value based on the original value.
romagnetic material is saturated when additional magneto- Backlash, due to clearance between the teeth of meshed
motive force, i.e., additional current through a winding, does gears, slack, due to clearance between components in a force
not increase the torque of the motor, or, more precisely, does load path or the lack of tension in belts, chains and cables,
not increase the magnetic flux producing the torque above and slip, due to insufficient shear force within a mechanism,
that, which would exist without the ferromagnetic material. result in temporary lack of power transmission on startup or
Clipping has the same appearance as saturation but is due loss of power transmission when a system reverses direc-
to a limit of the acquisition or processing of data rather than a tion. Backlash and slack are manifested as “dead-zones” of
power limitation in the physical phenomenon. For example, zero force or torque on either side of zero displacement when
8.5  Programming Non-Linearities and Input Functions 483

Fig. 8.13   a Check valve with a


threshold pressure below which
a b
it does not open. b Self-locking 20 20
mechanism, which will only
rotate in the positive direction
p12 T 0
0
MPa N •m

-20 -20

-0.01 -0.005 0 0.005 0.01 -2 -1 0 1 2


m3
Q, ___ rad
Ω1g , ___
sec sec

the velocity has been zero. This requires integration of the 8.5.2 Input Functions, Non-Linearities and the
mechanism velocity to yield displacement as part of the fi- Runge–Kutta Algorithm
nite difference code.
Thresholds are non-linearities, which manifest the op- The MATLAB code for non-linearities and input functions
posite behavior of a dead-zone, Fig. 8.12a. A threshold is a must be inside the ­for-end loop which contains the Runge–
level below which a finite force produces zero velocity. The Kutta algorithm, Fig. 8.14. Input functions are straightfor-
force must exceed the threshold level for the motion to occur. ward. Programming pulses into the Runge–Kutta algorithm
Thresholds are found in all types of physical systems, either is easier than constructing a response function using super-
by design or coincidence. Many electronic devices have volt- position. The Runge–Kutta algorithm only needs the n − 1
age thresholds below which they will not conduct. A Bing- values of the state variables, i.e., x1( n − 1), x2( n − 1), the val-
ham fluid is a fluid with a threshold due to shear strength. ues immediately preceding the step, and the nth value input.
It will not flow below its threshold pressure. E.C. Bingham The values of the state variables are saved as a matter of
(1879–1945) was a professor of chemistry at Lafayette Col- course since the vectors of the values of the state variables is
lege who pioneered the study of rheology. the solution of the algorithm. All that must be programmed
Directional dependency is often designed into compo- is the value of the input.
nents for functional requirements of the system, e.g., check Pulses are created by conditional statements. A single
valves to prevent backflow and diodes limit current flow pulse is created by an if statement before the Runge–Kutta
to one direction, Fig. 8.13. Diodes and many check valves algorithm. For example, the following MATLAB code cre-
also have a threshold voltage or pressure below which there ates a pulse of amplitude 10 and a duration of 0.5 sec.
is no flow. Mechanical systems can be “locking” or “self-
locking.” In some cases such as ratchets, the locking mech- %
anism permits motion only in one direction until the ratchet for n = 1:N;
is released. In other cases, such as a worm and worm gear, %
the system can be driven in either direction by an input % Use if an if() to calculate the input u. Set u = 0 then
torque on the worm but cannot be driven by output torque % overwrite with 10 if t < 0.5
on the gear. In addition to fluid non-linearities, the cross- u = 0;
sectional area of a hydraulic piston/cylinder’s operating if(n*T<=0.5);
rod reduces the piston area on that side. Consequently, the u = 10;
maximum tensile force a double acting (push-pull) piston/ end
% Beginning of Runge-Kutta Algorithm
cylinder can produce is less than the maximum compres-
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T;
sive force.
g1 = (a21*x1(n) + a22*x2(n) + a23*x3(n) + b2*u)*T;
Directional dependency can also occur in translational
systems when either the operating range or the output ex-
ceeds the expected limits and components designed to be in A more complicated pulse requires more conditional state-
tension are placed in compression or components intended ments and the form of an if-elseif statement.
to be in compress lose contact with the opposing surface and
exert no force. Directional dependency also occurs due to %
asymmetric wear or memory of a material. for n = 1:N;
%
% Use if an if-elseif to calculate the input u. Set u = 0
484 8  Finite Difference Methods and MATLAB

Fig. 8.14   Runge–Kutta algo- Initialize variables


rithm within a for-end loop. The
input value and state variable
dependent parameter values Calculate value of the input u.
are calculated at the top of the
for n=1:2000
Determine state variable dependent
loop, before the Runge–Kutta parameter values.
algorithm. At the bottom of the f1=(a11*x1(n)+a12*x2(n)+a13*x3(n)+b1*u)*T;
loop, after the state variables are g1=(a21*x1(n)+a22*x2(n)+a23*x3(n)+b2*u)*T;
h1=(a31*x1(n)+a32*x2(n)+a33*x3(n)+b3*u)*T;
updated, velocities are inte-
grated to displacements and state f2=(a11*(x1(n)+f1/2)+a12*(x2(n)+g1/2)+a13*(x3(n)+h1/2)+b1*u)*T;
variables are overwritten if they g2=(a21*(x1(n)+f1/2)+a22*(x2(n)+g1/2)+a23*(x3(n)+h1/2)+b2*u)*T;
h2=(a31*(x1(n)+f1/2)+a32*(x2(n)+g1/2)+a33*(x3(n)+h1/2)+b3*u)*T;
exceed limits
Runge f3=(a11*(x1(n)+f2/2)+a12*(x2(n)+g2/2)+a13*(x3(n)+h2/2)+b1*u)*T;
g3=(a21*(x1(n)+f2/2)+a22*(x2(n)+g2/2)+a23*(x3(n)+h2/2)+b2*u)*T;
Kutta h3=(a31*(x1(n)+f2/2)+a32*(x2(n)+g2/2)+a33*(x3(n)+h2/2)+b3*u)*T;
Algorithm
f4=(a11*(x1(n)+f3)+a12*(x2(n)+g3)+a13*(x3(n)+h3)+b1*u)*T;
g4=(a21*(x1(n)+f3)+a22*(x2(n)+g3)+a23*(x3(n)+h3)+b2*u)*T;
h4=(a31*(x1(n)+f3)+a32*(x2(n)+g3)+a33*(x3(n)+h3)+b3*u)*T;

f=(1/6)*(f1+2*f2+2*f3+f4);
g=(1/6)*(g1+2*g2+2*g3+g4);
h=(1/6)*(h1+2*h2+2*h3+h4);

x1(n+1)=f+x1(n);
x2(n+1)=g+x2(n);
x3(n+1)=h+x3(n);
t(n+1) = n*T; Integrate velocities to displacements.
Test state variables with conditional
end; statements and overwrite them if
they exceed a limit.
Plot state variables

% then overwrite with u = 10 if t <= 0.5


% and u = -10 if 0.5 < t <= 1 v1g(n+1) + v1g(n)
∆x1g = _____________ T
u = 0; 2
if(n*T <= 0.5);
u = 10;
v1g (n)
elseif(n*T <= 1.0);
u = -10 m
___
end sec
% Beginning of Runge-Kutta Algorithm
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T; n-2 T n-1 T n T n+1
g1 = (a21*x1(n) + a22*x2(n) + a23*x3(n) + b2*u)*T; Step Number
Fig. 8.15   Trapezoidal integration of velocity v1g to displacement x1g.
The n + 1 value of v1g was just calculated by the Runge–Kutta algorithm
Non-linear parameter functions that use the current value of and will be the n value when the for-end loop is iterated
the state variable to determine the value of a parameter are
also placed at the top of the for-end loop before the Runge–
Kutta algorithm. The parameter value is calculated first and Trapezoidal integration of state variables, which are ve-
then the coefficients containing that parameter are calculat- locities, occurs at the bottom of the for-end loop, after the
ed. In a linear Runge–Kutta calculation with constant param- Runge–Kutta algorithm and the update of the state variables
eters, the coefficients are calculated before the loop, since from their x( n) values to their x( n + 1) values.
they need only be calculated once.

Summary
8.5.3  Trapezoidal Integration
Finite difference methods are necessary to solve non-linear
Integrating velocity to displacement is needed to impose system equations. State equations are solved using finite
non-linearities created by geometric constraints. Trapezoidal difference methods in all cases. The state-space represen-
integration is the simplest, accurate method of numerical in- tation is particularly convenient for non-linear dynamic
tegration. As the name describes, the element added to the systems. The Euler method was the first method of finite
integral is trapezoidal, Fig. 8.15. differences and remains the simplest. The Runge–Kutta
Problems 485

Flywheel iv The angular velocity difference across the fluid


Rotational Inertia J coupling b.
8.1.c Express the state and output equations in matrix form.
Compliant Shaft 8.1.d For Cases I and II, calculate the system’s eigenvalues.
Torsional Spring K
If the system is underdamped, determine the ideal,
Fluid Coupling undamped natural frequency ωn, the damped natural
Damping b frequency ωd, and the damping ratio ζ.
8.1.e The system is de-energized when it is acted on by a
Angular Velocity step input of 10 rad/sec. Write a MATLAB program
Input Ω(t) for the Euler method to solve the state equations, and,
using the solution of the state equations, solve the out-
Fig. P8.1   A rotational mechanical system put equations for Cases I and II. Plot the responses of
the output variables.
8.1.f Repeat part e using the Runge–Kutta algorithm.
method is the most widely used because of its accuracy and 8.1.g Plot the Euler method and Runge–Kutta algorithm
numerical stability. results for the angular velocity of flywheel on the same
plot. How does the difference between the results of
the two methods vary over the solution?
Problems
Problem 8.2 A mass M supported on frictionless roller is
Problem 8.1 The rotational system shown schematically in shown in Fig. P8.2a. A dashpot with damping b and a spring
Fig. P8.1 comprises an angular velocity input driving a fluid with spring constant K are connected between the mass and
coupling attached to a compliant shaft modeled as a spring. ground. The mass is acted upon by an applied force F( t).
The compliant shaft drives a flywheel with mass moment The system has a linear and a non-linear configuration.
of inertia J. Two sets of parameters, Case I and Case II, are The linear configuration has linear dampers, which are de-
tabulated. scribed by the equation Fb = bv1g. The non-linear configura-
Table P8.1   Parameter values
tion has non-linear dampers designed so that they exert a
damping force when the damper is in compression but not
J K b
when the damper is in extension, as shown in Fig. P8.2b.
Case I 3 2 1
Table P8.2   Parameter values
Case II 1 2 3
M K b
Case I 3 2 1

8.1.a Derive the state equations for this system. Case II 1 2 3


8.1.b Derive the output equations for:
i The angular velocity of the inertia J. 8.2.a Derive the state equations for this system.
ii The torque acting through the compliant shaft 8.2.b Derive the output equations for:
spring K. i The velocity of the mass M.
iii The angular velocity difference across the compli- ii The force acting through spring K.
ant shaft spring K. iii The angular velocity difference across spring K.
8.2.c Express the state and output equations in matrix form.

Fig. P8.2  a Second-order


translational mechanical system. a x,v b
b Non-linear damper, which b 20
exerts a damping force only in
compression Fb
F(t)
M N 0

K -20

-10 -5 0 5 10
m
v1g, ___
Frictionless rollers sec
486 8  Finite Difference Methods and MATLAB

R1 is a check valve.
Permits flow only in R1 L
this direction. C
R1 I D
P(t)
vent to Pump
atmosphere R2 v(t) C
Fluid
R2
Reservoir

Fig. P8.3a   Fluid circuit schematic


Fig. P8.4a   Schematic of a RLC circuit with a diode D

2,000 8.3.b Derive the output equations for:


p(t) i The volume flow rate from the pressure source.
psi ii The pressure in the fluid accumulator.
1,000 iii The volume flow rate through the fluid inertance.
iv The pressure difference across the fluid inertance.
and check their units.
0
0 10 20 30 8.3.c Express the state and output equations in matrix form.
t, sec 8.3.d Calculate the system’s eigenvalues configured with
the linear fluid resistance R1. If the system is under-
Fig. P8.3b  Input pressure pulse
damped, determine the ideal, undamped natural fre-
quency ωn, the damped natural frequency ωd, and the
8.2.d For the linear Cases I and II, calculate the system’s damping ratio ζ.
eigenvalues. If the system is underdamped, determine The system was de-energized before the pressure pulse plot-
the ideal, undamped natural frequency ωn, the damped ted in Fig. P8.3b was applied to the system.
natural frequency ωd, and the damping ratio ζ. 8.3.e Write a MATLAB program to solve the linear state
8.2.e Write a MATLAB program to solve the linear state equations using the Runge–Kutta algorithm for the
equations using the Runge–Kutta algorithm for a step pressure input shown in Fig. P8.3b. Assume that the
input of 10 N. Assume that the system was de-ener- system was de-energized at time t = 0. Include the code
gized at time t = 0. Include the code to solve the out- to solve the output equations and plot their responses.
put equations and plot their responses. Run your code 8.3.f Save your MATLAB script of part e under a different
using the parameter values for Cases I and II. file name and edit it to solve the non-linear state equa-
8.2.f Save your MATLAB script to a different file name tions using the Runge–Kutta algorithm for the pres-
and edit the script to include the non-linear behavior sure input shown in Fig. P8.3b.
of the dashpot. Run your program for the same input 8.3.g Compare the pressure across the fluid inertance in the
and parameter values as part e. linear and non-linear cases and calculate the maxi-
mum change in pressure due to the check valve.
Problem 8.3 The fluid system shown schematically in
Fig. 8.3a consists of a fluid power unit, modeled as a pres- Problem 8.4 The electric circuit shown in the schematic,
sure source, two fluid resistances R1 = 100  MPa · sec/m3 Fig.  8.4a, consists of a voltage source, two resistors,
and  R2 = 200  MPa · sec/m3, a long run of piping, modeled R1 = 0.1W and R2 = 1W, a capacitor, C = 20 × 10−6 F, an induc-
as a fluid inertance, I = 40 × 107  kg/m4 and a fluid accumu- tor, L =1 × 10−6 H, and a diode D.
lator (capacitance), C = 6 × 10−8 m3/Pa. The system will be Diodes conduct current only in the direction that the
modeled twice, a linear and a non-linear model. In the linear triangle of their symbol points Fig. 8.4b. This is a diode’s
model, the fluid resistance R1 has a constant resistance and “forward” direction. Diodes have a threshold voltage, which
allows flow in both ­directions. In the non-linear model, the varies with the material and design of the diode. A typical
fluid resistance R1 is a check valve which only allows flow silicon diode has threshold voltage of 0.63 VDC before it
in the direction shown. The fluid resistance in the forward conducts in the forward direction, Fig. 8.4c. An actual diode’s
direction is the same as the linear case. voltage–current plot is similar to Fig. 8.4c. The small in-
8.3.a Derive the state equations for this system. crease in resistance to the forward conduction with increased
Problems 487

Fig. P8.4b   Diode schematic


symbol showing forward direc-
b Anode = Positive c 30
tion (direction of current flow). The band on physical +
c Diode voltage–current plot + diode and the bar on
20
the schematic symbol
iD indicate the cathode. iD
A 10

- 0
-
Cathode = Negative -0.5 0 0.5 1.0 1.5
vD , VDC

10 tions for the non-linear configuration, which includes


the diode, for the voltage input shown in Fig. P8.4d.
5
8.4.g C
 ompare the voltage across the inductance in the lin-
v(t) ear and non-linear systems and calculate the maxi-
0
VDC 0.05 0.1 0.15 2 mum change in voltage due to the diode.
t, msec
-5
Problem 8.5 Two rotational mechanical systems are shown
-10 schematically in Figs. P8.5a and P8.5b. Both systems contain
a torque source, a gear with rotational inertia, J1, supported
Fig. P8.4d   Input voltage by bearings with damping b1, a shaft with torsional stiffness,
K, and a flywheel with rotational inertia, J2, supported by
bearings with damping, b2. In System I, the gear is a spur
current is usually neglected and the diode modeled as having gear, driven by a pinion with negligible inertia. In System II,
the threshold voltage drop but no additional resistance. the gear is a worm gear, driven by a worm with negligible
The system will be modeled in two configurations, a lin- inertia. The gear ratios of Systems I and II are the same: 20
ear configuration without the diode, and a non-linear con- rotations of the pinion, or worm, to one rotation of the gear.
figuration with the diode.
Table P8.5   Parameter values for systems I and II
8.4.a Derive the state equations for the linear configuration
J1 J2 K b1 b2
and check their units.
8.4.b Derive output equations for the following variables of To be To be N · m · sec N · m · sec
0.5kg · m 2 10 20
the linear system. calculated calculated rad rad
i The voltage across capacitor C.
ii The current flowing through inductor L.
iii The voltage across inductor L. There is one significant difference between the two sys-
iv The current through resistor R1. tems. The spur gears of System I can be driven by either
v The current through resistor R2. the torque source or the output shaft, K. However, the worm
8.4.c Write the state equations and output equations in vec- gears of System II can only be driven by the torque source. A
tor-matrix form. worm drive is “self-locking.” It cannot be driven by its out-
8.4.d Calculate the system’s eigenvalues If the system is put shaft on the worm gear, only the input shaft on the worm.
underdamped, determine the ideal, undamped natural The angular velocity of the pinion in System I need not
frequency ωn, the damped natural frequency ωd, and have the same sign as the torque applied by the torque source.
the damping ratio ζ. However, the angular velocity cannot have the opposite sign
The system is initially de-energized before the voltage pulse of the torque from the torque source. The angular velocity of
shown in Fig. P8.4d was applied. the worm in System II can be of the same sign as the torque
8.4.e Write a MATLAB program to solve the linear state applied by the torque source or zero.
equations using the Runge–Kutta algorithm for the The spur gears are counter-rotating. Consequently, the an-
voltage pulse input shown in Fig. P8.4d. Assume that gular velocities of the two shafts have opposite signs. We are
the system was de-energized at time t = 0. Include free to define the signs for the rotation of the input and output
the code to solve the output equations and plot their shafts of the worm drive. For convenience, we will define the
responses. sense of rotation of the input and output shafts to have oppo-
8.4.f Save your MATLAB script of part e under a different site signs to correspond to the rotation of the spur gear system.
file name and edit the script to solve the state equa- 8.5.a Derive the state equations for Systems I and II.
488 8  Finite Difference Methods and MATLAB

Fig. P8.5a   System I spur gear Bearings,Negligible


drive damping Pinion, Negligible
Inertia
T(t)
Torque Load
Source Inertia J 2

Torsional Shaft
Spring Constant, K

Bearings Gear Bearings


Damping b1 Inertia J1 Damping b2

Fig. P8.5b   System II worm Bearings


drive Negligible
damping Worm Gear Load
Inertia J 1 Inertia J2
Worm
Negligible
Inertia Torsional Shaft
Spring Constant, K

Bearings Bearings
Damping b1 Damping b 2

T(t) Torque
Source

A
1 in

10
T(t)
N •m
4 in 2 in

0 1 2
32 in 36 in t, sec
Fig. P8.5d   Input torque pulse

8.5.c The flywheel, J2, shown in the plan and cross-section


in Fig. P8.5.c, is carbon steel. Calculate the mass
A Section A-A moment of inertia (rotational inertia), J2, using super-
position, as described in Chap. 4. The spokes are 1 in.
Fig. P8.5c   Flywheel J2 wide.
8.5.d Draw a flowchart of the logic needed in the state equa-
8.5.b The shaft, K, is carbon steel, solid with a circular tion solver loop, to prevent “backwards” (positive)
cross-section, 2 in. in diameter and 12 ft long. Cal- rotation of the worm gear, and to lock the worm gear,
culate its torsional spring constant, K, as described in when the input torque is turned off.
Chap. 4.
Problems 489

x,v
10
K F(t)
b2
F(t) N
M1 b1 M2
0 20 40
t, sec
Fig. P8.6c   Input force pulse
Fig. P8.6a   Third-order mechanical system

Table P8.6   Parameter values


20 M1 M2 K b1 b2
Case I 1.0 2.0 1.0 0.1 2.0
Case II 0.2 0.4 1.0 0.1 0.2
Fb2
0
N
8.6.a Derive the state equations for this system.
8.6.b Derive the output equations
-20
-10 -5 0 5 10
i The velocity of mass M1.
m
v2g, ___ ii The force acting through the spring K.
sec iii The velocity of mass M2.
Fig. P8.6b   Non-linear damper force, damper b2 Case I. The damping iv The force acting through damper b1.
coefficient in compression is that in Table P8.6. The damping coeffi- v The force acting through damper b2.
cient in extension is one half of the tabulated value 8.6.c Express the state and output equations in matrix form.
8.6.d For the linear configuration, calculate the system’s
eigenvalues for Cases I and II. If the system is under-
The system is at rest and relaxed for t < 0, before an input damped, determine the ideal, undamped natural fre-
torque pulse with magnitude of 10 N · m and a duration of quency ωn, the damped natural frequency ωd, and the
one second is applied at time, t = 0, as shown in Fig. P8.5d. damping ratio ζ.
8.5.e Write a MATLAB code to solve the state equations 8.6.e For Cases I and II of the linear configuration, write a
of System I with the Runge–Kutta algorithm for the MATLAB program to solve the state equations and
pulse input shown in Fig. P8.3d. The duration of the the output equations for the force pulse plotted in
simulation should be 2 sec. Plot the state variables. Fig. P8.6c. Plot the responses of the output variables.
8.5.f Modify the code to solve the state equations of System 8.6.f Edit a copy of the MATLAB program of part e to solve
II, again plotting the state variables. state and output equations of the non-linear configura-
8.5.g Calculate the maximum stress in the shafts of System tion for Cases I and II. Plot the responses of the output
I and System II. If the yield strength of the steel is variable to the force pulse of Fig. P8.6c.
60 ksi, did either shaft fail?
Problem 8.7 A translational mechanical system contains a
Problem 8.6 A translational mechanical system contains mass M sliding on a lubricating fluid film with damping b.
a spring, two masses, and two dampers acted upon by an The mass is acted upon by an applied force F( t). Spring K1 is
applied force is shown in Fig. P8.6a. The system has lin- attached between the mass and ground. Spring K2 is attached
ear and non-linear configurations. There are two cases for to ground but is not attached to the mass. When the system is
both configurations. The two cases for the linear configu- at rest and relaxed, there is a gap of width d between Spring
ration of the system use the parameters in Table P8.6. The K2 and the mass M. Spring K2 and is only in contact with the
two cases of the non-linear configuration use the mass mass when the position of the mass compresses spring K2.
and spring constant values from Table P8.6, but have non- Two linear and one non-linear configurations of the sys-
linear dampers which have half the damping coefficient tem will be investigated. The two linear configurations are
b in extension as they do in compression, as shown in (1) Spring K2 is absent from the system, and (2) Spring K2
Fig.  P8.6b. The values of damping coefficient b1 and b2 is attached to mass M. The non-linear configuration is that
in Table P8.6 are the higher compression values for the shown in Fig. P8.7 with gap d between the mass and spring
non-linear dampers. K2 when the system is at rest and relaxed. We will investigate
two cases. (1) The gap d of the de-energized system equals
490 8  Finite Difference Methods and MATLAB

x,v d 8.7.g Save your MATLAB script of part f under a different


name and edit it to solve the non-linear state equations
K2 which include a gap between the mass and spring K2.
F(t) Solve the state and output equations for the input and
M K1 conditions of part f and the two different gap widths.

Problem 8.8 The schematic of an electric circuit with a volt-


age source, two inductors, two capacitors, a resistor, and a
Lubricating fluid diode is shown in Fig. P8.8a.
Damping b
Table 8.8a   Parameter values
Fig. P8.7   Second-order translational mechanical system
C1 C2 L1 L2 R
1 µF 2 µF 3 µH 4 µH 2W

L1 L2
Diodes conduct current only in the direction that the tri-
angle of their symbol points Fig. 8.8b. This is a diode’s “for-
D
ward” direction. Diodes have a threshold voltage which var-
ies with the material and design of the diode. A typical silicon
v(t) C1 C2 diode has threshold voltage of 0.63 VDC before it conducts
R in the forward direction, Fig. 8.8c. An actual diode’s volt-
age–current plot is similar to Fig. 8.8c. The small increase
in resistance to the forward conduction with increased cur-
Fig. P8.8a   A non-linear RLC circuit
rent is usually neglected and the diode modeled as having the
threshold voltage drop but no additional resistance.
The system will be modeled in two configurations, a lin-
zero, i.e. the spring is in contact with but not attached to the ear configuration without the diode, and a non-linear con-
mass, and (2) the gap equals half the steady-state displace- figuration with the diode.
ment of the mass under the load of 1,000 N. There are two 8.8.a Derive the state equations for the linear system and
sets of parameters, Case I and Case II. check their units.
8.8.b Derive output equations for the following variables of
Table P8.7   Parameter values
the linear system.
M K1 K2 b
i The voltage across capacitor C1.
Case I 50 10,000 20,000 60
ii The voltage across capacitor C2.
Case II 50 20,000 10,000 60 iii The current through inductor L1.
iv The current through inductor L2.
v The voltage across inductor L1.
8.7.a Derive the state equations for this system using Kequiv vi The voltage across inductor L2.
as the spring constant. vii The current through resistor R.
8.7.b Derive the output equations for: 8.8.c Write the state equations and output equations in vec-
i The velocity of the mass M. tor-matrix form.
ii The force acting through spring K. 8.8.d Calculate the linear system’s eigenvalues. If the sys-
8.7.c Express the state and output equations in matrix form. tem is underdamped, determine the ideal, undamped
8.7.d For the linear configuration, calculate the system’s natural frequency ωn, the damped natural frequency
eigenvalues for the parameter values of Cases I and ωd, and the damping ratio ζ.
II. If the system is underdamped, determine the ideal, The system is initially de-energized before the voltage
undamped natural frequency ωn, the damped natural pulse shown in Fig. P8.8d was applied.
frequency ωd, and the damping ratio ζ. 8.8.e Write a MATLAB program to solve the linear state
8.7.f Write a MATLAB program to solve the linear state equations using the Runge–Kutta algorithm for the
equations using the Runge–Kutta algorithm for a step voltage pulse input shown in Fig. P8.8d. Include
input force F( t) = 1,000  N us( t). Assume the system is the code to solve the output equations and plot their
at rest and relaxed before the input acts on the system. responses.
Solve the output equations for the parameter values of
Cases I and II and plot their responses.
Problems 491

Fig. P8.8b   Diode schematic


symbol showing forward direc-
b Anode = Positive c 30
tion (direction of current flow). The band on physical +
c Diode voltage–current plot + diode and the bar on
20
the schematic symbol
iD indicate the cathode. iD
A 10

- 0
-
Cathode = Negative -0.5 0 0.5 1.0 1.5
vD , VDC

10
L y,v
5
x M
v(t)
0 h
VDC 1 2 3 4 5
t, millisec
-5 d

-10

Fig. P8.8d   Input force pulse


Fig. P8.9b   Support positioned under the beam at distance x from the
wall and a gap d from the bottom of the beam when it is unloaded and
at rest
Young’s modulus E
Density ρ
Width b 8.9.b Determine the damping coefficient b needed for the
system to have a damping ratio of ζ = 0.04, using the
M velocity the point on the beam at distance L from the
wall as the location of the velocity node 1. Determine
h
the eigenvalues of the system.
L 8.9.c The design is modified adding a support such that
there is a gap d = 1 mm below the bottom surface of
the beam, when the system is at rest and relaxed, at the
distance of x = 30 mm from the support, Fig. P8.9b.
Fig. P8.9a   Linear cantilevered beam with a cylindrical mass attached When the beam is in contact with the support during
at distance L from the support downward deflection, approximate it as a cantilevered
beam 20 mm long. The spring constant of the beam
8.8.f Save your MATLAB script of part e under a different is greatly increased. Conversely, the effective mass of
file name and edit it to solve the non-linear state equa- the beam is reduced when beam is in contact with the
tions, which include the diode in the circuit, for the support.
voltage input shown in Fig. P8.8d. The deflection equation for a point at distance x from the
8.8.g Compare the voltage across the inductances in the support of a cantilevered beam of length L is
linear and non-linear systems and calculate the maxi-
mum change in voltage due to the diode. Fx 2
δx = (3L − x )
6 EI
Problem 8.9 A cantilevered beam with rectangular cross-
section is shown in Fig. P8.9a. Attached to the beam is a The area moment of inertia of a rectangular cross-section
cylinder of mass M. The diameter of the cylinder equals the about a horizontal line through its centroid is
width of the beam w. The height of the cylinder is twice its
diameter. wh3
I=
8.9.a Determine the undamped angular frequency of the 12
beam and mass system if they are carbon steel and
L = 50  mm, w = 5  mm, and h = 1  mm.
492 8  Finite Difference Methods and MATLAB

Fig. P8.10a  Electromechanical Voltage Source v(t) Electrical Resistance R


system with a compliant shaft
belt slip Compliant Shaft
Torsion Spring K1

Flywheel 1
Rotational Inertia J 1 DC Electric Motor
Bearings Damping b1 Torque constant K T

Flywheel 2
Rotational Inertia J 2
6 in. Diameter Pulley Bearings Damping b 2

Belt

4 in. Diameter Pulley

Fig. P8.10b  The compliant vsurface vsurface


belt’s taut and slack sides switch
positions during an oscillation. Slack Taut
Ω 2g Ω1g Ω 2g
Ω1g Taut Slack
Driving vsurface Driven Driven
Pulley Driving vsurface
Pulley Pulley Pulley

K2 positioned under the beam. Run your program for the


n-1 n n+1 n+2 same step input of F0 = 10 N.

Problem 8.10 An electromechanical system is shown in


Fβ Fig. P8.10a. A voltage source drives a DC motor, which has
Tα Fα Tβ
resistance R and torque constant KT. The DC motor turns
a compliant shaft, modeled as a torsion spring with spring
constant K. The shaft turns flywheel1 with mass moment of
g inertia J1. Flywheel1 is supported on bearings with damping
b1. Flywheel2 has mass moment of inertia J2 bearings with
Fig. P8.10c  Linear graph representing a compliant belt as a transla-
tional spring between two rotational to translational transducer inter- damping b2, and is belt driven. A belt runs between flywheels
faces. one and two. The belt slips on the 4-in. diameter pulley when
the force carried by the belt equals 20 N, limiting the maxi-
Determine the effective mass and stiffness of the beam with mum force in the belt to 20 N.
the support as a function of the vertical displacement y of
Table P8.10   Parameter values
node 1, at the distance L from the wall.
R KT K1 J1 b1
8.9.d Derive the state equations for this system.
8.9.e Derive the output equations for: 3 Ω N·m N·m 6  kg  · m2  1 N · m · sec
9 800
i The velocity of node 1. A rad rad
ii The force acting through the equivalent spring K. J2 b2 K2
iii The angular velocity difference across spring K. N·m
5  kg · m2  N · m · sec 8,000
8.9.f Express the state and output equations in matrix form. 1
rad
rad
8.9.g Write a MATLAB program to solve the linear state
equations using the Runge–Kutta algorithm for a step
input of F0 = 10 N. Assume that the system was de-
energized at time t = 0. Solve the output equations and The system model without belt slip is the linear model.
plot the responses of the output variables. The system model with belt slip is the non-linear model.
8.9.h Save your MATLAB script with a different file name 8.10.a Derive the state equations of the linear model and
and edit the script to include the effect of the support check their units.
Chapter 8 Appendix 493

8.10.b Derive output equations for the following variables: Now that microprocessors capable of machine control cost
i The current drawn from the voltage source. less than one dollar, only the simplest machines are purely
ii The torque acting through the compliant shaft K1. mechanical. A machine’s functionality can be enhanced, and
iii The angular velocity of flywheel J1. its performance improved, by microprocessor control. A ma-
iv The angular velocity flywheel J2. chine designer must understand the basics of how micropro-
8.10.c Write the state equations and output equations in cessors function, to understand the machine they are part of.
vector-matrix form.
8.10.d Calculate the system’s eigenvalues for the linear
case with no belt slip. If the system is underdamped, A Brief History and Classification of Computer
determine the ideal, undamped natural frequency ωn, Programming Languages
the damped natural frequency ωd, and the damping
ratio ζ. Computer programming languages have evolved to meet the
The system was de-energized before a step input of 48 VDC changing demands of users. The very first computers were
was applied. programmed in a “machine-level” language, which specified
8.10.e Write MATLAB code to solve the state equations and how to implement each individual step required to manipu-
output equations of the linear model with the Runge– late the data. Machine-level “code,” or “instructions,” are at
Kutta algorithm for the step input of 48 VDC. Plot the level of detail needed to describe a complicated hand-
the responses of the output variables. held calculator computation, i.e., recall x, recall y, add x and
8.10.f Save the MATLAB script of part e under a differ- y, store the result in z, etc. The instructions, however, are in
ent name and edit the code to solve the state equa- binary code, since the computer’s digital circuitry functions
tions and output equations of the non-linear model with 0’s and 1’s. The computer machine code equivalent to
for the step input of 48 VDC. Plot the responses of a calculator’s “recall” is “move data,” which is the opera-
the output. tion of reading data at a location in memory, and writing it
to a different location. Typical machine code for the opera-
tion, “move data,” is the four-digit binary code, 1011. Binary
Chapter 8 Appendix code is difficult for humans to work with. The codes, 1101,
1011, and 1110, are different instructions. It was not long
Introduction to Programming and MATLAB before the second-generation computer languages, called
“assembly languages,” were created. Assembly languages
Why study computer programming, if engineering software use an alphanumeric code in place of binary code, making
is available? Further, the sophisticated computer programs them much easier for humans to understand. The alphanu-
used in engineering design and analysis are enormous. They meric code incorporates three-letter abbreviations for the
often required teams of programmers and years to develop. machine’s operations. For example, MOV is the assembly
What can an individual engineer and novice computer pro- code for “move data” and MUL is the assembly code for
grammer accomplish of value? Fortunately, one does not “multiply.” The most important memory locations, known as
need to be an expert programmer to write useful computer registers, are given alphanumeric codes, such as AX and B1.
programs. A little knowledge of computer programming Octal (base 8) or hexadecimal (base 16) can be used in place
goes a long way. of their 8, 16, or 32 binary-digit addresses.
Competent engineers must be able to write simple com- Assembly language code is still commonly used to pro-
puter programs for data acquisition and reduction, as well gram the small microprocessors used for machine control. It
as design computations. It is helpful to know some common was superseded for general use in the early 1950s, however,
programming syntax and techniques, since they appear in when FORTRAN (an acronym for FORmula TRANslation),
many user interfaces. An understanding of the fundamentals the first of the third-generation computer languages, was
of computer programming is important to maximize the util- introduced. FORTRAN was the dominant program for en-
ity of most engineering software. Most engineering software gineering computations through the mid-1970s. FORTRAN
packages allow the user to write either “macros” or “scripts,” is now known as a “procedural” language. The syntax and
which are short programs, to automate repetitive processes structure of a procedural computer language are oriented
or iterative computations. towards algorithm, i.e., the logic of the computation. Early
Computers are not only an important tool in the design procedural languages include ALGOL (ALGOrithmic Lan-
process. Microprocessors are now widely used as “embed- guage, mid-1950s), COBOL (COmmon Business Operations
ded” controllers in machines. Microprocessors are micro- Language, late 1950s), LISP (LISt Processing language,
computers which contain the central processing unit (or early 1960s), and BASIC (Beginner’s All-purpose Symbolic
CPU which performs arithmetic and logical operations on Instruction Code, mid-1960s). BASIC was implemented as
data), memory, and input and output devices on a single chip. an “interpreted,” rather than a “compiled,” language. An
494 8  Finite Difference Methods and MATLAB

interpreted language is converted into machine code line variables. The addition operator for real variables is part of
by line, as the program executes. Compiled languages are the programming language. The algorithms to create an ad-
converted into machine code, before the program executes. dition operator for complex variables would be part of the
Compiled programs run significantly faster than interpreted class, “Complex.” The class, Complex, would include other
programs. Procedural languages remain sufficient for pro- properties, such as calculation of the magnitude, and opera-
gramming engineering computations. FORTRAN was last tions, such as multiplication and division of complex num-
revised in 2008. LISP is used in Autocad. The current mani- bers. The class, Complex, would be added to any program
festation of BASIC is Microsoft’s Visual Basic 2013. which includes complex variables. Each complex variable
Two diametrically opposed trends impacted the devel- used in the program would be an “object.”
opment of computer languages in the late 1960s. Computer The first widely adopted object-oriented programming
programs increased in size, while computers themselves language was C++ (early 1980s). C++ was developed from C.
decreased in size. Computer scientists found that some as- C programs can run in most implementations of C++. Today,
pects of early procedural languages became unmanage- C++ seems to be the dominant language used in academic
able, as programs grew in size and were written by teams research. Microsoft gives away stripped-down versions of
of programmers. Pascal (mid-1970s) was the first of the new C++, and its modernized version, called C#, with no multi-
“structured” programming languages. Structured program- core support or 64-bit math. C# has simpler syntax than C++.
ming languages modularized a large programming task into a Unfortunately, object-oriented languages are more difficult to
set of smaller tasks, and provided a structure for the modules learn than procedural languages. There is twice as much to
to interact. Modularizing the program code also allowed for understand. Object-oriented languages have all of the syntax
easier reuse of portions of the code in other programs. Si- and operators for their procedural aspect, plus as much again
multaneously, “minicomputers” using newly developed inte- in syntax and operators for their object-oriented aspect.
grated electronics provided low-cost competition to large and We will work with the programming language, MAT-
expensive “mainframes.” These smaller computers prompted LAB. MATLAB is a “high-level” programming language
a minimalist approach to developing applications to run on developed for engineering and scientific applications. “High
them. The language C (early 1970s) is a procedural language level” means that many of the tasks, which must be pro-
with an enriched set of operators. C also provides access to grammed in other languages, are performed automatically
bit-level operations, formerly only possible using assembly in MATLAB. Specifically, all of the mathematical functions
language. C remains an important language today. Many mi- you can imagine, along with an engineer-friendly plotting
croprocessors programmable in assembly language are also function, are provided. MATLAB is an interpreted not com-
programmable in C. Although the richness of C permits ar- piled language. It runs slowly, compared to C or C + + . Fur-
cane programming practices, it is also possible to program in ther, an academic license limits MATLAB in ways which
C in a style close to FORTRAN, BASIC, or MATLAB. C is can affect the accuracy of computations. However, programs
an “open source” language, meaning it is available for free. will be fairly simple, as well as run quickly and accurately
Current programming languages are “object-oriented,” enough for our purposes.
as opposed to the earlier procedural-oriented programming
languages. The intent of object-oriented programming is to
create modular and “encapsulated” code, where “encapsu- Fundamentals of Procedural Programming
lation” means that the code can safely be used in a mod-
ern Windows-based program with millions of lines of code, One purpose in learning procedural programming is to au-
without causing conflict with other portions of the program. tomate numerical methods that were developed in the late
In the abstract, the modern, enormous computer programs 1800s and early 1900s, before the invention of the digital
can be written in the early procedural languages. In practice, computer. The algorithms were initially implemented using
however, developing and debugging gigantic programs writ- “desk calculators,” which could add and multiply. The early
ten in procedural languages would result in software over- desk calculators were purely mechanical, crank-operated de-
priced and obsolete, by the time it went to market. vices, which later evolved into electromechanical designs.
Object-oriented programming languages focus on the They were enormously complex and quite expensive. The
data, not on algorithms. At first, this sounds impossible. Of cost and complexity of these machines stimulated the design
what use are data without algorithms to manipulate and in- of purely electronic calculators, which led to the invention of
terpret them? Object-oriented programming languages pack- the microprocessor.
age the algorithms needed into a module called a “class,” Computer programs are instructions “coded” in the syntax
specific to each type of data, or “object,” in the program. specific to a particular computer language. Dozens of com-
A familiar example of a data type is a complex variable. puter languages exist. Different computer languages were
Complex variables are manipulated differently from real written to provide convenient methods of performing certain
Chapter 8 Appendix 495

tasks, such as creating attractive graphical user interfaces, Discrete (as opposed to integrated) transistors and “mag-
manipulating strings of characters, or performing numeri- netic core” memory took the place of vacuum tubes in the
cal calculations. Although computer languages can be very 1950s. Core memory was the dominant form of computer
different, there are fundamental, logical, and mathematical memory in the late 1950s through the early 1970s, and is
entities and operations which are common to all computer still used in a few military and aerospace applications for
languages. its reliability in extreme conditions. Core memory is a rect-
Three fundamental functions must be performed by a angular grid of orthogonal wires with tiny rings, or cores,
computer to execute an engineering algorithm. of a ferromagnetic material at the intersections. Currents
1. Store and retrieve data. through two wires magnetize the core at their intersec-
2. Evaluate conditional statements comparing the values of tion in either a clockwise or counter-clockwise direction,
two variables. depending on the direction of the currents. The direction
3. Perform arithmetic operations. of magnetic flux in the core is defined as 0 or 1. Discrete
We will focus on these fundamental aspects of program- memory was replaced by integrated circuits, beginning in
ming. Our development will begin with a brief overview of the late 1960s. The density of integrated circuits, as mea-
the physical and logical design of data storage. We will then sured by the number of transistors per unit area, has ap-
introduce “flowcharts” as a technique for graphical presenta- proximately doubled every two years.
tion of programming logic. Our first programming will use
the TI-89 calculator. This is historically appropriate, since Bits, Bytes, and Words
Texas Instrument introduced some important mid-scale in- We now consider the basic structure of computer memory.
tegrated circuits in the late 1960s. We will then move to The smallest element of computer memory is a single BIna-
MATLAB, a powerful high-level language. Finally, we will ry digiT, a bit. Groups of eight bits are called “bytes.” Bytes
compare MATLAB and the programming language, C. are elements of “words.” Microprocessors have fixed word
sizes from 8 to 32 bits in length, which increase with the
cost of the microprocessor. Current personal computers can
A Brief History of Computer Memory have words up to 64 bits in length, but vary the size of words
used, to make best use of memory. We will assume a 16-bit
Data are stored in the computer’s memory, exploiting a phys- word comprising two 8-bit bytes, which is the size of words
ical phenomenon or device, which can exist in one of two in many mid-level microprocessors used in machine control.
states (such as off or on; low or high; open or closed; de- Further, we will only discuss microprocessors to focus on
energized or energized; etc.), associated with the binary dig- the basics, and avoid the complexity of the clever addressing
its, 0 and 1. The earliest digital computers built during World schemes used by computers.
War II used vacuum tubes, or electromechanical relays as In order to store data in and retrieve data from a comput-
memory elements, where the state of the device represented er’s memory, there must be an “address”, a unique name, for
a binary digit, or “bit.” Vacuum tubes and electromechanical each word. The number of words, which can be addressed
relays were eventually replaced by “reed” relays, which are using N binary bits, equals the maximum value of binary
still commercially available. Reed relays are pairs of very number of N bits—in other words, a binary number of N bits,
thin steel cantilevers sealed in tiny glass vials. A coil around each of which is 1
the vial closes the contacts. Vacuum tubes consume a great
n = N −1
deal of power with lower reliability than reed relays, but they
“transition,” or change states, faster. Memory devices after
Number of Addresses = ∑
n=0
2n = 2 N − 1
World War II included “delay lines,” devices which use the
propagation time of a wave through a media to store informa- Example: The number of words which can be addressed with
tion, such as an acoustic wave through mercury; and “drum an 8-bit word is:
memory,” similar in function to current hard disk memory
but with a small fraction of the capacity and read, write, and
erase heads for each track.

Number of Addresses = 111111112


n = 8 −1
= ∑2
n=0
n
= 20 + 21 + 22 + 23 + 24 + 25 + 26 + 27

= 1 + 2 + 4 + 8 + 16 + 32 + 64 + 128 = 255 = 2 N − 1
496 8  Finite Difference Methods and MATLAB

The number of words which can be addressed by a 16-bit bi- is not an exact decimal fraction, since it cannot be expressed
nary number is 216 − 1 = 65, 536 − 1 = 65, 53510. Sixteen-digit as a finite sum of natural decimal fractions.
binary numbers are difficult to work with. The hexadecimal Positional notation is used within all number systems,
equivalent is used instead. including binary. Although the term, “binary point,” is com-
monly used in the United States, the proper term is “radix.”
11111111111111112 = FFFF16 Integer binary place values are to the left of the radix. Natu-
ral binary fraction place values are to the right, just as with
A common notation for addressing a bit within a byte or decimal numbers. Natural binary fractions and their decimal
a word is to number the bits from the least significant bit equivalents are familiar, because US customary units use
(LSB), to most significant bit (MSB). For example, the bit fractions which are powers of one half.
addresses in a 16-bit word range from 0 for the LSB, to F for
Table A8.2   Binary place values from 2–4 to 25
the MSB expressed in hexadecimal.
25 24 23 22 21 20 2−1 2−2 2−3 2−4
The bit number is appended to the word address after a 100000 10000 1000 100 10 1 0.1 0.01 0.001 0.0001
dot. For example, the address of the MSB in the 16-bit word 32 16 8 4 2 1 1 1 1 1
with the address, 352A, is 352A.F. The bit address, 352A.F, 2 4 8 16
is not a binary fraction. It is an address code. The dot is not
32 16 8 4 2 1 0.5 0.25 0.125 0.0625
a radix. It is just a dot.

Binary, Octal, and Hexadecimal Numbers The place values for the first 16 binary digits to the left of
The “base” of a number system is the number exponentiated the radix, expressed as decimal numbers, are
to yield the place values of that system. Decimal is base 10.
Table A8.3   Binary place values from 215 to 20
Binary is base 2. We will also work with octal, which is base
215 214 213 212 211 210 29 28
8, and hexadecimal, which is base 16. We write numbers using 32,768 16,384 8,192 4,096 2,048 1,024 512 256
“positional notation,” which we commonly refer to as “place 27 26 25 24 23 22 21 20
value.” In a decimal number, the decimal point, or radix, marks 128 64 32 16 8 4 2 1
the division between place values with exponents less than
zero, and place values with exponents greater or equal to zero Octal (base 8) was used to address memory in early com-
puters which had very little memory. Programmable logic
16-Bit Word controllers (PLCs) still use octal (base 8) for addressing
F E D C B A 9 8 7 6 5 4 3 2 1 0 memory, because they need relatively little.
1 0 0 1 1 0 0 1 1 0 0 1 1 0 0 1
Table A8.4   Octal place values from 87 to 80
8-Bit Byte 8-Bit Byte 87 86 85 84 83 82 81 80
Fig. A8.1   A 16-bit word is composed of two 8-bit bytes 2,097,152 262,144 32,768 4,096 512 64 8 1

Table A8.1   Decimal place values, including natural decimal fractions Hexadecimal (base 16) is used in modern computers,
105 104 103 102 101 100 10−1 10−2 10−3 10−4 because it allows large numbers to be represented easily,
100,000 10,000 1,000 100 10 1 1 1 1 1 and converts easily and exactly to binary. The decimal
10 100 1,000 10,000 number system has 10 digit symbols, zero through nine,
100,000 10,000 1,000 100 10 1 0.1 0.01 0.001 0.0001 and the hexadecimal system needs 16 digit symbols, zero
through fifteen. The hexadecimal digit symbols for the
decimal numbers, 10 through 15, are the Roman letters,
“Natural” decimal fractions are the place values to the A through F.
right of the radix. The only decimal fractions which are exact
Table A8.5   Hexadecimal digits
are those which can be represented by sums of the natural
Decimal 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
decimal fractions. For example,
Hexadecimal 0 1 2 3 4 5 6 7 8 9 A B C D E F 10
1 5 1 1 1 1 1
= = + + + + = 0.5
2 10 10 10 10 10 10

is an exact decimal fraction, whereas

1 3 3 3 3 3 3
= + + + +  + n +  = 0.333333
3 10 100 1, 000 10, 000 100, 000 10
Chapter 8 Appendix 497

Base Conversion Yielding the conversion from decimal to binary:

Base conversion allows one to represent a given number in 47.810 = 101111.11002


different number systems. It can always be accomplished by
performing successive division, starting with the largest di- Note the error of 0.0510 between the binary and decimal
visible place value, and working down in place value. The numbers, even though the base conversion was to four bi-
process is repeated on the remainder of the each division, nary places to the right of the radix. It is always possible
until the conversion is carried to the precision desired. to perform an exact conversion of integer values between
We will illustrate the technique by converting the decimal bases. However, it is often impossible to convert fraction
number, 47.8, into binary number, using the binary place val- values between bases exactly.
ues from 25 to 2–4. The decimal number, 32 = 2 5 , is divisible, A second example: Convert 978 (base 10) to octal (base 8).
leaving a remainder of 15.8. The next binary place value, First, determine the octal place values in base 10 to a suffi-
16 = 2 4 , is not divisible. The process is repeated. ciently large value to perform the division.

   
       
    HWF
   

The results are Perform successive division on the number to be convert-


ed to octal, 97810
Binary Place Values 25 24 23 22 21
Dividend 47.8 15.8 15.8 7.8 3.8
Divisor 32 16 8 4 2
Division possible? Yes No Yes Yes Yes
Binary Digits 1 0 1 1 1
Remainder 15.8 15.8 7.8 3.8 1.8
Binary Place Values 20 2−1 2−2 2−3 2−4 84 83 82 81 80
Dividend 1.8 0.8 0.3 0.05 0.05 Base 10 Divisors 4,096 512 64 8 1
Divisor 1 0.5 0.25 0.125 0.0625 Value of 97810 in Octal 0 1 7 2 2
Division possible? Yes Yes Yes No No
Binary Digits 1 1 1 0 0
Remainder 0.8 0.3 0.05 0.05 0.05 The result is 97810 = 17228. The method of successive divi-
sion always works, but is not necessary when one is convert-
ing between number systems that are powers of a common
base, such as binary (base 2 = 21), octal (base 8 = 23), or hexa-
decimal (base 16 = 24). There is a shortcut.
Numbers can be converted among binary, octal, and hexa-
decimal by exploiting the power relationship between the bases.
Compare the place values of binary, octal, and hexadecimal:

%LQDU\3ODFH9DOXHV
   
              
            

2FWDO3ODFH9DOXHV
    
    

+H[DGHFLPDO3ODFH9DOXHV
   
   
498 8  Finite Difference Methods and MATLAB

The relationship between the nth octal place value and the Summarize the decimal, binary, octal, and hexadecimal
corresponding binary place value is 8n = (2n)3. Similarly, values:
the relationship between the nth hexadecimal place value,
and the binary place value is 16n = (2n)4. Consequently, 93510 = 11101001112 = 16478 = 3A716
it requires three binary digits, or places, to represent an
octal number, and four binary digits to represent a hexa-
decimal number. The shortcut is to convert the number to Computational Error on Conversion from Natural
the common base, binary in this case, and then to either Decimal to Natural Binary Fractions
octal or hexadecimal base, by separating the binary num-
ber into groups of three or four binary digits, respectively, As we have seen, integer values can be converted between
and converting the binary group into a single digit in the different bases exactly. However, fractional values may be
higher base. impossible to express exactly within a base. If they can be
Example. Express the decimal number, 935, in binary, expressed exactly within one base, we may not be able to
octal, and hexadecimal. Base 10 does not share a common express them in a second base. We touched on the compu-
base with base 2, base 8, or base 16. The conversion from tational error inherent in conversion between decimal and
decimal to binary must be accomplished by successive divi- natural binary fractions above. It is an important topic, which
sion, as above. Summarizing the result, the binary value of deserves elaboration.
93510 is

         
         
         

 

We can now readily convert to octal, by dividing the binary


number into blocks of three binary digits. Then sum each
block of three binary digits to yield its octal value

         
         
         
     
    

 

Likewise, we can convert from binary to hexadecimal, by


dividing the binary number into blocks of four digits. Sum
them to yield their hexadecimal value.
Now, convert from binary to hexadecimal:

           
           
           
           
       { $     

 $
Chapter 8 Appendix 499

Natural binary fractions can be expressed exactly as Clearly, simply summing the binary fraction approximation
decimal fractions, if enough significant figures are used. of a decimal fraction is an unacceptable timing algorithm.
Example: Sadly, many engineering errors have been made through

1 0 1 5 6 2 5
= + + + + + = 0.015625
64 10 100 1, 000 10, 000 100, 000 1, 000, 000

However, beware, it does not work the other way. Most natu- ignorance of the magnitude of the error, in converting be-
ral decimal fractions cannot be expressed exactly as binary tween natural decimal fractions and natural binary fractions.
fractions! For example, we can approximate 1 These errors could have been prevented by using natural bi-
10 10 as nary fractions, which can be converted to decimal fractions
­exactly.
 1 1  1  1 1 1 
 16 + 32  < 10 <  16 + 32 + 64 
10 10 10 10 10 10

Data Types
0.0001102 < 0.110 < 0.0001112
The binary numbers used for word addresses are positive bi-
nary integers. How does a computer store negative binary in-
0.0937510 < 0.110 < 0.10937510 tegers or “real” binary numbers, i.e., binary fractions? Also,
how are large numbers, those greater than 2N − 1, stored?
Note additional binary places do not solve the problem. Also, They are coded, and in order for a computer to manipulate
binary fractions may appear smaller than their corresponding numerical data, it must know the coding, known as the “data
decimal fractions, but they are not. type,” associated with the data, in a specific word (or words)
Example. Suppose you have a timer, timing by one hun- in memory. Some languages will assume a default data type,
dredth of a second by summing the equivalent binary frac- if the programmer does not “declare” a variable to be a cer-
tion. How much error would you accumulate in an hour? tain data type. Other languages will not compile a program,
First, we have to choose our binary fraction approximation if all variables used are not declared and assigned initial val-
of 1 ues. The discussion that follows is general, since the specific
10010 term used for a generic data type, and the number of bits
associated with it, varies with both computer and program-
1 1  1 1 
< < + ming languages.
12810 10010  12810 25610 

0.007812510 < 0.0110 < 0.01171875010


Integer and Signed Integer Variables
0.0000000102 < 0.0110 < 0.000000112
The simplest data type is a positive binary integer. No cod-
Both available approximations have substantial error: ing is needed, other than to associate place value with loca-
tion within a word, since increasing numbers of bytes can be
0.007812510 − 0.0110 = −0.00218750010 used to store integers. Various adjectives and acronyms are
0.01171875010 − 0.0110 = 0.00171875010 used to distinguish between lengths of the integer data types.
(An example of awkward terminology is C++ ’s “long long
integer” which is, as you would expect, longer than C++ ’s
“long integer”.)
Using the approximation with less error per count, The “signed integer” data type requires coding to dis-
0.0110 ≈ 0.000000112 , in one hour, the accumulated error is tinguish between positive and negative values. The sign of
the integer is represented by setting the most significant bit

 sec error   counts   sec   min  sec error


 0.00171875010  100   60   60  = 618.75
count sec min hour hour
500 8  Finite Difference Methods and MATLAB

(MSB) to 1, to indicate a negative number. Negative num- Boolean or Logical Variables


bers are stored in two’s complement form. Two’s comple-
ment form is created in two steps. First, every bit in the num- Boolean variables, also known as “logical” variables are a
ber, except the sign bit, is inverted. If a bit is 0, then it is late addition to programming languages. They are unneces-
changed to 1. If a bit is 1, then it is changed to 0. Next, the sary, since one can use integer variables with values of either
number, 1, is added to the number. 0 or 1 in their place. They were added to programming lan-
Example. Create the two’s complement of the binary guages to prevent errors that result from integer variables,
number, − 1101011, by using a 16-bit word. which were to be restricted to the two Boolean states, being
Sign bit (MSB) set to 1, followed by the binary number, accidently assigned other values.
1101011:

1000000001101011 Procedural Logic and Flowcharts

Bits 0 through E are inverted: The numerical methods used in engineering can be very
sophisticated. Some required decades to reach their current
1111111110010100 form. Fortunately, writing a computer program to imple-
ment an existing numerical method is far less of a challenge.
One is added to create the two’s complement: Most of the effort is spent in developing an understanding of
the numerical method, before the program is written, then
1111111110010101 searching for any coding errors afterwards.
Flowcharts are a graphical representation of the sequence,
We will test the usefulness of two’s complement by perform- in which the instructions of a program are to be executed.
ing the addition, − 1101011 + 1101011, which should yield Although not all types of programs can be depicted in a flow
zero: chart, our programs can. A flowchart presents the logic of
a program, without the details of the specific programming
1111111110010101 ← Two’ s complement of 1101011 language syntax. Flowcharts serve the same function for
+ 0000000001101011 computer programs, as outlines serve for papers. Flowcharts
0000000000000000 show the flow of “logical control” through an algorithm.
They consist of “blocks,” which contain instructions or op-
Floating Point Variables erations, and arrows connecting the blocks to indicate the
flow of logical control.
“Floating point” is the term used by computer scientists for Rectangular blocks are “executable” blocks, which repre-
scientific notation, i.e., the product of a decimal fraction be- sent actions or calculations. The action or calculation may be
tween 1 and 10, and a multiplier which is a power of 10. a single, simple step, or it may be many complicated steps.
Varying the exponent has the effect of moving the radix or The action or calculation may be vaguely described in Eng-
decimal point, hence the term floating point. The coding of lish or expressed explicitly in mathematics. There can only
floating point numbers is more involved than for signed in- be one logical path entering the block and only one logical
tegers, since both the mantissa and the exponent are signed. path exiting the block. The direction of the logical flow is
We will not delve into details, other than to note that more indicated by arrows.
bits are allotted to the mantissa than the exponent. For ex- Diamond-shaped blocks are “decision” blocks, per
ample, if the floating point data type was 4 bytes or 32 bits Fig. A8.3. They contain conditional statements, i.e., greater
long, the mantissa may be 24 bits and the exponent 8 bits. than, less than, equal to; or a Boolean or Logical variable;
“Double” is a common name for the data type with twice, or or a result which must be either True or False. There can
double, the internal storage and, hence, numerical precision only be onelogical path into the decision block. There are
of the floating point data type. two logical paths which exit. Each path must be identified by
Computer code is restricted to the characters of a key- the logical result or value it represents, i.e., Yes or No; True
board. Consequently, the product of the decimal fraction or False; etc. The logical flow “branches” at decision blocks.
multiplied by the power of 10 is expressed, omitting the base The exit from the decision block taken by the logical control
of 10 and preceding the exponent of 10 with E (or e) for is governed by the result of evaluating the conditional state-
“Exponent,” e.g., ment, value of the Boolean variable, or logical result. A dia-
mond shape is standard for a decision block, but that shape
352.478 → 3.52478 × 102 → 3.52478E02 can become cramped for long expressions. This text will also
Procedural Logic and Flowcharts 501

Fig. A8.2   An “executable”


block represents an action or
calculation

Compute output variable y


y = n2 True
x < 0.001 OR y < 0.2 ?

False

Fig. A8.3   The standard Fig. A8.4   An alternative, non-standard, decision block
diamond-shaped decision block

Fig. A8.5   A logical


loop. “Control” exits
Enter Loop
No
n > 10 ? the loop, when the
condition expressed in
the diamond-shaped
decision block is
Yes satisfied

Instructions for
actions or computations

use, as needed, a non-standard, “alternative style,” decision


block, per Fig. A8.4.

Logic Loop No
Done ?
An important use of a decision block is to form a “loop,” per
Fig. A8.5. A loop is a logical path that passes through the
same blocks more than once. Loops are used for repetitive Yes
calculations. The instructions within the loop are executed
repeatedly, either for a fixed number of times, or until a logi-
cal condition is satisfied. If the programmer errs and neglects Exit Loop
to create an exit condition, or creates an exit condition that
is impossible to satisfy, then the logic is trapped in an “end-
less loop.” one’s own use. There are two rules, which must be respected
for the flowchart to be logical.
1. There can only be one logic path leaving a block, except
Nested Loops decision (diamond) blocks which have Yes and No or
True and False paths.
It is common to use “nested” loops with one loop inside an- 2. Branching can only occur at a decision block.
other, Fig. A8.6. Each loop has a counter variable. The inner The following are a few guidelines which will eliminate er-
loop’s counter variable must complete its index through its rors and improve the readability of flowcharts.
range for the outer loop’s counter variable to index once 1. Every block in the diagram should be on at least one logi-
nested loop are useful for working with arrays and matrices. cal path from the Start block to the Finish block. If a block
cannot be reached from the Start block, then the instruc-
tions in that block will never execute. If the path from a
Flowchart Rules and Guidelines block does not eventually reach the Finish block, then the
logic will never properly terminate. A common error is to
Although an international standard for flowcharts exists, terminate a path at a result or action, other than Finish.
there is no reason to abide by it, when drawing flowcharts for
502 8  Finite Difference Methods and MATLAB

Enter Outer Loop


of the Nested Loops

Instructions for
actions or computations

Instructions for
actions or computations

Enter Inner Loop

Instructions for Instructions for


actions or computations actions or computations

Instructions for
actions or computations

Fig. A8.7   Flowchart with illogical branching occurring outside of a


decision block. No logical conditions for the branching are shown

No Inner
Loop Tasks
Done ?
Start

Yes
Exit Inner Loop
Instructions for
Instructions for actions or computations
actions or computations

Instructions for No
Conditonal
No Outer actions or computations True ?
Loop Tasks
Done ?
This is a logical “dead-end”
Yes
Yes

Instructions for
actions or computations
Exit Outer Loop

Fig. A8.6   Nested loops

Finish
2. Have only one path into a block. If there are multiple
paths leading to the block, have them join at a node before Fig. A8.8   Flowchart with dead-end logic
the block, then have a single path into the block.
3. Draw the logic, so that sequential blocks are arranged
from top down, or side to side. Do not arrange sequential the two flowcharts. In Flowchart A, the counter variable, n,
blocks to read from bottom to top. is “incremented,” or increased, by one with the instruction:
Flowcharts A and B, Figs. 8.11 and 8.12, illustrate the use of a
decision block to control the number of times that logic loops n = n +1
through a set of instructions. In both flowcharts, a “counter”
variable, n, is assigned the value of one before the logic en- before the logic enters the decision block with the rela-
ters the loop. The loops in both flowcharts contain identical tional expression, n = = 10? In Flowchart B, the logic enters
instructions. The first instruction in both loops computes the the decision block and evaluates the relational expression,
output variable, y, by squaring the counter variable, n. How- n = = 10?, before the counter is incremented.
ever, the order of the two remaining blocks is different in
MATLAB 503

Fig. A8.9   a Two branches en-


tering one block, logical but can
a b
be difficult to read. b Preferred
flowchart. Flow merges at a Instructions for Instructions for Instructions for Instructions for
node before entering the block actions or computations actions or computations actions or computations actions or computations

Instructions for Instructions for


actions or computations actions or computations

Start

Instructions for
Instructions for
actions or computations Initialize counter variable
actions or computations
n=1

Instructions for
Instructions for actions or computations
actions or computations

Compute output variable y


y = n2
No
Done ?

Increment counter by one


Yes n=n+1

Fig. A8.10   Flowchart with upside down flow: logical, but difficult to
read and, therefore, error prone
No
n == 10 ?

The program shown in Flowchart B will loop 10 times,


as n is incremented from n = 1 to n = 9. When n = 9, the Yes
program terminates. The program shown in Flowchart B will
loop 10 times, as n is incremented from n = 1 to n = 10.
When n = 10, the program terminates. Finish
The effect of the relative position, where the counter vari-
able is incremented within the loop, and the relational expres- Fig. A8.11   Flowchart A
sion, which terminates execution of the loop, is straightforward.
It is a very common error, however, to have a loop execute one
more time, or one fewer time, than the engineer intended. figuration when it is started up. In a multi-user location, such
as a computer lab, the MATLAB “environment” (or set of
windows), when opened, will reflect the preferences of the
MATLAB previous user.

Modern programming languages provide an “environment,”


which is a desktop and a set of windows. The menu, “Desk- MATLAB “Environment”
top,” contains items which allow you to open needed win-
dows. We will use the “Command” window and the “Editor” The MATLAB “environment” includes the Command win-
window. MATLAB retains the configuration of the environ- dow, the Editor window, and the Help window, among others.
ment when the program is closed, and returns to that con- The Editor is well hidden. MATLAB’s Editor opens, when
504 8  Finite Difference Methods and MATLAB

while to recognize that script is not yours, particularly if you


Start
failed to add comments.
MATLAB is an interpreter, meaning, it will happily ex-
ecute a script, or the same commands entered at the prompt
Initialize counter variable
in the Command window. The variables used during a work
n=1 session are retained, until the program is closed. They are
displaced in the Workspace window. Retention of previous
variables is very convenient, when you are using MATLAB
as a calculator, and entering computations at the Command
window prompt. Once you have defined a variable, you may
use it for the entire work session.
Compute output variable y However, there is a dark side to MATLAB’s workspace.
y = n2
The existence of variables in the workspace can mask errors
and omissions in scripts. For example, suppose you write a
script to calculate a response function. You intend to define a
variable named tau as being equal to a certain value, but you
Increment counter by one No simply forget, and omit that line of code. If you run the script,
n == 10 ?
n=n+1 and tau was not used in any previous calculations, so that it
is not in the workspace, MATLAB will identify the error. A
Yes ding sounds. The Command window contains an error mes-
sage, indicating that an undefined variable was used in line x
of the script. On the other hand, if you omit the definition of
Finish tau in the script, but another tau already exists in the work-
space, MATLAB will use the workspace value. The script
Fig.  A8.12   Flowchart B will execute but produce an erroneous result.
Engineers tend to use and reuse the same meaningful
variable names. MATLAB does not create a new instance
the user clicks on the New Script button. A script is a short of an existing variable. So if a MATLAB execution error oc-
program. The term, script, came into use after the size of curs, or the results of a successful execution of a script don’t
programs became enormous. Scripts, or programs, are writ- make sense, clear the workspace, and run the script again.
ten in the Editor window, saved as an “m-File” (a file with The entire workspace is cleared by selecting the title bar of
extension file_name.m), and then run from the Command the Workspace window, right clicking, and choosing Clear
window. MATLAB does not permit spaces in file names. Workspace. Individual variables are deleted, by selecting the
MATLAB looks in a number of locations to find m.files, variable in the Workspace window, right clicking, and choos-
including the default directory. The collection of directories ing delete. MATLAB’s retention of variables is a particular
and folders checked for m.files is known as MATLAB’s nuisance, when debugging a script. It is common to change
path. If you write a script and attempt to save it to your flash the number of iterations within a for loop, while debugging.
drive, you will see a MATLAB message stating the folder is When results of a computation are written to a vector, or an
not in MATLAB’s path, and ask if you would like to add it. array, and the number of iterations is increased, no problem.
Accept this choice, so that MATLAB will look at your flash If the number of iterations is decreased, the values beyond
drive for a script. the most recently overwritten elements are unchanged. A
A MATLAB program is run, by typing the name of the plot may make sense, nearly to the end, before becoming
m-file without the extension, “.m,” at the prompt in the quite strange. When faced with inexplicable results, clear
command window. Example: If you have an m-file named MATLAB’s workspace first, before tearing apart your script.
Model5.m, run the program by typing only Model5.m, at the
prompt in the command window. MATLAB will look in its
default directories to find the m-file. Consequently, when Variables
you open MATLAB on a public computer, click on the Set
Path button in the Environment tab, and add your flash drive MATLAB does not require variables to be “declared,” mean-
folder to the path. If you forget to do this, and you have cho- ing the types of data assigned to variables do not have to
sen a common name for your script, then MATLAB may find be identified before they are used. Integer and floating point
someone else’s script of the same name saved to one of its variables are identified by their use. Integer variables are as-
default directories, and run that one instead. It may take a signed integer values, i.e., numbers without a decimal point.
MATLAB 505

Floating point variables are assigned either decimal numbers Many computer languages will not allow you to use a logi-
or numbers expressed in scientific notation. cal variable in a numerical calculation. MATLAB is more
All variables must have a name. In MATLAB, variables forgiving. Type
names must begin with a letter and may contain numbers
and underscores, but no other special characters. MATLAB & DPDVV
is case sensitive.
Values are assigned to variables with an “assignment” state- and you will find the result is 4.5000.
ment. The assignment operator in MATLAB is the conventional
equal sign, =. Type the following statement in the Command
window to assign the value of 3.5 to the variable “mass.” Scalar, Matrix, and Array Variables

PDVV  MATLAB was written to facilitate vector-matrix mathemat-


0$7/$%HFKRHV\RXULQSXWDV ics. MATLAB does not require the user to declare a variable
as a scalar, vector, matrix, or an array. Scalar, vectors, and
!!PDVV  matrices have their conventional mathematical definitions,
PDVV  based on the number of elements and the number of rows
 and/or columns. Arrays resemble vectors and matrices, as
both comprise elements in rows, columns, or both rows and
An assignment statement assigns the value of the right side columns. MATLAB distinguishes between matrices and ar-
to the variable named on the left side. Assignment statements rays by the way one can operate on them. Matrix operators
are not equations! Typing the statement in the reverse order obey the rules of linear algebra. Array operators operate ele-
(capitalizing the “M” in Mass to make it a different variable) ment by element.
produces an error. The following assignment statement creates a vector or an
array. Important: Note the use of square brackets!
 0DVV
9 >@
MATLAB thought you tried to name a variable, 3.5. This
violates two of its rules: (1) variable names must start with a Note spaces separate the elements within brackets. Alterna-
letter, and (2) the only permissible non-alphanumeric char- tively, the elements within brackets can be separated by com-
acter is an underscore. Periods, points, or dots are not permit- mas, or by a mixture of commas and spaces.
ted in variable names.
Variables can be assigned strings of characters. Charac- 9 >@RU9 >@
ters or strings of characters are identified by pairs of sin-
gle quotes. Example: type the following valid assignment A semicolon terminates a row. Example:
statement
0 >@
$ µ(QWURS\URFNV¶
Is echoed to the screen as:
Note how MATLAB color-codes the input as you are typing.
!!0 
When you type the lead single quote, MATLAB highlights

the single quote and the following characters in dark red,

until the closing single quote is typed. It then changes the
color to purple to indicate the expression is complete.
Logical variables are defined by assigning a value of ei- Individual elements can be addressed using row and column
ther true or false to the variable. MATLAB recognizes true subscripts within parentheses. Example: The element, 7, is in
as equivalent to the Boolean value 1, and false as equivalent the second row and third column of the matrix (or array), M4.
to the Boolean value 0. When you type
!!9  
D WUXH DQV 

MATLAB echoes

D  The lowest index in MATLAB is 1, not 0!


506 8  Finite Difference Methods and MATLAB

MATLAB is very tolerant with vector-matrix mathemat- Table  A8.6   MATLAB’s arithmetic operators
ics. It will execute operations which are not defined in linear Operator Description
algebra, by interpreting your vectors and matrices as arrays. + Addition
For example, in conventional vector-matrix mathematics, if − Subtraction
b is a scalar, and B is a row vector, e.g., Array multiplication
.*
E  ./ Array right division
.\ Array left division
+ Unary plus
and
− Unary minus
% >@ Colon operator
.^ Array power
then the sum .’ Transpose
’ Complex conjugate transpose
& E%
* Matrix multiplication
/ Matrix right division
is not defined. MATLAB, however, will execute this state-
\ Matrix left division
ment and echo
^ Matrix power
& 

Arithmetic Operators
MATLAB has a type of array, called a “cell array,” to handle Operators and the symbols used for operators vary tremen-
non-numerical data. It refers to the element of an array as a dously across programming languages. The only operators,
“cell.” The assignment statement for an array uses “braces” whose meaning will not waver, are the arithmetic operators
(which MATLAB refers to redundantly as “curly braces”). for multiplication, division, addition, subtraction, as long as
The assignment statement they are used as “binary” operators (with two operands), as
they were in elementary school. Many operators, other than
*BFHOOBDUUD\ ^WZRWKUHHIRXU` dear Aunt Sally, are included in a programming language.
The set of operators varies with the intended application of
is echoed as the language. For example, languages developed to program
webpage applications will have operators to parse through
*BFHOOBDUUD\  strings of character data. MATLAB was written for scientific

WZR

WKUHH

IRXU
 and engineering computation, so it has more mathematical
operators than most languages.
Note the individual cell data are maintained. Contrast this Symbols to represent operations were limited to those on
result with a conventional array, defined by the use of square the computer keyboard. Consequently, operators are repre-
brackets, containing the same data. The assignment state- sented by combinations of symbols that made sense to the
ment programmer, but do not follow any convention and vary
widely between languages. For example, exponentiation of
*BDUUD\ >
WZR

WKUHH

IRXU
@ a to the b power, ab, is represented as the function pow(a,b)
in C, a**b in Python and FORTRAN, and a^b in MATLAB.
is echoed as We will use these arithmetic operators in MATLAB: the
elementary school arithmetic operators of multiplication,
*BFHOO  division, addition, subtraction; the “unary” plus and minus
WZRWKUHHIRXU operators (which are simply the plus and minus signs); and
exponentiation. The complete set of arithmetic operators is
presented in Table A8.6.
Operators This set requires some explanation. MATLAB uses the
All programming languages provide at least three differ- terms, “matrix operations” and “array operations,” to distin-
ent types of operators: arithmetic, relational, and logical. guish between conventional linear algebra operations per-
Arithmetic operators are used in computations. Relational formed with vectors and matrices from those operations not
operators are used to define conditions to control the flow defined in linear algebra, which are performed with two ar-
of a program’s execution. Logical operators are used for rays of the same size and shape, or on an array and a scalar.
Boolean logic.
MATLAB 507

An array operation is performed “element by element,” Table A8.7   MATLAB’s array and matrix operators
which means the operation is performed repeatedly, using Operator Description Conventional Linear
Algebra?
elements in the same row and column position in each of
+ Addition Yes
the two arrays, until the operation has been performed on − Subtraction Yes
all the pairs. Note array operations are preceded with a dot .* Array Multiplication No
(“.”), except for addition and subtraction, which are the same ./ Array Right Division No
as matrix addition and subtraction, since they are performed .\ Array Left Division No
element by element. For illustration, define two square ar- + Unary Plus Yes
rays A and B: − Unary Minus Yes
Colon Operator No
 a11 a12   b11 b12  .^ Array Power No
A= and B = 
 a21 a22   b21 b22  .’ Transpose Yes
’ Complex Conjugate Transpose Yes
* Matrix Multiplication Yes
/ Matrix Right Division No
Array Multiplication:
\ Matrix Left Division No
^ Matrix Power Yes
 a11b11 a12b12 
A. * B = 
 a21b21 a22b22 
then the C transpose is

Array Right Division:  c11 c 21 


C.′ =  c12 c 22 
 a11 a12   
b  c12 c 22 
11 b12 
A. / B =  
a
 21 a22 
 b b22  MATLAB recognizes both i and j as imaginary numbers
21
when a complex number is entered as

Array Left Division: z = 2 + 3i or z = 2 + 3j

 b11 b12  However, when MATLAB echoes back a complex number, i


a a12  is used as the imaginary number.
11
A. \ B =  
The MATLAB operator, which transposes a complex
b
 21 b22 
 a a22  number to obtain its complex conjugate, is “ ‘ ”.
21

]
  L
Array Power:
The Colon operator, “:”, creates a range variable. Example:
 a b11 a12b12  Type:
A. ^B =  11b21
a 21 a22b22 
!!D 
An exception to the syntax of preceding array operators with
a dot is the transpose operator, “.’ ”. The transpose operator and MATLAB will echo
can operate on vectors, matrices, and arrays. It is a “unary”
operator, since it has only one operand. The transpose opera- D 
tor interchanges the order of the elements to move an ele- 
ment from the rth row and cth column to the cth row and rth
column. Example: If In this case, the variable, a, is now a vector or an array. The
fifth element is
 c11 c12 c12 
C= !!D  
 c 21 c 22 c 22 
DQV 

508 8  Finite Difference Methods and MATLAB

To increment by a value other than unity, indicate the first would accept the statement, 0.99999 = = 1, as true in most
value of the series, the increment, and the upper limit. Ex- circumstances, but to MATLAB, the statement is false. Do
ample: The statement not use the Equal relationship operator with floating point
(scientific notation) variables, since they are likely to fail the
!!E  comparison due to minute numerical error.

yields two values: the initial value and one increment. The Logical Variables
second increment would exceed the upper limit of the series Boolean logic is performed in MATLAB using “logical”
variables which are defined or created by assigning them a
E  logical value of either true or false. Note that true and false
 are lowercase. Numerical data can be converted to logical
values with the function logical(). If the argument is posi-
MATLAB includes two odd operators, Matrix Right Divi- tive, the result is true. If the argument is negative or zero, the
sion and Matrix Left Division. Matrix division is NOT de- result is false.
fined in linear algebra. The matrix “division” implemented
in MATLAB is a shortcut for matrix inversion. Although ma- Logical Operators
trix “division” may be advantageous in some circumstances, MATLAB provides the three fundamental Boolean opera-
it is better to use matrix inversion. tions, And, Or, and Not, in two forms, as operators and as
functions.
Relational Operators
Table A8.9   MATLAB’s logical operators
A computer language’s arithmetic operators are used for
Logical operation Equivalent function
crunching numbers. The language’s relational operators are
A&B and(A, B)
used to compare the values of variables. The operators allow
A|B or(A, B)
a computer to “think,” by branching the flow of the execu-
~ A not(A)
tion on the outcome of relational expressions. Relational op-
erators produce a result that is a Boolean, or “logical,” value,
i.e., either true or false. For example, if A and B are numbers, Boolean operators are used with relational operators to
then the answer to the question, “Is A greater than B?” must construct conditional statements that control which program
be either true or false. instructions are executed. The Boolean operators are best
MATLAB’s relational operators are given in Table A8.8. described using “truth tables,” where A and B are Boolean
variables which must have values of either true or false.
Table A8.8   MATLAB’s relational operators
< Less than Table A8.10   and truth table
> Greater than A B A&B
< = Less than or equal to 0 0 0
> = Greater than or Equal to
1 0 0
= = Equal
~ = Not equal 0 1 0
1 1 1

Only the last two operators require discussion. Note the


Table A8.11   or truth table
relational operator, Equal, consists of two equal signs. A com-
A B A|B
mon error is to confuse the single equal sign, which is the as-
signment operator, and the double equal signs, which signify a 0 0 0
relational equal operator. Also note the symbol for Not Equal, 1 0 1
which is unusual. Most programming languages use the excla- 0 1 1
mation point, “!” for the “Not,” or logical inversion operator. 1 1 1
Most engineers would read ~ = as approximately equal, which
is why MATLAB chose it to represent Not Equal.
Table A8.12  not truth table
The distinction between approximately equal and equal is
A ~A
a warning for the use of the Equal operator in an if statement.
The Equal relational operator requires an exact mathematical 0 1
or logical equality, not an engineering equality. An engineer 1 0
MATLAB 509

Programming Statement Syntax is executed. Echoing the code and the results to the monitor
during execution of a program slows the execution consider-
A computer programming language’s syntax is why com- ably. Consequently, you will want to place semicolons at the
puter code is called code. A programming language’s syntax end of line in loops which will execute repeatedly.
is restrictive and specific, so that the characters read by the
machine can be interpreted as instructions.
Control Flow Statements

Assignment Statements if Statement


A fundamental instruction in all languages is an “if” state-
A program is a list of instructions coded in the programming ment. The syntax varies with the language but the essence is
language’s syntax. It is important to recognize that program if (expression A Relational Operator expression B) is
statements are not equations. For example, the following as- true then execute the following instructions
signment statement in the flowcharts Figs. A8.12 and A8.13 In MATLAB, the syntax of an if statement is:
is a valid statement in all computer languages, even though it
is not a valid equation, except in Boolean algebra if (expression A Relational Operator expression B)
Instruction executed if conditional is true
Q Q  Instruction executed if conditional is true


As an algebraic equation, this expression is nonsense. No Instruction executed if conditional is true
number can be equal to itself plus one. As a program state- end
ment (or instruction), however, the expression not only
makes sense, it is very important. It is the instruction to in- The instructions appear between the conditional statement
crement the value of the variable, n, by one. Recall this ex- with the relational operator and the end statement. Important:
pression is called an assignment statement, and a variable MATLAB is case sensitive. The keywords, “if” and “end,”
name is an address in the computer’s memory. This assign- are both lower case. MATLAB highlights the leading “if”
ment statement is understood by the computer as saying, and the closing “end” in blue to demarcate the limits of the
“Write (or copy) the contents of the memory address(es) as- instructions, which will execute if the conditional statement
signed to the variable, n, to the specialized hardware in the is true. To further demarcate the if statement as a “block” of
central processing unit called the accumulator, which per- code, it is good practice to indent the instructions, which are
forms operations on binary numbers. Add one to the value executed by the if statement.
in the accumulator. Write the contents the accumulator to the MATLAB and many other languages provide two addi-
memory address(es) assigned to the variable n.” tional instructions to supplement the if statement. They are
Assignment statements are read from right to left. The “else” and “elseif.”
right side can be any valid mathematical statement expressed
using MATLAB’s notation, with the restriction that variables else Statement
cannot be used, before they have been assigned a value. If The “else” instruction is used to divide the block of instruc-
we use the expression, n = n + 1, before giving n an initial tions between an if statement and its corresponding end
value (referred to as “initializing” the variable), a red error statement into two blocks. The first block is executed, if the
message would state conditional statement is true. The second block is executed,
if the conditional statement is false. Example:
??? Undefined function or variable ‘n”’
if (expression A Relational Operator expression B)
The following code would execute properly Instruction executed if conditional is true
Instruction executed if conditional is true
Q   

Q Q  Instruction executed if conditional is true


else
Many languages require a character, such as a semicolon, Instruction executed if conditional is false
to indicate the end of an instruction. MATLAB does not. Instruction executed if conditional is false

In MATLAB, a semicolon at the end of a line prevents the


Instruction executed if conditional is false
code from being written to the monitor, when the program
end
510 8  Finite Difference Methods and MATLAB

elseif Statement for counter = lower integer limit: upper integer limit
The “elseif” instruction is a second conditional statement Instructions
that follows an if statement. It is executed only when the Instructions
conditional statement of the if statement is false. Example: Instructions
end
if (expression A Relational Operator expression B)
Instruction executed when if conditional is true The following code would execute the computation flow
Instruction executed when if conditional is true charted above

Instruction executed when if conditional is true for n = 1:10


elseif (expression C Relational Operator expression D) y = n^2
Instruction executed when elseif conditional is true end
Instruction executed when elseif conditional is true

“Nested” for loops are often used. The following code pro-
Instruction executed when elseif conditional is true
duces a matrix, or array, named A with two rows and three
end
columns. The “inner” for loop increments through the col-
umns. The “outer” for loop increments through the rows.
while Loop
A “while loop” uses a conditional statement to control the re- for r = 12
peated execution of a block of code. The structure of a while for c = 13
loop resembles an if statement. A(r,c) = (r+c)^2
end
while (expression A Relational Operator expression B) end
Instructions
Instructions The result is
Instructions
end $ 

The while loop differs from the if statement, in that the block 
of instructions of an if statement are executed only once,
if the conditional statement is true. The block of code in a Example: for Loop and while Loop
while statement is executed repeatedly, as long as the con- The following example is a first-order finite difference algo-
ditional statement remains true. It is possible to code an “in- rithm, programmed using a for loop to control the number of
finite loop.” If you ever need to interrupt the execution of a iterations through the loop:
MATLAB program, type CTRL + C simultaneously.
The following code with a while loop executes the compu- 1 
tation flow charted above. WBHQG 
D 
n=1 E 
while (n ~= 10) [   
y = n^2 X 
n=n+1 7 WBHQG1
end W   
IRUQ 1
'[ Q  D [ Q 7E X 7
for Loop [ Q  '[ Q [ Q 
A “for loop” is a block of code which executes for a fixed W Q  Q 7
number of times. The number of executions is controlled by HQG
giving a counter variable a range, using the colon operator. SORW W[
The block of code which executes repeatedly is demarcated
with the key word, “end.” A MATLAB for loop has the fol-
lowing form:
MATLAB 511

Here is an alternative program, using a while loop to control % Input matrix elements
the number of iterations b1 = 1/M
b2 = 0.0
1  % Initial values of state variables
WBHQG  %
D  x1(1) = 0.0 % mass velocity v1g
E  x2(1) = 0.0 % Force acting through spring FK
[    %
X  u = 10.0 % step input F
7 WBHQG1 T = t_end/N % time step
W    t(1) = 0.0 % initial element of time vector
ZKLOH 1 !Q %
'[ Q  D [ Q 7E X 7 % Difference equation loop
[ Q  '[ Q [ Q  for n = 1:N;
W Q  Q 7 Dx1 = (a11*x1(n) + a12*x2(n) + b1*u)*T;
Q Q Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
HQG x1(n+1)= Dx1 + x1(n);
SORW W[ x2(n+1)= Dx2 + x2(n);
t(n+1) = n*T;
end;
Comments %
plot(t,x1,t,x2)
It is essential to include “comments” in a computer program
to describe the algorithm and identify variables. Comments plot Statement
serve to document the program and make it easier to read.
Comments must be distinguished from the program’s code. The MATLAB command to create an x–y plot is
It seems every language uses a different symbol to identify
a comment. MATLAB uses the percent sign, %. MATLAB SORW [\
will ignore all characters which follow a % on the same line.
Comments may be included on the same line as a program where x and y are column vectors, or one-dimensional arrays
instruction, or on a separate line. Example: of equal length. Multiple plots can be created using the same
axes, by adding x, y pairs
% This is a comment.
A = B + C % This is a comment following code. SORW [\[\

Compare readability of the m-file code for the spring-mass- If you wish to plot two output variables, y1 and y2, against
damper system above, with the m-file with comments below. the same independent variable, t, then you must repeat the
Commented script for the spring-mass-damper system independent variable in each “x, y” pair
state equations.
SORW W\W\
% Second order system state equation solver
% Plotting and Labeling Multiple Figures
N = 2000 % number of steps MATLAB plots onto a “graphic object” called a “figure.”
t_end = 100.0 % duration of simulation In other words, a plot is the trace drawn on (or in) a figure.
% MATLAB’s function, plot ( x, y), draws a plot of the x, y pairs
% Parameter values of the vector x and the vector y on a figure. If a figure has not
M = 3.0 been created, then the function, plot ( x, y), creates a figure.
K = 1.0 If a figure exists, then plot ( x, y) wipes that figure clean, and
B = 0.2 draws a plot on it.
% Coefficient matrix elements
Plotting multiple figures requires creating a new figure
a11 = -B/M
with a number as its “handle” to identify it. The command is
a12 = -1/M
a21 = K
a22 = 0.0 ILJXUH KDQGOHQXPEHU 
512 8  Finite Difference Methods and MATLAB

The following MATLAB code yields only the last plot, the Figure titles are added with the function title(‘Plot Title’).
plot of x3 vs. y3, because the figure created by the first plot X-axis labels are added with the function, xlabel(‘meaningful
would be reused twice. x-axis label’). Y-axis labels are added with the command,
ylabel(‘meaningful y-axis label’). The text used in these
SORW [\ functions must be enclosed in single quotes. Example:
SORW [\
SORW [\ WLWOH
3UREOHPH6SXU*HDU6\VWHP

[ODEHO
WLPHVHFRQGV

The command figure creates a graphic object. The following \ODEHO
6KDIW7RUTXHQHZWRQV

MATLAB code would yield three new plots numbered suc-
cessively from 1, each time the code was run. Run the code Formatting Plots
four times, and a total of 12 plots would be created and stay There are various symbols, line types, and line colors avail-
open. able in MATLAB. They are selected by adding arguments
to the plot(x, y) command after the pair of vectors. See
ILJXUH Table A8.13. The default is a solid, black, thin (0.5) line. For-
SORW [\ matting the line type, color, and width is particularly helpful
ILJXUH when there are multiple plots (traces) on a figure. Line type
SORW [\ formats are coded graphically; solid is '_', dashed is '–', dot-
ILJXUH ted is ':', and dot-dash is '–.'. Black is ‘k’, blue is ‘b’, red is
SORW [\ ‘r’, and green is ‘g’. The following code draws plots of a
dashed, blue line with a width of 2. Note the use of single
There are two methods to limit the number of figures created quotes around ‘b’ and ‘LineWidth’.
by successive runs of the same code. One method is to use
the close command, prior to the figure command. The com- SORW [\


E

/LQH:LGWK

mand, close all, will close all open figures.
Color is assigned with a color letter code in single quotes.
FORVHDOO The colors are cyan, magenta, yellow, black, red, green,
ILJXUH blue, and white. The line width is specified by ‘LineWidth’
SORW [\ in single quotes and without a space between the words, fol-
ILJXUH lowed by a comma and a number indicating the width before
SORW [\ the plot command. Examples:
ILJXUH
SORW [\ figure(1)
plot(rkt,rkx1,'k','LineWidth',2)
The second method is to identify the figure with a “handle,” title('Problem 8.5.e, Worm Gear System')
which, in our case, is a number. This is done by adding the fig- xlabel('time, seconds')
ure number to the command, as an argument within parenthe- ylabel('Angular Velocity of Gear, rad/sec')
ses. The following code creates the plot of x1 vs. y1 on figure %
number 1. If figure number 1 does not exist, the command, plot figure(2)
(x1,y1), creates it. If figure number 1 does exist, then the com- plot(rkt,rkx2,':','b','LineWidth',2)
mand, plot (x1,y1), will wipe it clean and reuse it. title('Problem 8.5.e, Worm Gear System')
xlabel('time, seconds')
ylabel('Angular Velocity of Load, rad/sec')
ILJXUH  
%
SORW [\
figure(3)
ILJXUH  
plot(rkt,rkx3,'-.','r','LineWidth',2)
SORW [\
title('Problem 8.5.e, Worm Gear System')
ILJXUH  
xlabel('time, seconds')
SORW [\
ylabel('Shaft Torque, newtons')

Labeling Plots
If you are going to plot multiple figures, you must title them
to identify them. A plot’s axes should be labeled to identify
the variables and units.
MATLAB 513

Table A8.13   MATLAB’s line type, color, and marker format codes Reading From and Writing To Files
Line-type format Code Line marker format code
– Solid line (default) + Plus sign
We will need to read data from files and write data to files,
-- Dashed line o Circle
in order to use MATLAB with laboratory data. MATLAB
: Dotted line * Asterisk
conveniently has interactive methods using a graphical user
–. Dash-dot line . Point
Line Color Format x Cross
interface. However, our present focus is on programming.
Code Because Microsoft’s Excel is so widely available, it is com-
r Red ‘square’ or s Square mon for engineering software, including MATLAB and
g Green ‘diamond’ or d Diamond Mathcad, to be able to read and write numerical data to and
b Blue ^ Upward-pointing triangle from an Excel worksheet. MATLAB also includes more
c Cyan V Downward-pointing general and flexible functions for reading and writing data
triangle
to files, which we would need to use for reading and writ-
m Magenta > Right-pointing triangle
ing strings of characters or data to a “text.” Unfortunately,
y Yellow < Left-pointing triangle
complex syntax is the price of flexibility. The programmer
k Black ‘pentagram’ Five-pointed star
or p (pentagram) is responsible for “opening” and “closing” files and “for-
w White ‘hexagram’ Six-pointed star matting” the data. The latter is a throwback to the days of
or h (hexagram) programming computers using punch cards. We will stick
with Excel worksheets.
In order to use MATLAB’s syntax to read and write Mi-
A Brief History of Control Characters crosoft Excel files, we need to either assemble an array to
The syntax MATLAB uses to format data is a mixture of write to an Excel worksheet or disassemble an array read
format strings from FORTRAN, a programming language from an Excel worksheet.
which dominated engineering applications in the 1960’s and
1970’s, and control characters, which are a century older, Concatenation of Arrays
dating from 1870’s International Telegraph Alphabet, the “Concatenation” is the operation of joining arrays to create
5-bit code used to drive the first teletypes, invented by Emile a new array. It is easiest to understand the operation through
Baudot. The unit baud, the SI unit for symbols per second, is illustration. Starting with arrays A and B, where
named for Baudot.
The carriage return control characters, /r, refer to the car- a11 a12 b11 b12
A= and B=
riage of a typewriter. The first teletypes were electromechani- a21 a22 b21 b22
cal typewriters. The carriage of a typewriter is the assembly,
which carries the paper; the “platen,” a rubber-covered roller Arrays A and B can be concatenated as either
which backed up the paper; the feed rollers; and guides. The
carriage moved the paper from right to left in front of the key a11 a12
bars. The linkages connected the key of the keyboard with the a21 a22 a11 a12 b11 b12
type slug, which struck the ink-saturated cloth ribbon, and C= or D=
b11 b12 a21 a22 b21 b22
transferred the image of a character to the paper. Teletypes
b21 b22
were used for a century and evolved significantly over that
period. The last teletype model found a market as an input/
output device with the old “main frames” and “minicomput- MATLAB’s concatenation function, cat(α , β , γ ), takes three
ers.” Teletypes were used in place of computer monitors, arguments. The first argument, α, controls whether arrays β
because the teletypes were less expensive, even though they and γ are concatenated with β above and γ below, or side by
cost approximately $ 700 in 1970. To put that in perspective, side with β to the left and γ to the right. The argument, α, is
a Volkswagen cost less than $ 2,000 in 1970. Teletypes died the “dimension” along which the concatenation occurs. We
out, when the introduction of inexpensive dot matrix print- will most often work with one or two-dimensional arrays.
ers finally made them obsolete in the late 1970s. You will A one-dimensional array has only rows, xrow. Therefore, the
never work with one, but you will see them in movie scenes dimension of the element index for rows is 1. A two-dimen-
of newspaper newsrooms from the twentieth century. (Soon, sional array has rows and columns, yrow,column. The dimension
you will only see newspapers in movies too.) of the index for columns is two. The MATLAB syntax which
produced arrays C and D is:

Programming a Function in MATLAB C = cat (1,A,B) and D = cat ( 2,A,B )

See Sect. 8.3 of this chapter.


514 8  Finite Difference Methods and MATLAB

If you prefer not to work with the dimensions, MATLAB “Tools” menu to change the file settings. The complete file
also provides the functions, vertcat() and horzcat() path will have a form similar to

C = vertcat ( A,B) and D = horzcat ( A,B ) C\Documents and Settings\Class User\My Documents\
Systems\datafile.xls
The most common error in concatenating arrays is to attempt
to concatenate two-dimensional arrays which do not “fit” It is easier to read from or write to a flash drive, which will
next to each other properly. Two arrays must have the same have a short full path name. If your flash drive were assigned
number of columns if they are to be concatenated above and the drive letter R, then the full path name would be
below, or the same number of rows if they are to be concat-
enated side by side. 5?6\VWHPV?GDWDILOH[OV
By the way, when MATLAB help refers to “multidimen-
sional arrays,” it means arrays with more than two dimen- The syntax for creating Excel worksheet file and writing to
sions. it is

Extracting a Row or Column from a [OVZULWH µFRPSOHWHILOHSDWK¶DUUD\WREHZULWWHQ 


Two-Dimensional Array
MATLAB uses the colon operator “:” as shorthand for a where “array to be written” is the variable name of the array.
range variable which indexes by unity. The lower limit, j, Example: Say you plot the result of a Runge–Kutta simu-
and the upper limit, k, can be assigned as j:k. lation, and wish to save it as a Microsoft Excel worksheet
When a range variable is used as one of the indexes of file. The plot() function requires vectors holding the x data,
an array, and no limits are assigned to the range variable, in time in our case, and y data, say force F. We will concate-
other words, the colon is used by itself, then the range vari- nate these data into a two-dimensional array, name the array,
able is assigned the limits of the row or column of the array. plot_output, and write the data as an Excel worksheet file to
This shorthand notation makes it simple to extract a single a flash drive R with the following code.
column or row from an array. For example, the following
SORW W) 
code extracts the second column from the array C and as-
SORWRXWSXW KRU]FDW W) 
signs it to the variable, data2:
[OVZULWH
5?0(?UHVXOWV[OV
SORWBRXWSXW 
data2 = C(:,2)
yielding the column vector Reading and Writing to a “Text” File
Although we refer to many different “types” of files, such
a12 as a Microsoft “doc” file, a Mathcad “mcd” file, a MAT-
a22 LAB “m” file, etc., what we are referring to are the file-
data2 =
b12 name “extensions” (what follows the final dot in the file
b22 name), which allow a program to recognize a family of files
it can read and write to. There are, in fact, only two funda-
mental file types binary files, in which data are encoded
We achieve the same result using the lower limit of one and in binary and text files, in which data are encoded in char-
the upper limit of four acters. Most of the files created by the programs we work
with are text files.
GDWD &   When you read or write to an Excel file in MATLAB, you
gain the convenience of letting MATLAB handle many of
the required details for you, but you sacrifice your indepen-
Reading and Writing to a Microsoft Excel dence, in that you do things the way MATLAB wants you to.
Worksheet File Reading and writing text files in MATLAB is more involved
The MATLAB syntax for reading a Microsoft Excel work- than Excel files. You regain your freedom, and the responsi-
sheet file and assigning the data to the variable, data3, is bility that comes with it, to take care of the details.
There are three steps for writing to or reading from a text
GDWD [OVUHDG µFRPSOHWHILOHSDWK¶  file.
1. “Open” the file, meaning
Note the name of the file to be read is the complete file path, a. identify or create a file
enclosed in single quotes. If you do not see the file path in b. declare whether you plan to write to it or read from it
the colored “address bar” at the top of the dialog, use the c. assign the file to a variable name
MATLAB 515

Fig. A8.13   fprintf function, annotated


to show the association of format strings
and data
fprintf(File_Variable_Name,'%format %format new line',data 1,data 2);

2. Write (or read) line by line using a loop


format, which is a conventional decimal number, or e to
a. “format” the data by defining the width of the “field” mean “exponential,” or scientific notation. Note that f must
and the data type be lower case, but either e or E is acceptable. Example: The
b. end the line of data with a carriage return or new line number, 31,415.92, converted using both fixed and expo-
“control character” nential format strings:
3. “Close” the file
The command to open a file is I
 
)LOHB9DULDEOHB1DPH IRSHQ µILOHBQDPH¶µZ¶ 
(
 (
Note the use of SINGLE quotes around the “file_name” and
the argument “w,” where “w” means the file is opened to A common error is to use too narrow a field width. Remem-
write to. Example: ber that the field width includes the spaces between the data
also, unless those are added as a text string, or with a \t, for
tab, which are control characters.
ILG IRSHQ
/DEB5.B/LQBW[W

Z
 The formatting information ends with “control charac-
ters” which date from the 1930s and teletypes. Two useful
The command to write a line of two formatted data to an control characters, when written to a file, are \n, which com-
open file is mands a “new line” and \r, which commands a “carriage re-
turn.”
fprintf(File_Variable_Name,'format format new line',… Closing the file is the easy part. The command is
data 1,data 2);
IFORVH )LOHB9DULDEOHB1DPH 
The command fprintf() stands for File Print Formatted.
The syntax MATLAB uses to format data is literally from Forgetting to close a file means it cannot be opened and read.
the 1960’s. All of the formatting information is between two If you can access the file you created, check to see if you
single quotes. A format “string” begins with a percent sign, closed the file in your code.
%, to indicate what follows is the format code. There must The following is MATLAB code which writes the arrays,
be as many format strings as there are data. The first format rkt and rkx1, to the file, Lab3_RK_Lin_09172014.txt.
string formats the first datum. The second format string for-
mats the second datum. Note there are no commas between fid = fopen('Lab3_RK_Lin_09172014.txt','w');
format strings, just a space, while the data variables are sepa- for kk=1:N;
rated by commas. fprintf(fid,'%8.4f %12.8f \n',rkt(kk),rkx1(kk));
An example is end
fclose(fid);
ISULQWI ILG
II?Q
UNW NN UN[ NN 

The first number following the percent sign is the maxi-


mum number of numbers to be written. This is the “field MATLAB’s step() and impulse() Functions
width.” It will be padded with blanks to the left of the most
significant digit. Next, there is a decimal point followed by MATLAB’s Control System Toolbox contains the func-
a number. The number after the decimal point, called the tions, step() and impulse(), which calculate the unit step
“precision,” is the number of digits to be written after the response and the unit impulse response of a transfer func-
decimal point. The format string ends with a letter called tion. The ability to calculate a time-domain response from
a “conversion character,” which specifies the notation the a transfer function is a tremendous convenience. The step()
number is to written in. We will use either f to mean “fixed” and impulse() functions take the same arguments and work
516 8  Finite Difference Methods and MATLAB

the same way. The following discusses the function step() model = K*wn^2/(s^2+2*z*wn*s+wn^2);
but also applies to impulse(). If step() is given the name of %
a transfer function, say Model, as its single argument, tend = % Retain the previous plot with the next.
0.5, step(Model), the duration of the response may or may %
not meet our needs. We can set an end time for the calcula- hold on
tion, by adding a scalar argument, say step(Model, tend), % Calculate the unit impulse of the model at the
% times in the vector time. Assign the model
to force unit step calculation to run from t = 0 to t = tend .
% response to the variable modelResponse.
We can also pass step() a vector with the times at which to
% Assigning the output of impulse( ) to a variable is
evaluate the unit step response. In our application, the time
% necessary to format the plot in code. This is also
vector would be that of the observed response, step(Model,
% true for the functions step( ) and rlocus( ).
timeVector). %
The following MATLAB m-file program reads the RCL modelResponse = impulse(model,time);
circuit discharge data used in the Mathcad code above from %
Excel files, plots the data, then calculates and plots the im- % The formatting string is '-k'. "-" means a solid line.
pulse response of the model. % "k" stands for black. The remaining attributes
% are set as pairs of 'attribute name', attribute value.
% Program Lab 0 RLC Discharge Data and Model % Here 'LineWidth' is set to 2.
% %
% Read the time and voltage data from Excel files. plot(time,modelResponse,'-k','LineWidth',2);
% %
time = xlsread('j\timeVector090214.xls'); % The axes labels and the plot title are assigned
data = xlsread('j\dataVector090214.xls'); % after the plot is created.
% %
% Close all existing figures. xlabel('time, seconds');
% ylabel('volts, VDC');
close all; title('RLC Circuit Data vs. Model');
%
% Create a figure for the plot.
Vector Calculations in MATLAB
%
figure;
% MATLAB has two functions to determine the number of ele-
% Plot the x, y vector pair. The format string is '.b'. ments in a vector, length( ), used on vectors, and size( ),
% No line type is indicated. The period indicates which is used on arrays of any number of dimensions, in-
% that the "marker", or data point type, is a point. cluding vectors. When size( ) is used on a vector, it will re-
% Normally, the default when the line type is omitted turn the number of rows, n, and the number of columns, 1, as
% is a solid line. However, when no line type is a row vector, [n 1]. Remember that a space or a comma can
% indicated and a line marker is given, the plot will be used to separate elements in a row in MATLAB. Rows of
% have no line, just the line markers. vectors and arrays are separated by semicolons.
% The "b" stands for blue. Calculation of the integrals of the error, the absolute value
% of the error, and square of the error is performed in MATLAB
plot(time,data, '.b'); with for end loop. The MATLAB ­m-file program follows:
% Assign values for the natural frequency,
% damping ratio and the gain of the model. % Program to calculate of the integral of the error,
% % the absolute value of the error and the error
wn = 6.293E5 % squared between a model’s response and the
z = 0.258 % observed response.
K = 3.29E-6 %
% % Read Excel files with the time and amplitude data.
% Define "s" to be a transfer function in order to use %
% it as the Laplace variable. time = xlsread('j\timeVector090214.xls');
% data = xlsread('j\dataVector090214.xls');
s = tf('s') %
% % Assign values for the natural frequency,
% Assign the transfer function to the variable "model". % damping ratio and the gain
%
References and Suggested Reading 517

RLC Circuit Data vs. Model %


1.5
% MATLAB’s lowest vector or array index is 1, not 0.
% Backward- looking recursive
1
% calculations cannot start with the index 1, since
% they would look back to the non-existent index
volts, VDC

0.5 % value of 0. Calculate the first value of the error


% and integrals of the error, absolute value of the
0 % error, and of the error squared before the for – end
%loop.
-0.5 %
error(1) = modelResponse(1)-data(1);
-1
0 0.5 1 1.5 2 2.5 3 3.5 4
intError(1)= error(1)*dt;
-5 intAbsError(1)= abs(error(1))*dt;
time, seconds x 10
intSqError(1)= error(1)*error(1)*dt;
Fig. A8.14   Plot created by Program Lab 0 RLC Discharge Data and %
Model % The for – end loop uses backward-looking
% recursion to calculate the remaining values.
%
wn = 6.293E5 for i = 2 N;
z = 0.258 error(i) = modelResponse(i) - data(i);
K = 3.29E-6 intError(i) = intError(i-1) + error(i)*dt;
% intAbsError(i) = intAbsError(i-1) + abs(error(i))*dt;
% Define "s" to be a transfer function in order to use intSqError(i) = intSqError(i-1) + error(i)*error(i)*dt;
% s as the Laplace variable. end;
% %
s = tf('s') % A variable’s name followed by the Enter key is the
% % MATLAB command to evaluate and display the
% Assign the transfer function to the variable "model". % result. We can display a result on the screen
% % without format by entering the variable or elements
model = K*wn^2/(s^2+2*z*wn*s+wn^2); % name. To make our results more readable, we
% % use the fprintf( ) function write to a file, but we omit
% Calculate the unit impulse of the model at the % the File_Variable_Name aka fid. When the fid is
% times in the vector time. % omitted, MATLAB interprets the function fprintf( )
% % as a command to write to the screen.
modelResponse = impulse(model,time); %
% fprintf('\nThe integral of the error …
% Determine the number of elements in the observed = %12.6E\n',intError(N))
% response data. %
% % The syntax to continue a command onto the next
N = length(time) % line is three dots “…”
% %
% Calculate the time step dt. fprintf('The integral of the absolute value of the ...
% error = %12.6E\n',intAbsError(N))
dt = time(2) - time(1) %
% fprintf('The integral of the error squared ...
% Preallocate vectors for the variables to be = %12.6E\n',intSqError(N))
% calculated in the for – loop. The function zeros( )
% creates a vector or array of the size needed
% and fills it with zeros. Preallocation speeds
% execution of the program. References and Suggested Reading
%
error = zeros(1, N); Hamming RW (1973) Numerical Methods for Scientists and engineers,
intError = zeros(1, N); 2nd edn. Dover, New York
intAbsError = zeros(1, N); Press WH, Teukolsky SA, Vetterling WT, Flannery BP (2007) Numeri-
cal recipes, 3rd edn. Cambridge University, Cambridge
intSqError = zeros(1, N);
Transfer Functions, Block Diagrams,
and the s-Plane 9

Abstract
Transfer functions are input–output relationships in the Laplace-domain. They are multipli-
cative operators. Multiplying a transfer function by the Laplace transform of an input vari-
able yields the corresponding output variable. A block diagram represents input–output re-
lationships graphically. The operators are contained within rectangular “blocks.” The input
and output variables, or “signals,” are the lines which connect blocks. Block diagrams can
combine transfer functions representing mathematical operations, such as differentiation,
with transfer functions created from system equations, to predict the response of systems
created by interconnection. Feedback loops allow calculation of the difference between
the commanded value and the response of the system, which is termed the error signal,
and is the basis of feedback control. The s-plane is a complex plane, in which the real and
imaginary axes are the components of the eigenvalues of a system. The association between
regions of the s-plane with the homogeneous response of a system allows the s-plane to be
used as a graphical design tool.

9.1  Linear Operators and Transfer Functions displacement and force of a linear spring, where K is the
spring rate or spring constant
An operator is a function or an operation, in which one or
more input variables yield a single output variable. Opera-
F (t ) = Kx(t )
(9.1)
tors are input–output relationships. We will restrict our dis-
cussion to single input–single output operators. Hence, we
will not consider multivalued functions to be operators. By Alternatively, we can express K as a “multiplicative” opera-
our definition, the quadratic equation is not an operator, tor, by creating the ratio of the output variable, F( t) over the
because it yields two values. Neither is multiplication of input variable x( t)
two variables an operator, since multiplication is a binary
operation that requires two inputs. We will, further, restrict
ourselves to linear operators. A linear operator satisfies the  Output F (t )
= = K ≡ Operator (9.2)
test, that doubled input yields doubled output. Trigonometric Input x(t )
functions are not linear operators, since they are non-linear
functions. The next step is to generalize the familiar linear equation,
Eq. 9.1, and express it in a functional form, Eq. 9.3. Al-
though it seems awkward, we will find this form useful for
9.1.1  Linear Operators extending our definition of operators from the time-domain
to the Laplace-domain
The simplest example of a linear operator is the operation
of multiplying a variable by a constant, say, the constant K. 
Output F (t )
We are all familiar with linear input–output relationships = = K{ } ≡ Operator (9.3)
in the time-domain, such as the relationship between the Input x(t )

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_9, © Springer Science+Business Media New York 2014 519
520 9  Transfer Functions, Block Diagrams, and the s-Plane

We need three more linear operators to demonstrate the Example: where u1 (t )  = [ ft ] and u 2 (t )  = [in ]
properties of linear operators. Let us add the value of 1,000
and the units, N/m, to the operator, K N m m 
1, 000  0.305 u 1 (t ) + 0.00254 u 2 (t )
m ft in 
N N m N m
K{ } ≡ 1, 000 = 1, 000 0.305 u 1 (t ) + 1, 000 0.00254 u 2 (t )
m m ft m in

Define the operators, L , F , and I as Commutative property: L {K {u (t )}} = K {L {u (t )}}

Output Flb (t ) lb Example: where [u (t ) ] = [ m ] :


= = 0.224 = L { } ≡ Operator
Input F (t ) N
lb N N lb
0.224 1, 000 u (t ) = 1, 000 0.224 u (t ).
L {1 N} = 0.224 lb N m m N

and These linear operators possess one more important prop-


erty, that of inversion, where the inverse of an operator is
Output x(t ) m indicated by the exponent, − 1. For example, the inverse of
= = 0.305 = F { } ≡ Operator linear operator F { } is F { } . The property of inversion is
−1
Input x ft (t ) ft
defined as
F {1 ft} = 0.305 m
F { F {u (t )}} = u (t )
−1
(9.4)

Output x(t ) m Example of the property of inversion
= = 0.00254 = I { } ≡ Operator
Input xin (t ) in
1 N
K −1 {K {u (t )}} = 1, 000 u (t ) = u (t )
I {1 in} = 0.00254 m 1, 000
N m
m

9.1.2  Properties of Linear Operators The multiplicative, linear operators, F { } , I { }, K { }, and


L { }, can be manipulated algebraically, as if they were con-
Linear operators have associative, distributive, and commu- stant factors, since, in fact, each is the operation of multiply-
tative properties. To review these properties ing an input by a constant. We will state that all linear op-
erators can be manipulated, as if they were constant factors,
Associative property: A · ( B · C ) = ( A · B) · C because they all possess the four properties illustrated above.
We will not prove this statement, but we will illustrate it. To
Distributive property: A · ( B + C ) = A · B + A · C illustrate this, let us define two linear operators. Let G be the
operation of multiplying input signal u( t) by 3
Commutative property: A · B = B · A
Output
Associative property: G{ }≡ =3
Input

{ {
L F ( K {u (t )}) }} = L {(F {K {u(t )}})} Let H be the operation of differentiating the input signal with
respect to time
Example: where [u (t ) ] = [ m ] :
d
H{ }≡
dt
lb m N 
0.224 0.305  1, 000 u (t ) Then
N ft  m 
d
lb  m N  H {G {u (t )}} = (3u (t )) = y (t )
= 0.224  0.305 1, 000 u (t ) dt
N ft m 
and
Distributive property: du (t )
G { H {u (t )}} = 3 = y (t )
dt

{ { } {
K F u1 (t ) + I u 2 (t ) }} = K {F {u (t )}} + K {I {u (t )}}
1 2
9.1  Linear Operators and Transfer Functions 521

 d 2u  du
These are equivalent statements since
L
b0  2  + b1
 dt 
L  dt  + b L {u}
2
d du (t )
(3u (t )) = 3 d 2y dy
dt dt
= a0 L  dt 2 

+ a1 L  dt  + a L { y}
2

9.1.3  Incrementally Linear Functions


b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s ) = a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s ) .
The operation of multiplying the input by a constant is clear-
ly a linear operator. Although straight lines are described by 2. Factor out the transformed input and output variables
linear functions, only straight lines which pass through the
origin pass our test for linear operators. For example, the fa- (b s
0
2
) ( )
+ b1 s + b2 U ( s ) = a 0 s 2 + a1 s + a 2 Y ( s )
miliar equation of a straight line
3. Create a ratio of the output variable over the input vari-
m x(t ) + b = y (t ) able

Output ( s ) Y ( s ) b0 s + b1 s + b2
2
is not a linear operator, since doubling x( t) does not double = = (9.5)
y( t)  Input ( s ) U ( s ) a 0 s 2 + a1 s + a 2

m 2 x(t ) + b ≠ 2 y (t ) A transfer function is a linear operator in the Laplace-


domain. The numerator polynomial contains the opera-
This equation of a straight line also does not yield an opera- tions performed on the input, transformed into the Laplace-
tor which is the ratio of output y( t) over input x(t) domain, where the Laplace variable, s, assumes the role of
differentiation, with respect to time in the time-domain. Cor-
Output y (t ) mu (t ) + b b respondingly, the denominator polynomial, a 0 s 2 + a1 s + a 2,
= = = m+ ≠ Operator
Input x(t ) x(t ) x(t ) contains the operations performed on the output variable in
the differential equation. Note it is also the characteristic
A straight line which does not pass through the origin is re- function of the differential equation. Setting the characteris-
ferred to as “incrementally” linear, Sect. 4.2.4.1. tic function equal to zero forms the characteristic equation of
the differential equation

9.1.4 Differential Equations and Transfer a 0 s 2 + a1 s + a 2 = 0 (9.6)
Functions as Linear Operators
A transfer function is an unusual function, because it is used
The differential system equation is an input–output relation- in two very different ways. A transfer function can be evalu-
ship. It is a dynamic operator which operates on the input ated as a conventional function by assigning a value to its
to yield the output, where output is the response function of argument, s = σ + jw . Transfer functions are complex func-
the dynamic system. A differential system equation is also tions, meaning, in the general case, the argument, s, and the
a linear operator, if each of its terms passes the “double the output are complex numbers.
input yields double the output” test. Linear differential sys- The second manner of using a transfer function is as a
tem equations result from modeling the energetic properties dynamic operator. Rather than substituting the input func-
of a real object or system, using linear elemental equations. tion into the time-domain differential equation, the Laplace-
Recall that we can create a transfer function from a dif- domain transfer function acts as an input–output relation-
ferential system equation, as follows: ship, when multiplied by the Laplace transformation of the
1. Use the Laplace transform, neglecting the initial condi- input function. A transfer function is a multiplicative opera-
tion terms, to create an algebraic equation tor in the Laplace-domain. Multiplying a transfer function by
the Laplace transform of input function yields the Laplace
d 2u du d 2y dy transform of the output (response) function.
b0 + b + b u = a + a1 + a2 y
dt 2 1
dt
2 0
dt 2
dt Refer to the relationship between the displacement and
force of a spring, F (t ) = Kx(t ) . If we rearrange this as a ratio
 d 2u du   d 2y dy  of output F( t) over input x( t), we have the form of a transfer
L b 0 2 + b1 + b 2u  = L a 0 2 + a 1 + a 2 y function
 dt dt   dt dt 
F (t )
F (t ) = Kx (t ) → =K
x (t )
522 9  Transfer Functions, Block Diagrams, and the s-Plane

A transfer function is a linear, multiplicative operator. For where C is the “output” matrix, and D is sometimes called
example, if the input is x(t ) = 3 sin(2t ), then the output is the the “feed-through” matrix. The output equations are derived
product from the equation list, presuming knowledge of the values of
the state variables over the duration of interest, and assuming
F (t )
x (t ) = K 3sin ( 2t ) → F (t ) = K 3sin ( 2t ) knowledge of the input, which is necessary for any solution.
x (t ) We will solve the two matrix equations

If we perform the Laplace transformation on the time-domain dx


= Ax + Bu
equation, where the unknown functions in the time-domain, dt
x( t) and F( t), are transformed into the unknown functions,
X( s) and F( s), in the Laplace-domain, and then create the y = Cx + Du
ratio of the output function over the input function, we have
a Laplace-domain transfer function: by first solving the state equations, so that state vector x is
expressed in terms of the constant matrices, A and B, and the
L {F (t )} = L {Kx (t )} → F ( s ) = KX ( s ) input vector, u. We will then substitute the solution for x into
the output equations, yielding output vector y as a function
F (s) of input vector u.
=K
X (s) There are only two steps in this solution, which require
mathematical insight or, for most of us, education. The first is
The advantage of a transfer function is its ability to serve as a the fundamental approach to the solution. We will transform
multiplicative constant, such as spring constant K. It can also the set of simultaneous first-order differential equations into
serve as a dynamic operator, meaning it can be formed from a set of simultaneous algebraic equations, by operating on
the Laplace transformation of a differential system equation. both sides of the state equations with the Laplace transform
The time dependencies represented in the differential system
dx
equation, as derivatives with respect to time, are represented
in the transfer function as powers of the Laplace variable s.
L  dt  = L {Ax + Bu}
L {y} = L {Cx + Du}
9.2 Laplace-Domain Solution of a Set of State
and Output Equations The Laplace transform is a linear operator and, as such, can
be manipulated as if it were a multiplicative constant. It can
The final topic in our development of matrix mathematics is be distributed to each term in a vector or matrix, in the same
use of the Laplace transformation to solve a set of state equa- manner that it is distributed to each term in a summation
tions and output equations expressed in matrix form. The re-
dx
sulting solution is a vector of transfer functions of each of the
output variables. The solution eliminates the state variables,
L  dt  = L {Ax} + L {Bu}
unless they are included in the output vector. We begin with
state equations in standard vector-matrix notation L {y} = L {Cx} + L {Du}
dx Matrices A, B, C, and D can be factored out of the Laplace
= Ax + Bu transforms, since they are constants
dt
dx
where x is the vector of state variables; A is the “coefficient”
matrix comprised of terms consisting of the elemental pa-
L  dt  = AL {x} + BL {u}
rameters; B is the “input” matrix, again comprised of terms
consisting of the elemental parameters; and u is either a sca- L {y} = CL {x} + DL {u}
lar, in the case of a single input, or a vector in the case of
multiple inputs. We will keep u as a vector during the deriva- Recall the Laplace transformation of a derivative with re-
tion of the solution, then evaluate the solution, using a single spect to time yields the Laplace variable, s, in the Laplace-
(scalar) input for u. domain. Mathematical convention is to use capital letters for
The output equations in matrix form are the transformed variables

y = Cx + Du sX( s ) = AX( s ) + BU( s )


Y( s ) = CX( s ) + DU( s )
9.2  Laplace-Domain Solution of a Set of State and Output Equations 523

Considering the transformed state equations, subtracting Y( s ) = CX( s ) + DU( s )


AX( s) from both sides yields
We can now substitute for the transformed state vector X(s)
sX( s ) − AX( s ) = BU( s ) in the output equation

Y( s ) = C ( sI − A ) BU( s ) + DU( s )
−1
This is where the second mathematical insight, or education,
is required. We would like to factor transformed state vari-
able X( s) out of the two terms on the left side of the equation, Factoring out transformed input vector U( s), we obtain the
but we cannot, because the result desired result

( s − A ) X( s) = BU( s)

(
Y ( s ) = C ( sI − A ) B + D U ( s )
−1
) (9.8)

contains the factor the transformed output vector, Y( s), as a function of the
transformed input vector, U( s).
(s − A) The crux of the calculation is the inverse matrix ( sI − A) −1 .
We will investigate the meaning of this expression by re-
which is the difference between the scalar variable s and the turning to the RLC circuit, Fig. 1.39, which we reduced in
matrix A. That operation is not defined. This term cannot be Sect. 1.5.2.2 to the second-order system equation relating the
evaluated. The mathematical insight is that we may pre- and input, v1g, to the output, v3g,
post-multiply any vector or matrix we wish by the identity
matrix, I, without changing the equation. If we pre-multiply d 2 v3 g dv3 g
v1g = LC + RC + v3 g
the transformed state vector, X( s), in the first term on the dt 2
dt
left side
and to a set of state equations, Sect. 7.6.3
sIX( s ) − AX( s ) = BU( s )
 R 1
− −  i 1
We have changed the equation no more than if we were to dx d  iL   L L  L   
= Ax + Bu → v  =   + L v
have multiplied a term in a scalar equation by the scalar iden- dt dt  3g   1   v3 g   
tity, the number one. However, we are now able to factor out  C
0
  0 
the transformed state vector, X( s)
and output equations, Sect. 7.6.4
( sI − A ) X( s) = BU( s)
(9.7)
 iC   1 0 0
since the factor i   1 0 0
 R   iL   
( sI − A ) y = Cx + Du →  i = 1 0    + 0 v
    v3 g  
 v12   R 0    0
is the difference between two square matrices of the same
v23   − R −1 1 
size. We solve for X( s) by multiplying both sides by the in-
verse of the factor, ( sI − A),
Recall the inverse of a matrix M is
( sI − A )−1 ( sI − A ) X( s) = ( sI − A )−1 BU( s)
cof ( M )
T

M −1 =
yielding M

X( s ) = ( sI − A ) BU( s )
−1
where cof (M ) is the matrix of the cofactors of M and |M| is
the determinant of M.
Returning to the set of state equations and output equations Expressing ( sI − A) −1 in terms of the determinant and co-
factor of ( sI − A)
X( s ) = ( sI − A ) BU( s )
−1

 R 1  R 1  R 1
− −  − −  s +
1 0  L L  s 0  L L L L
sI − A = s  − = − = 
0 1  1 0 s   1 1
0   0  − s
 C   C   C 
524 9  Transfer Functions, Block Diagrams, and the s-Plane

T coefficients of the exponents, which form the homogeneous


 R 1
s + L L  solution
cof  
−1 s v3 g H (t ) = A1e s1t + A2 e s2t
(9.10)
( sI − A ) −1
=
[ sI − A ]
T

=
 C
 
sI − A R 1 The homogeneous solution represents the natural response
s+
L L of the system to a disturbance which transfers energy to the
1 system.
− s
C The transfer function, Eq. 9.10, formed from the Laplace
transform of the differential system, Eq. 9.8, contains the op-
We will begin by evaluating the determinant erators from the output variable side of the differential equa-
tion as the denominator polynomial
R 1 1
s+
L L Output ( s ) V3 g ( s ) LC
sI − A = = = (9.11)
1 Input ( s ) V1g ( s ) R 1
− s  s2 + s +
C L LC

 R  1  1 R 1 Hence, the denominator of a transfer function is the charac-


sI − A =  s +  s −    −  = s 2 + s + teristic function of the system.
 L  L  C  L LC
Although the state-space representation of a dynamic system
is quite different from the same system represented as a higher-
This expression is a second-order polynomial in s. It is, in order differential equation, both mathematical representations
fact, the characteristic function of the system, which we can must contain the same physical information. From a mathe-
verify by comparing to the characteristic function of the matical perspective, “invariant” quantities do not change, when
RLC circuit we derive from the second-order system equa- we manipulate our mathematical representations. In dynamic
tion. Taking the Laplace transform of both side and neglect- systems, the eigenvalues, or roots of the characteristic equa-
ing the initial condition terms, tion, and the corresponding eigenvectors, the set of orthogonal
vectors which correspond to the eigenvalues, do not change,
d 2 v3 g dv3 g when we manipulate the mathematical representation, because
v1g = LC 2
+ RC + v3 g they have physical meaning. They are, respectively, the decay
dt dt rate and initial magnitude of the natural response of a system.
 d 2 v3 g dv3 g  Turning to the calculation of the cofactor of ( sI − A) , re-
L {v } = L  LC
1g
dt 2
+ RC
dt
+ v3 g 

call the cofactor of a matrix is created by forming the minor
 or submatrix, by deleting the row, sI − A, and column of an
element, calculating the determinant of the minor, and sub-
 d 2 v3 g   dv3 g  stituting it in place of the element, Sect. 7.5. Calculation of
L {v } = L
1g  LC
 dt 2 
+ L  RC dt 
+ L {v }3g
the cofactor of a two by two matrix requires use of the spe-
cial case of the determinant of a single element, since there
is only one element remaining in the matrix, when we drop
V1g ( s ) = LC s 2V3 g ( s ) + RC sV3 g ( s ) + V3 g ( s )
the row and column containing our target. Recall the deter-
minant of a single element is defined as that element. Also,
V1g ( s ) = ( LC s 2 + RC s + 1)V3 g ( s ) recall the elements of the cofactor of a matrix are multiplied
by ( −1)( r + c ) = ( −1)(i + j ) . Hence,

1  R 1   R 1
V1g ( s ) =  s 2 + s +
(9.9)   V3 g ( s ) s + L L
LC L LC  cof ( sI − A ) = cof  
−1 s
What does it mean to have the characteristic function of the  C 
system in the denominator of ( sI − A) −1? Recall from the
Method of Undetermined Coefficients, Sect. 2.5, when the
 1 
characteristic function of a system is set equal to zero, it is  ( −1)
(1+1)
s ( −1)(1+ 2)  −  
the system’s characteristic equation. The roots of the charac- C 
cof ( sI − A ) = 
 ( 2 +1)  1  R 
teristic equation are the eigenvalues of the system. They are
( −1)   ( −1)(2+ 2)  s +  
 L L 
9.2  Laplace-Domain Solution of a Set of State and Output Equations 525

Transposing the cofactors of ( sI − A) yields after we have taken the Laplace transform of output vector
y( t) for the RLC circuit,
T
 1   1 
 s C  s −
L 
  iC    I C (s) 
cof ( sI − A )  
T
=  =  .  I (s) 
− 1 R 1 R   iR    R 
s+  s+
 L L  C L  L {y } = L  i  → Y( s) =  I ( s) 
v   
  12   V12 ( s ) 
Division by the determinant sI − A yields the inverse ma-  v23   V23 ( s ) 
trix
and its scalar input, u( t),
T
 R 1
s + L L  L {u} = L {v} → U (s) = V (s)
cof  
−1 s
[ sI − A ] =  C
T

( sI − A )−1 =
R 1
 
s − 
1  
sI − A   L 
s+  IC (s)    1 0   0
L L  I (s)    1  1 R 0
   0  C s + 1  
1 R
L 
− s  I ( s )  =   1 0   L  +  0  V ( s )
C     2 R 1    
V12 ( s )    R 0  s + s+  0   0 
L LC
 1  
T
1  V23 ( s )    − R −1 1 
 
 s C  s − 
L  
   
− 1 s + R   1 s + R

= L L  =  C L 
( sI − A )−1 R 1 R 1
It will be easier to evaluate the right side, if we factor the
2
s + s+ 2
s + s+ characteristic function, which is a scalar function, out of its
L LC L LC matrix product. We will divide the result of the matrix mul-
tiplication by the characteristic function as the last step. Fac-
We can complete the solution by substitution toring the scalar function to the front of its term

(
Y ( s ) = C ( sI − A ) B + D U ( s )
−1
) (9.12)

 IC (s)    1 0  0 
 I (s)    1 0   s 1   
 R     −   1   0 
1 L    
 I (s)  =   1 0   L + 0  V (s)
   s2 + R s + 1  1 R    
V12 ( s )   R 0  s +   0  0
L LC  C L  
V23 ( s )   
   − R −1 1 

We now evaluate the matrix multiplication

  1  
  s −  
L
   
 IC ( s)    1 
s − 0 
 I ( s)    L   
 R      1  0
1  1   L  + 0 V ( s )
 I ( s)  =  s −
   s2 + R s + 1  L     
V12 ( s )   L LC  R   0  0
V23 ( s )   Rs −  1 
  L  
  1 R  R  
 − Rs − −s+  
  C L  
L  
526 9  Transfer Functions, Block Diagrams, and the s-Plane

  1   The final result is a vector of transfer functions for the output


  s   variables of the system.
L
    When there is the need to derive transfer functions of
 IC (s)    1  0 
s higher-order systems, the vector-matrix form of the re-
 I ( s)    L   
 R     0  sult,Y( s ) = (C( sI − A) −1 B + D)U( s ), Eq. 9.12, is a tremen-
1  1 
 I ( s)  =  s + 0 V ( s ) dous time-saving method. As a rule of thumb, the time re-
   s2 + R s + 1  L    quired to reduce the energetic equations of a system to a
V12 ( s)   L LC  R  0
V23 ( s)    s  1   single, higher-order system equation is roughly proportional
 
  L    to the square of the number of equations. In contrast, the
  R 1   time required to reduce the same set of energetic equations
 − s −  
 L LC  to state equations is roughly proportional to the number of
equations.
Modern handheld engineering calculators, i.e., the TI-89
 s  
and TI-Nspire, can invert a four by four matrix with sym-
  2 R 1  
Ls + s +    bolic elements. The symbolic algebra capability of these ma-
  L LC    chines is superior to that of Mathcad and MATLAB. What
 s   formerly were onerous linear or scalar algebraic expansions
  
  L  s2 + R s + 1    and simplifications can now be performed quickly and ac-
I ( s )    L LC
 
0  curately with these calculator.
 C     
 I (s)    s  0
 R     
 I ( s )  =   L  s 2 + R s + 1   + 0 V ( s ) 9.3  Block Diagrams
    L LC    
V12 ( s )    Rs
0 
  
V23 ( s )     1  Thus far, we have developed two mathematical descriptions
  R 1     of the dynamic relationships present in energetic systems: (1)
 Ls + s +2
 
  L LC    classical, higher-order differential equations, which repre-
 R 1
  sent the dynamic relationship between a single input variable
 − s−   and a single output variable, and (2) the state-space method,
 L LC  
  2 R 1    using a set of simultaneous first-order differential equations,
   s + L s + LC    expressed in vector-matrix notation and written in terms of
 

 s   s 
  R 1     R 1  
 L  s2 + s +    L  s2 + s +  
  L LC     L LC  
 s   IC (s)   s 
   V ( s)   
  2 R 1      L  s2 + R s + 1  
 L  s + L s + LC    I R ( s )    L

LC  
 IC (s)       
 I ( s)   s   V ( s)   s 
 R     I ( s)    2 R 1  
 I ( s )  =  L  s 2 + R s + 1   V ( s ) →   =  Ls + s +  
    L LC    V ( s)    L LC  
V12 ( s )     V12 ( s )   
V23 ( s )   Rs Rs
      
  2 R 1    V ( s)   L  s 2 + R s + 1  
 L  s + L s + LC    V ( s )    L

LC  
   23   
 R 1   V ( s )   − R s − 1 
 − L s − LC   L LC 
 + 1  + 1
  s2 + R s + 1     s2 + R s + 1  
  L

LC     L

LC  
9.3  Block Diagrams 527

Fig. 9.1  a Block with a linear


operator. The input and output
a b
signals conform to the direction U(s)
Input Linear Output Y(s)
of the arrowheads. b Block with G(s)
transfer function G( s), input U( s) Operator
and output Y( s)

the energy storage, or state variables, and an input vector. G {U ( s )} = Y ( s )


Linear operators permit both higher-order differential equa-
tions and sets of simultaneous differential equations to be The output signal of the block can be labeled as a variable
represented graphically as “block diagrams.” Block diagram (signal), Y( s), as the result of operation on the input, G {U ( s )},
representation of differential equations is advantageous, be- or, with both terms, as an equation, G {U ( s )} = Y ( s ). The
cause it yields a deeper understanding of the symbolic math- only use of an equal sign on a block diagram is to indicate a
ematics. Block diagrams have the additional importance as a signal has two names.
means of designing feedback control systems. Because linear operators can be manipulated algebra-
A block diagram is a method of representing information ically, as if they were constant factors, we will simplify and
flow, and the sequence of operations performed on informa- clarify our notation by omitting the braces, “{},” which we
tion. The term, “block,” refers to the rectangles used to box- used in Sect. 9.1.1. We may indicate a linear operator is op-
in the operators. Block diagrams are used for some engineer- erating on a quantity, by using the same notation as one does
ing software graphical user interfaces. The drag-and-drop for multiplication. For example, the distributive property ex-
diagrams you may have created using the data acquisition pressed in Sect. 9.1.1 as
software LabView are block diagrams.
The lines between the blocks in a block diagram represent
the variables in the system. These variables are also called { { } { }} = K {F {u (t )}} + K {I {u (t )}}
K F u 1 (t ) + I u 2 (t ) 1 2

“signals,” because the block diagram notation was developed


by electrical engineers. The lines representing signals must will henceforth be expressed as
include arrows to indicate the direction of information flow,
i.e., which is input and which is output of the block. With the
exception of the operation of summation, all linear operators ( )
K FU 1 ( s ) + I U 2 ( s ) = KFU 1 ( s ) + KI U 2 ( s )
have a single input and a single output. Do not show more
than one signal entering a block, except for a “summation We may also omit the independent variable, s, and rely on
junction.” Never show more than one signal exiting a block. upper case font to identify variables in the Laplace-domain.

K ( FU1 + I U 2 ) = KFU1 + KI U 2
9.3.1  A Block
For convenience, we will apply the associative, distributive,
The simplest block diagram contains a single block with and commutative properties to combinations of signals and
one line entering the block, the input signal, labeled U( s), operators. In other words, the variable (signal) need not be
and one line exiting the block, the output signal, labeled Y. on the right end of a term
The operator in the block is identified by the letter, G( s),
but does not need to be a function of the Laplace variable, LKI U = LKU I = LU KI = U LKI
s. G( s) can be a constant. Since transfer functions are linear
operators, they can be used in block diagrams. Depending When we interpret expressions, we must distinguish between
on circumstances, we may choose to name the transfer func- variables (signals) and operators. Operators operate on vari-
tion, Eq. 9.13, G( s); express it as Y( s)/U( s), or as a ratio of ables (signals), not the other way around.
polynomials in s.

Y (s) b0 s 2 + b1 s + b2 9.3.2  Cascaded Blocks


(9.13)
≡ G (s) =
U (s) a 0 s 2 + a1 s + a 2 Two operators in series are referred to as “cascaded.” Using
U( s) as the input signal, W( s) as an intermediate signal, Y( s)
The functional relationship represented by the block diagram as the output signal, and G( s) and H( s) as the two operators,
Fig. 9.1b is expressed symbolically as the cascade is shown in Fig. 9.2.
528 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.2   Cascaded blocks G( s)


and H( s) U(s) W(s) Y(s)
G(s) H(s)
Signal Signal Signal
Linear Linear
Operator Operator

Fig. 9.3   Four equivalent block


diagrams
a c
U Y U Y
G H GH

b d
U Y U Y
H G GH

Fig. 9.4  a Integration block with


Laplace-domain notation in the
a dv3g
b
dv3g
time-domain. b Integration block — v3g — v3g
with time-domain notation dt 1
__ dt ó
dt
s
ó

The functional relationship of the cascade is represented low output impedance can provide a large amount of power
symbolically as to its downstream neighbor. Often, there is an impedance
mismatch in a system design, and the cascade model cannot
UG = W and WH = Y (9.14) be used. In these cases, the dynamic model between the input

and the output variables must be represented as a whole by a
The commutative, associative, and distributive algebraic single transfer function.
properties of linear operators extend to their representation
in block diagrams. The order of cascaded blocks can be in-
terchanged, or the block diagram simplified, by combining 9.3.3 Differentiation, Integration, and Transfer
the cascaded blocks. Block diagrams are equivalent, if an Functions Blocks
input signal yields the same output signal. The four block
diagrams shown in Fig. 9.3 are equivalent. The two cascaded The example above illustrated how differentiation with re-
blocks simplify to the product of the two linear operators, spect to time is a linear operation. Integration and differen-
G( s) and H( s), as given in Eq. 9.15: tiation are inverse operations, since
 UGH = Y (9.15) du
∫ dt dt = u and
d
dt
(∫ udt ) = u
The term “cascade” derives from a natural cascade of water.
If one were to affect the water at a point in a cascade, say, There are different notations used to represent the operations
by disturbing the bottom and muddying the water, the effect of differentiation and integration in time-domain (where the
of that action would be seen downstream but not upstream. signals are variables of time) block diagrams. Rather than
The same assumption is made, when blocks are cascaded d
( ) and ∫ ( ) dt , some authors use D to represent differen-
in a block diagram, that actions downstream do not affect the dt
values of signals upstream. Whether or not this is a reason- tiation, and D−1 to represent the inverse operation of integra-
able model in a dynamic system depends on the input and 1
tion. We will use s to represent differentiation, and s−1 or
output “impedance” of the components or subsystems which s
comprise the blocks. Impedance is a generalized resistance, to represent integration in the time-domain, Fig. 9.4. They
addressed in Chap. 11. A block with high input impedance are concise and their use in the time-domain will reinforce
draws little power from its upstream neighbor. A block with their meaning in Laplace-domain block diagrams.
9.3  Block Diagrams 529

9.3.4  Summation Junctions a Input A

Summation junctions are represented as circles, rather than Output


Input B + A+B-C
rectangles. A summation junction usually has multiple in- +
-
puts, but a single input is allowable. There is only one out-
put. A positive or negative sign is shown where the input Input C
enters the summation junction. This allows the summation
junction to add orsubtract an input from the resulting out-
put. Engineering terminology here is somewhat awkward. b A
+
A+B
+
A+B-C
+ -
The signal is described as “inverted” or “non-inverted.”
The advantage of this nomenclature is the reduction of po-
tential sign errors, when a negative signal is inverted. Input B C
signals are indicated as inverted or non-inverted, upon their
entry to the summation junction. If the output signal must
be inverted, it is done with an “inverter” block (a block with c A A-C A-C+B
+ +
negative one as its operator), after the output signal leaves - +
the summation junction. Cascaded summation junctions are
interchangeable. C B

9.3.5  Branch Points Fig. 9.5   Equivalent block diagrams. a Single summation junction with
three inputs. b Two summation junctions. c Summation junctions of b,
interchanged
Often it is necessary to send the output signal from one block
to two or more blocks. A block has only one output signal.
However, a signal can be “branched” or “tapped,” in order 9.3.6  Block Diagram Algebra
to provide additional copies. The signal does not divide at
a branch point, because what is transmitted is information. A set of mathematical rules is an “algebra.” Block diagram
The same information is sent down to both branches. The algebra are the rules which permit manipulation of a block
graphical convention is to indicate a branch point with a dot, diagram into an equivalent block diagram. We have already
in order to distinguish it from signal lines which simply cross introduced three rules:
each other without a connection, Figs. 9.5 and 9.6. These 1. Cascaded blocks can be interchanged or combined into a
conventions are identical to those used in circuit diagrams. single block.
There is an important physical difference between branching 2. A summation junction with three or more inputs can be
in a block diagram, and in an electric circuit. The quantity divided into cascaded summation junctions.
which flows in a block diagram is information; whereas, the 3. Signals can be branched to replicate signals.
flow quantities in an electric circuit are current and power. We need two more rules to have a working knowledge of
In the abstract block diagram, there is no limit to the number block diagram algebra: how to move blocks and branch
of branches, which can be added to an ideal signal. An ideal points and how to collapse two different types of loops.
signal never degrades, as it is divided. 4. A block can be moved, relative to a branch point, if an
A real signal in an electric circuit has a limit, called the equivalent block diagram is created by adding or remov-
“fan-out,” which is the number of branches that can be added, ing blocks, as necessary.
before the information of the signal degrades due to insuffi- Rule four is best explained through two examples.
cient power. In a digital circuit, a typical fan-out is eight.

Fig. 9.6  a One signal with


a branch point. b Two crossing
a Signal B
b Signal A
signals Dot “ties” signals
together

Signal A Signal A Signal A Signal A

Signal B Signal A
530 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.7   Two equivalent block


diagrams. a Block diagram with
a AG
b AG
two blocks G after branch point, G
b Block diagram with a single
block G before branch point
A AG A AG
G G

Fig. 9.8   Two equivalent block


a b 1
diagrams. a Block diagram with A __ A
block G after the branch point,
b Block diagram with a block G
G before branch point and the
inverse block, 1/G, on the branch
A AG A AG
G G

Fig. 9.9   Feedforward loops


a b
Information Flow Information Flow
A AG AG+AH AG A A
G ++
G

A A AH ++
AG+AH AH A
H H
Information Flow Information Flow

Fig. 9.10   Two equivalent block


diagrams. a A feedforward loop.
a b
b The feedforward loop replaced A AG
G
by a single transfer function

A A AH ++
AG+AH A AG+AH
H G+H

Example 1  Move two identical blocks “back” (against the block diagram are relative to the input and output signals.
direction of information flow) through a branch point. The The forward direction is toward the output of the block di-
blocks merge (one is removed), Fig. 9.7. agram. The backward direction is toward the input to the
block diagram.
Example 2  Move a branch point forward through the block Unfortunately, in the case of feedforward loops, liberty
G. When a branch point is moved forward through a block is taken with the meaning of “forward.” Two “feedforward”
G, the inverse block, 1/G, must be added to the branch to loops are shown in Fig. 9.9. In Fig. 9.9a, both signals of the
remove the operation of block G on the signal, Fig. 9.8. feedforward loop are in the forward direction. In Fig. 9.9b,
both signals of the loop are in the backward direction. What
makes both loops feedforward loops is that two signals of a
9.3.7  Feedforward and Feedback Loops loop are in the same direction.
They can both be replaced by a single block with the same
We will state block diagram Rule Five and then develop the transfer function, Fig. 9.10. Replacing a loop with a single
background needed to understand it and apply it. transfer function is known as “collapsing” the loop.
5. Loops must be untangled before they are collapsed from, A feedforward loop can be collapsed, i.e., replaced with
the inner most outward, using the feedback and feedfor- a single block, by inspection of the output of the summa-
ward loop reduction formulae. tion junction in Fig. 9.10a. Operating on input signal A with
Branch points and summation junctions allow block dia- the transfer function (the sum of the transfer functions of the
grams to include “loops,” where a signal is branched, op- two branches of the feedforward loop, G + H) yields the same
erated on, and then summed. “Forward” and “back” in a output. Hence, the two block diagrams are equivalent.
9.3  Block Diagrams 531

R C puts voltage. Why? The inputs to a summation junction must


+
- G have the same units, or they cannot be summed. The out-
put of the summation junction, R − CH, would be a voltage
H which represents the error or difference between the desired
speed and the actual speed.
The transfer function, which is equivalent to a negative
Fig. 9.11   Negative feedback loop. Input R stands for the reference, and
output C stands for the controlled variable feedback loop, cannot be determined by inspection of the
block diagram. The key to understanding block diagrams is
Input R Error E Output C to identify the input and output signals of each block in the
+
- G diagram. There are three types of equations which can be
written to represent the input–output relationship of a block.
In many cases, the output of a block is an internal signal
H
Feedback CH within the system, which, though essential for the operation
of the closed-loop system, is not significant enough to be
Fig. 9.12   Signals identified in a negative feedback loop named, Fig. 9.13a. The equation which represents that input–
output relationship is trivial, AF = AF. In some cases, the
output signal of a block has a name, such as the output of
A “feedback” loop is shown in Fig.  9.11. This block system C, Fig. 9.13b. This yields the non-trivial equation,
diagram is the basis of single input–single output feedback AF = C. The third case is an internal signal significant enough
control systems. The standard notation uses R for the input to be named. Figure 9.13c shows the summation junction of
signal, and C for the output signal, where R stands for Refer- a closed-loop negative feedback system. The output of the
ence and C for the Controlled variable. The feedforward path summation junction is named error E. This yields the non-
is from the summation junction, through transfer function G, trivial equation, R − CH = E.
to the output. The feedback path is from the output, through To derive the single transfer function which can replace
transfer function H to the summation junction. The feedfor- or collapsed in a feedback loop, we need to identify the sig-
ward transfer function, G, represents the system under au- nals in the block diagram, write symbolic equations which
tomatic control, and includes the controller. The feedback represent the input–output relationships of the blocks, and
transfer function, H, represents the sensor used to measure do a little algebra. We will use Fig. 9.12 where the output of
the output variable, C, and whatever signal processing is per- the summation junction is named E for Error. There are two
formed on it, before it is subtracted from the input signal at blocks and the summation junction, so we can write three
the summation junction. Note the feedback signal is inverted equations. The equation for the summation is
at the summation junction, creating negative feedback. The
output of the summation junction is the difference between R − CH = E
input R and feedback CH.The sensed output variable is error
signal E, Fig. 9.12. We can also write an equation for the controlled variable, C,
Suppose you wish to control the speed of a motor, using in terms of E
closed-loop negative feedback control. Controlled variable
C would be the angular velocity of the motor. Input refer- EG = C
ence signal R could be a number of physical quantities, such
as the position of a throttle, the torque or force on a machine The remaining equation is trivial, since the output signal is
element, a current, or a voltage. If reference signal R were named as the product of the input signal and the block, but
a voltage, then block H represents a tachometer which out- we will write it, to be comprehensive,

Fig. 9.13   Signal names in block


diagrams. a An internal signal
a b c
with no name. b An existing Output Input Error
name, in this case, the output
A AF A C=AF R E=R-CH
signal of system C. c The name F F +
-
error assigned to the output of the
summation junction

CH
Feedback
532 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.14   Two Equivalent block


diagrams. a Block diagram of a
a b
closed-loop negative feedback R C R G
_____ C
system. b Equivalent transfer +
- G
function of the closed-loop 1+GH
system
H

Control Control
Input Error Signal Output Input Error Signal Output
R E U C R E U C
+
- GController G Plant +
- GController G Plant

HSensor HSensor
Feedback Feedback
CH CH
Open loop here
Fig. 9.15   Closed-loop negative control system with the feedforward
transfer function split into GController and GPlant. Control signal U( s) is
usually amplified to power the plant Fig. 9.16   Open-loop system. The feedback loop is broken before the
summation junction. The error signal, E, equals input R

CH = CH increases, the feedback signal, CH, grows, and error E de-


creases.
These three equations can be reduced to yield an input–out- An absolute requirement for closed-loop feedback control
put relationship for the input and output signals, R and C, of shown in Fig. 9.14 and Eq. 9.16 is that a positive input to the
the system by eliminating the signal, E: closed-loop system, R, produces a positive output, C. A sign
inversion in the forward path between the summation junc-
C tion and the output leads to positive feedback.
EG = C → E =
G
R − ( −CH ) = R + CH ≠ E
C
R − CH = E → R − CH =
G
Positive feedback is good for people but destabilizing for
physical systems. Positive feedback causes the error signal
RG − CGH = C → RG = C + CGH → RG = C (1 + GH )
to grow when, in fact, the system is acting to reduce the error.
The system is out of control.
C G Recall the sign inversion introduced by the linear graph
GC .L. ≡ = (9.16)
 R 1 + GH transformer and transducer sign convention. It was stated
that the signs would be “fixed” after the reduction of the
Equation  9.16 is known as the “closed-loop transfer func- linear graph to a system equation, if necessary. Use of a
tion” since it replaces the entire feedback loop. transfer function derived by the linear graph method in a
The block diagram of Fig. 9.14a is the simplest and most closed-loop feedback control system is such a situation.
condensed block diagram of a closed-loop negative feedback There must not be a net sign inversion in the path from the
system. Transfer function G represents all subsystems in the summation junction back to the summation, the open-loop
forward path combined. A slightly expanded block diagram signal shown in Fig. 9.16. If there is a sign inversion, then
is shown in Fig. 9.15. Feedforward transfer function G has it must be fixed either mathematically by simply removing
been split into two transfer functions, GController and GPlant, the negative sign or physically by inserting an “inverter”
where “plant” is control jargon for the system under control. into the signal path.
The controller uses error signal E to produce control sig- Figure 9.16 shows the “open-loop” configuration, where
nal U, which drives the plant. Suppose the plant was a DC the feedback loop is opened, or broken, before the summa-
motor running in steady-state, when the input, or reference tion junction. If feedback signal CH is zero, then input R is
signal R, was increased by a step change. The difference be- the only information available to the controller. Regardless
tween input R and feedback signal CH, also known as error of the load on the motor, the controller sends the same con-
E, is greatest at the instant the step command is given to the trol signal. Open-loop control is the least expensive, in terms
system, but before the system has responded. As motor speed of hardware. However, the control algorithm is designed for
9.3  Block Diagrams 533

Control select a controller type, then determine its gains to optimize


Input Error Signal Output the response of the system.
R E U C
+
- K G Plant
9.3.7.1  Collapsing Nested Loops
Feedback and feedforward loops are often “nested” within
HSensor one another. The simplest case is when none of the branches
Feedback of the inner and outer loops overlap. For example, the block
CH diagram in Fig. 9.19a has two nested feedback loops.
Reduction of nested loops always proceeds from the in-
Fig. 9.17   Proportional closed-loop negative feedback control
side out. This strategy applies to both feedback loops and
feedforward loops. Collapsing the inner loop using the re-
the expected operating conditions, which are not necessarily G
duction formula GC .L. ( s ) = yields the block diagram
the actual conditions the system experiences. 1 + GH
The simplest closed-loop negative feedback controller is Fig. 9.20.
a “proportional” controller, Fig.  9.17. A proportional con- The reduction formula is used again to collapse the outer
troller multiplies error signal E by a constant, the controller loop, Fig. 9.21. The final step is to clear fractions. Clearing
gain, K, to produce control signal U. Hence, the control sig- fractions often requires more effort than the block diagram
nal is proportional to the error signal. algebra since the transfer function F, G, and H are ratios of
When a block diagram with a controller is reduced to a polynomials.
single transfer function, the controller’s gain, K, is kept as a The final form of a transfer functions is a proper ratio,
factor in the feedforward transfer function, Fig. 9.18. Fig.  9.22. This result is in symbolic form. After the trans-
Recall the denominator of a transfer function is the trans- fer functions F, G, and H are substituted, fractions must be
fer function’s characteristic function, which, when set equal cleared again. The denominator is then expanded into a poly-
to zero, yields the transfer function’s characteristic equation. nomial and factored.
The closed-loop transfer function: Using the following transfer functions as an example,

C KG 4 3 5
GC .L. = = (9.17) F (s) = G (s) = H (s) =
 R 1 + KGH s+6 s+9 s+7

yields the characteristic equation of the closed-loop system C G


=
R 1 + GH + FG

 1 + KGH = 0 (9.18)
3
C s+9
The eigenvalues of the closed-loop system and, hence, the =
closed-loop system’s transient response are a function of the R  3  5   4  3 
1+  +
controller gain, K. The process of control system design is to  s + 9   s + 7   s + 6   s + 9 

Fig. 9.18   a Block diagram of


closed-loop negative feedback
a b
control with controller gain K as R C R KG
______ C
a factor. b Closed-loop system’s +
- KG
transfer function 1+KGH

Fig. 9.19   Nested feedback loops


a b Outer Loop
Inner Loop
R C R C
+
- +
- G +
- +
- G

H H

F F
534 9  Transfer Functions, Block Diagrams, and the s-Plane

R G
_____ C The numerator is in factored form and can be left alone.
+
- 1+GH Expand the denominator

F C 3 ( s + 6) ( s + 7 )
= 3
R ( )
s + 22 s + 159 s + 378 + (15s + 90) + (12 s + 84)
2

Fig. 9.20   Inner loop collapsed


Collect the denominator into a polynomial

G
______ C 3 ( s + 6) ( s + 7 )
=
R 1+GH
_________ C R s 3 + 22 s 2 + 186 s + 552

FG
______ Factor the denominator into first-order and underdamped
1+
1+GH second-order factors

Fig. 9.21   Outer loop collapsed C 3 ( s + 6) ( s + 7 )


=
(
R ( s + 6.42) s 2 + 15.58s + 85.98 )
R G
_________ C 9.3.7.2  Untangling Loops
1+GH+FG If loops overlap and are, therefore, “entangled,” then block
algebra must be used to “untangle” the loops before the loops
Fig. 9.22   Transfer function equivalent to the block diagram in Fig. 9.19 can be collapsed. The reduction formulae for feedback and
feedforward loops can only be used to collapse untangled
loops. Most of the effort expended in block diagram algebra is
The TI-Nspire calculator function propFrac ( ) will evaluate invested untangling entangled loops so that the loop reduction
this transfer function symbolically and return a proper ratio. formulae can be applied from the inner most loop outward.
The manual technique follows. The process of untangling loops requires a combination
of moving branch points, adding blocks, and dividing sum-
( s + 6)( s + 7)( s + 9)
Multiply by the unity ratio mation junctions. Moving summation junctions through a
( s + 6)( s + 7)( s + 9) block can be done, but it is more difficult than the other ac-
3 tions. Consider the block diagram in Figs. 9.23 and 9.24.
C ( s + 6)( s + 7 )( s + 9) s+9 There are two options. Either move Branch Point A so
= that it is to the right (output side) of Block H2, Fig. 9.25, or
R ( s + 6)( s + 7 )( s + 9)  3  5   4  3 
1+  +
 s + 9   s + 7   s + 6   s + 9  move Branch Point B so that it is to the left (input side) of
Block H2, Fig. 9.26.
Both options involve “sliding” a branch point through
Distribute and cancel common factors Block H2. Consequently, a block must be added in either case
to compensate for moving the branch point. If Branch Point
A is slid forward through Block H2 to the right of Branch
C
=
( s + 6) ( s + 7)(3) 1
R ( s + 6) ( s + 7 ) ( s + 9) + ( s + 6)(3)(5) + ( s + 7 )( 4)(3) Point B, then the inverse operation, must be added to
H2
the Branch Point A’s branch to remove the effect of H2. Con-
versely, if the Branch Point B is slid backward through Block

Fig. 9.23   Block diagram with


entangled feedforward and feed- H1
back loops

R + G1 H2 ++
C
-

G2
9.4  Time-Domain Block Diagrams of Differential System Equations 535

Fig. 9.24   Block diagram with


entangled feedforward and feed- H1
back loops Branch Branch
Point A Point B
R + G1 H2 +
C
- +

G2

Fig. 9.25   Branch Point A slid


1
___
forward through block H2 to the H1
right of Branch Point B Branch H2
Point B
R + G1 H2 +
C
- Branch +
Point A

G2

Fig. 9.26   Branch Point B slid


backward through Block H2 to H1
the left of Branch Point A Branch
Point B
R H2 C
+
- G1 +
+
Branch
Point A
G2 H2

H2 to the left of Branch Point A, then the operation H2 must 9.4.1 Block Diagram Without Differentiation
be added to Branch Point B’s branch. of the Input
The loops can be collapsed once they are untangled. Re-
member to identify whether a loop is a feedforward or a feed- We will illustrate the technique with a third-order differential
back loop by the relative direction of the information flow. equation, with u as the input variable, and y as the output
A feedforward loop has the information flowing in the same variable
direction (possibly backward) on both sides of the loop paral-
lel to the input-to-output direction. A feedback loop has infor- d 3y d 2y dy
b 3u = a 0 + a + a2 + a3 y
mation flow in both directions. Do not rely on the signs at the dt 3 1
dt 2
dt
summation junction to identify either type of loop.
The first step is to put the differential equation into standard
form, by clearing the highest-order derivative of the output
9.4 Time-Domain Block Diagrams variable of its coefficient
of Differential System Equations
 b3  d 3 y  a1  d 2 y  a2  dy  a3 
A differential equation can be represented graphically as a  a  u = + + + y
dt 3  a0  dt 2  a0  dt  a0 
block diagram. First, the differential equation is manipulated 0

to represent the highest-order derivative of the output vari-


able, as a summation of the other output variable terms plus where the constant terms are shown in parentheses. What are
the input variable terms. Then, the highest derivative of the the unknowns in this equation? The constants are known,
output variable is integrated to yield the next highest deriva- since they are terms comprised of element parameters. Input
tive. The process is repeated by a series of integrators, until u( t) is also known. We must know the input, i.e., what we
the output of the last integrator is the output variable itself. are doing to the system, if we wish to predict the response of
536 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.27   A chain of integrators


yielding progressively lower-
d3 y
___ d2 y
___ dy
___
order derivatives dt 3 1
__ dt 2 1
__ dt 1
__ y
s s s

Fig. 9.28   Integrators and coef- d3 y d2 y


___ ___ dy
___
ficients of output variable terms
dt 3 1
__ dt 2 1
__ dt 1
__ y
s s s
d2 y
a3 ___
__
a0 dt 2 a3
__
a0
a2 ___
__ dy
a0 dt a2
__
a0
a1
__ y
a0 a1
__
a0

Fig. 9.29   Time-domain d3 y d2 y dy


___ ___ ___
block diagram of y
u b3
__ dt 3 1
__ dt 2 1
__ dt 1
__
d 3y d2y
b3u = a 0 3 + a1 2 + a 2
dy
+ a3 y a0 +_ _ _
s s s
dt dt dt
a3
__
a0
a2
__
a0
a1
__
a0

the system. Consequently, only output variable y( t) and its The time-domain block diagram representation of the dif-
derivatives are unknown. ferential equation can be used to define a set of state vari-
Now, express the differential equation as a summation, ables. Recall the standard form of state equations is a set of
yielding the highest-order derivative of the output variable first-order differential equations
dx
d 3 y  b3   a1  d 2 y  a 2  dy  a 3  = Ax + Bu
= u −  a  dt 2 −  a  dt −  a  y dt
dt 3  a0  0 0 0
Remember that following the flow of a block diagram back-
ward through a linear operator is equivalent to performing
We integrate the highest derivative, d 3 y /dt 3 , with respect the inverse operation represented in the block
to time to yield the second highest derivative, d 2 y /dt 2 . The State variables are defined on the block diagram, start-
process is repeated by arranging the integrators in series, ing with the output variable. Define the output variable as
so that the output of an integrator is the input to the next, x1. The input to the final integrator is the derivative of x1 ,
until the final integrator’s input is dy /dt and its output is dx1 /dt . This signal is given two names by also defining it
y (t ). The chain of integrators, Fig. 9.27, produces the output as the second state variable, x 2. Assigning two names to the
variable and all of its derivatives, allowing us to create the same signal creates an equation
summation.
The rule is simple: Never differentiate the output signal
or its derivative. Only integrate them. The output signals of  dx1
= x2 (9.19)
the integrators are then multiplied by a constant term and fed dt
back to the summation junction, Figs. 9.28 and 9.29.
9.4  Time-Domain Block Diagrams of Differential System Equations 537

Fig. 9.30   Moving in the reverse


direction through a block inverts
a b
dy
__ dy
__
the operation of the block.
Forward through an integrator, or dt 1
__ y dt ó y
dt
reverse through a differentiator, s
ó
is the same operation, because
the operations are inverses
dy
__ dy
__
dt y dt y
s dt

Fig. 9.31   Define state variables Define state variables working


starting with x1 ≡ y, where y is
the output variable, then work
BACKWARD through the integrators.
backward through the integrators. dx3 dx1
dx2
Define as many state variables x3 = x2 = x1
as there are integrators, i.e., the dt 1
__ dt 1
__ dt 1
__
number of state variables equals
the order of the differential d3 y s d2 y s dy s y
equation
dt 3 dt 2 dt

Remember there are only as many state variables as there State variables generally have no physical meaning. If the
are integrators. Do not define the derivative of the final state output variable, x1, were displacement, its third deriva-
variable as a state variable, or you will define one too many. tive would be the derivative of acceleration, also known as
The derivative of the final state variable is the output of the “jerk.” If the output variable were pressure, the only physical
summation junction, Figs. 9.30 and 9.31. meaning of its third derivative would be as the derivative
The state equations are those equations, which can be of the second derivative of pressure. State variables are not
written for the inputs to each integrator. All but the final state unique. Since we use the linear graph method to extract state
equation will define a state variable as the derivative of the equations from energetic systems, our state variables are the
next one. The final state equation will be a summation, in- energy storage variables of the system. However, once we
cluding all of the state variables and the input. perform any matrix algebra on the state equations, say to
The state equations are manipulate them into a form preferred for Modern Control
theory, we create a new set of state variables to describe the
dx1 same system.
= x2
dt
dx2
= x3
 dt 9.4.2 Block Diagram with Differentiation
dx3  a3   a2   a1   b3  of the Input
= −   x1 −   x2 −   x3 +   u (9.20)
dt  a0   a0   a0   a0 
The third-order system equation we expressed as a block
diagram was a special case. There was no differentiation of
and, in matrix form the input variable. The general case of a third-order system
equation is

 0 1 0   0 
 x1    x1   0  d 2u d 2u du
 b0 + b1 2 + b 2 + b3u
d    0 0 1     dt 2
dt dt
x2 = x + u
dt     a 3   a2   a 1    2   b 3   d2y d2y dy
 x3   −  a  −  −     x3     = a0 2 + a 1 2 + a 2 + a3 y
  0   a0   a0    a 0   dt dt dt

where the input variable can be differentiated up to the degree


(9.21) of the output variable, but not higher. A system equation, in


538 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.32   Block diagram of the


summation, Eq. 9.23 u b2
___
a0
du
__
dt b1
___
a0
d2 u
___ d2 y
___
dt 2 b0
___ ++ dt 2
+ __
a0
dy
__
a1
___ dt
a0
a2
___ y
a0

which the input variable has higher-order differentiation than d 2 y  b0  d 2 u  b1  du  b2   a1  dy  a 2 


=   2 +  + u−  − y
the output variable, describes a system which responds to  dt 2
 a0  dt  a0  dt  a0   a0  dt  a0 
what the input is going to do, not what the input just did.
Energetic systems (without intelligence) do not anticipate (9.23)
the future. The mathematical term for a system with higher 
differentiation of the input than the output is “non-causal,” The block diagram of the differential equation is constructed
meaning that it violates cause and effect. by starting with the summation junction. The output of the
Differentiation of the input function presents difficulties summation junction is the highest-order derivative of the
for the state-space representation of a dynamic system, be- output variable, Fig. 9.32.
cause state equations can only have derivatives of the state The block diagram is completed, by adding differentia-
variables, not of the inputs. There are two special cases tors to create the derivatives of the input variable, integrators
which can be implemented. to create the output variable, and its derivatives, Fig. 9.33.
Case One: The input functions are known. In this case, Note there are two integrator blocks in the block diagram.
the derivatives of the input function can be evaluated analyti- The number of integrators equals the order of the differen-
cally, and treated as additional inputs. This case is identical tial equation. Note the input signal terms are fed forward,
to a system with multiple undifferentiated inputs. and the output variable terms are fed back. Although we
Case Two: The input functions are known to be con- can implement this block diagram, if input u( t) is a continu-
tinuous and smooth. If the input function is continuous and ous, smoothly varying function which can be differentiated
“smoothly” varying, then a finite difference approximation numerically, that is not the preferred form. The difficulty
of a derivative will be reasonably accurate. However, the lies in numerical differentiation being sensitive to the size
finite difference approximation of the derivatives of discon- of time step ∆t. If the input function changes smoothly, the
tinuous yields meaningless results which are inversely pro- approximation of its derivative can be accurate. However,
portional to the time step ∆t. when the function changes rapidly, the size of the time step
determines the value of the derivative.
9.4.2.1  Numerical Differentiation of the Input
We will illustrate the method using a generic second-order 9.4.2.2 Variable Redefinition to Eliminate
differential equation, with u as the input variable, and y as Differentiation of the Input
the output variable It is possible to manipulate the differential equation to elimi-
nate differentiation of input. Recall we introduced a method
d 2u du d2y dy of representing system equations with derivatives of the
b0 + b1 + b 2 u = a0 2 + a 1 + a 2 y (9.22)
 dt 2 dt dt dt input terms in Chap. 7. The method was presented as an
algorithm symbolically (or formulaically) for expressing a
which we rearrange as a summation, equaling the highest- transfer function in “Controllable Canonical” Form. Recall
order derivative of the output variable that “canonical” means standard. Controllable canonical
9.4  Time-Domain Block Diagrams of Differential System Equations 539

form is very important in Modern Control (state-space based


control), because it is the form in which coefficient matrix A
is most easily “diagonalized,” thereby allowing individual

y
controller gains to be applied to each state variable.
We will now derive the similar algorithm, to express a

__
1
higher order, differential system equation in “Observable

s
Canonical” state-space form. We will develop the method
graphically using a block diagram, and then extract the re-

___
a2
a0
cursive algorithm from the block diagram. We will use the
dy
__
dt
same general fourth-order system equation used in Chap. 7.
__
1
s

___
a0
a1
d4y d3y d2y dy
a0 4
+ a1 3
+ a2 2
+ a3 + a4 y
dt dt dt dt
d 4u d 3u d 2u
___

du
d2 y
dt 2

= b0 4
+ b1 3
+ b2 2
+ b3 + b4 u
dt dt dt dt
+ __
++

The first step is to express the differential equation in stan-


dard form, by clearing the highest-order derivative of the
output variable of its coefficient.
___
a0
b0

d 4 y  a1  d 3 y  a2  d 2 y  a3  dy  a4 
+ + + + y=
dt 4  a0  dt 3  a0  dt 2  a0  dt  a0 
___
a0
b1

 b  d 4 u  b  d 3u  b  d 2 u  b  du  b4 
+ 0 4 + 1 3 + 2 2 + 3 + u
 a0  dt  a0 
___
d2 u
dt 2

 a0  dt  a0  dt  a0  dt
___
a0
b2

Now, isolate the highest derivative of the output variable on


s

the left-hand side of the equation (or, in other words, solve


for it)
du
__
dt

d4y  a1  d 3 y  a 2  d 2 y  a 3  dy  a 4 
dt 4
= −  a  dt 3 −  a  dt 2 −  a  dt −  a  y
0 0 0 0
s

 b  d 4u  b1  d 3u  b2  d 2u  b3  du  b4 
+ 0 4 +  3 +  2 +  + u
 a0  dt  a0  dt  a0  dt  a0  dt  a0 
u

Next, integrate both sides the number of times necessary, to


integrate the highest derivative output variable down to the
Fig. 9.33   Block diagram of the second-order differential equation,
output variable, four times in this case
Eq. 9.21, with differentiation of the input function

d4y
∫ ∫ ∫ ∫ dt 4
dt dt dt dt

  a 1  d 3 y  a 2  d 2 y  a 3  dy  a4   b 0  d 4 u  b1  d 3u  b 2  d 2 u  b 3  du  b4  
= ∫ ∫ ∫ ∫ −   3 −   2 −   −  y+  4 +  3 +  2 +  + u  dt dt dt dt
  a0  dt  a0  dt  a0  dt  a0   a0  dt  a0  dt  a0  dt  a0  dt  a0  
540 9  Transfer Functions, Block Diagrams, and the s-Plane

On the right-hand side, collect the input and output terms of This integral equation represented as a block diagram in
the same order of differentiation, and distribute the integrals Fig. 9.34.

 b  d 4u    a  d 3 y  b  d 3u    a  d 2 y  b  d 2u 
y = ∫ ∫ ∫ ∫  0  4  dt dt dt dt + ∫ ∫ ∫ ∫  −  1  3 +  1  3  dt dt dt dt + ∫ ∫ ∫ ∫  −  2  2 +  2  2  dtdtdtdt
 a0  dt    a0  dt  a0  dt    a0  dt  a0  dt 
  a  dy  b3  du    a4   b4  
+ ∫ ∫ ∫ ∫ −  3  +   dtdtdtdt + ∫ ∫ ∫ ∫  −   y +   u  dt dt dt dt
  a0  dt  a0  dt    a0   a0  

Evaluate the integrations, neglecting the constants of inte- We can represent this fourth-order system equation
grations. (The integration constants represent initial condi-
tion terms, which will be satisfied when the set of state equa- d4y d3y d2y dy
a0 4
+ a1 3
+ a 2 2
+ a3 + a4 y
tions are solved.) All integrations will be performed on the dt dt dt dt
fourth-order derivative of the input and output variable. Only d 4u d 3u d 2u du
three will be performed on the third-order derivative, leaving = b0 4 + b1 3 + b2 2 + b3 + b4 u
dt dt dt dt
one left over. Two will be performed on the second deriva-
tives, leaving two left over, etc. in state-space by extracting the state and output equations
from the block diagram. The first step is to identify the state
b    a1   b1   variables as outputs of the integrators, and derivatives of the
y =  0  u + ∫  −   y +   u  dt state variables as input to the integrators, starting at the right
 a0    a0   a0  
with the first state variable, x1, and working toward the left.
  a2   b2   Now, write equations for the output of each of the sum-
+ ∫ ∫  −   y +   u  dt
mation junctions, again, working from left to right
  a0   a0  
  a3   b3   b 
+ ∫ ∫ ∫  −   y +   u  dt dt dt y = x1 +  0  u
  a0   a0    a0 
 a  b   dx1  a1   b1 
+ ∫ ∫ ∫ ∫  −  4  y +  4  u  dt dt dt dt = x2 −   y +   u
a
  0  a0   dt  a0   a0 
dx2  a2   b2 
The four integrals are terms of the input variable, and the = x3 −   y +   u
dt  a0   a0 
output variable integrated to increasing degrees. One way
to represent the “flow” of this calculation is to “nest” the dx3  a3   b3 
= x4 −   y +   u
integrals. Integration proceeds from the innermost integral, dt  a0   a0 
which, when evaluated, is summed with the next term. That a  b 
dx4
sum is then integrated, and the process proceeds = − 4  y +  4  u
dt  a0   a0 

b 
y= 0u
 a0 
   
   
     
     
  a1   b1      a 2   b2     a 3   b3    a  b     
+ ∫   −   y +   u  + ∫  −   y +   u  + ∫   −   y +   u  + ∫  −  4  y +  4  u  dt  dt  dt  dt
   a0   a0      a0   a0      a0   a0     a0 

 a0     
 
     
   First Integration   

  Second Integration  
   
9.4  Time-Domain Block Diagrams of Differential System Equations 541

b 
y = x1 +  0  u

Output y
 a0 
dx1  a1   b   b 
= x2 −    x1 +  0  u  +  1  u
dt  a0    a0    a0 

a0
b0

+
dx2  a2    b    b2 

+
= x3 −    x1 +  0  u  +   u
dt  a0    a0    a0 

x1
dx3  a3    b    b3 
= x4 −    x1 +  0  u  +   u
dt  a0    a0    a0 
s
1

dx4 a  b   b 
= −  4   x1 +  0  u  +  4  u
—1

dt  a0    a0    a0 
dx
dt

a0
a1
a0
b1

+
+
-

Distribute and separate variables to yield the proper form of


the state and output equations:
x2

b 
s
1

y = x1 +  0  u
 a0 

dx2

dx1  a1    b1   a1   b  
dt

= −   x1 + x2 +    −    0   u
dt  a0    a0   a0   a0  
a0
a2
a0
b2

+
+
-

dx2  a2    b2   a 2   b  
= −   x1 + x3 +    −    0   u
dt a   a0   a0   a0  
x3

 0
dx3  a3    b3   a 3   b  
= −   x1 + x4 +    −    0   u
s
1

dt a
 0   a0   a0   a0  
a   b   a   b 

dx4
dx3
dt

= −  4  x1 +   4  −  4   0   u
dt  a0    a0   a0   a0  
a0
a3
a0
b3

+
+
-
x4

The state and output equations expressed in matrix form are


s
1

  a1     b1   a 1   b0  
−  1 0 0   −     

  a0    a0   a0   a0  
dx4


dt

   
 x1    a 2   x   b2   a 2   b 
a0
a4
a0
b4

0 1 0  1     −    0  
+
-

  − 
d  x2    a0   x2  a0   a0   a0  
=  + u
Input u

dt  x3    a 3    x3    b 3   a 3   b0  
  − 0 0 1      −     
 x4    a0  x
  4    a0   a0   a0  
 a    b   a   b  
Fig. 9.34   Block Diagram of Observable Canonical form − 4 0 0 0  4 − 4 0 

  a0     a0   a0   a0  

The first equation is the output equation, which, importantly,


 x1 
allows us to eliminate output variable y from the state equa- x   
tions. Substitute for y in the state equations b
y = [1 0 0 0]   +  0  u.
2

 x3   a0 
 
 x4 
542 9  Transfer Functions, Block Diagrams, and the s-Plane

Reversing the order of the state variables, which interchang- R L


es the rows and columns of coefficient matrix A, puts the 1 2 3
state and output equations into Observable Canonical form:

 a    b4   a4   b0  
0 0 0 − 4     −      v(t) C
  a0     a0   a0   a0  
   
 x4    a 3   x   b3   a   b 
  1 0 0 −     4    −  3   0  
d  x3    a0   x3   a0   a0   a0   g g
=  + u
dt  x2    a 2    x2   b 2   a2   b0  
  0 1 0 −        −      Fig. 7.6   Linear graph of the series RLC circuit of Fig. 7.5
 x1    a0    x1   a0   a0   a0  
   b 
0  a1    1  −  a 1   b0  
0 1 −  
  a0     a0   a0   a0   9.5.1 Drawing a Block Diagram of an Existing
  
State Equation
 x4 
x   
b If a set of state equations has been derived for a system, then
y = [0 0 0 1]   +  0  u
3

 x2   a0  constructing a block diagram for the system using the state


  equations is straightforward. We will use the energy storage
 x1 
variables of this system, as its state variables. The state equa-
tions, derived in Sect. 7.6.2.3 from using current through the
inductor iL as the energy storage variable and voltage across
9.5 State Equations as a Time-Domain Block the capacitor v3g, as the state variable, are
Diagram
diL R 1 1
= − iL − v3 g + v1g
Presented below are three methods to represent the state  dt L L L (7.45)
equations of a system as a block diagram. One method is to dv3 g 1
reduce the equation list of the state equations, then construct = iL
dt C
the block diagram from the state equations. A second meth-
od is to draw the block diagram directly from the equation Note the state equation for iL is a summation, but the state
list. This method can save time but requires experience. The equation for v3g is not a summation.
third method is to define a new set of state variables from a The structure of a block diagram of state equations usual-
higher-order system equation. The latter can also be used for ly has as many summation junctions as there are state equa-
system equations, which do not have derivatives of the input tions, since state equations are usually sums, Fig. 9.35. The
variable. output signals of the summation junctions are the derivatives
All three techniques will be illustrated, using the series of the state variables. The block diagram also has as many in-
RLC circuit from Sects. 1.4.2.2 and 7.6.2. The linear graph tegrators as state equations, since the derivatives of the state
and the energetic equations of the series RLC circuit from variables must be integrated to yield the state variables.
Sect. 7.6.2.1 are reproduced below, for reference.

Energetic Equations 9.5.2 Drawing a Block Diagram


from the Energetic Equations
Compatibility:  v(t ) = v12 + v23 + v3 g
The method of drawing the state equations directly from the
Continuity:  isource = iR iR = iL iL = iC set of energetic equations of system is not very practical, but
dv3 g it is useful as an aid in understanding the parallels between
diL
Elements:  v12 = RiR v23 = L iC = C the symbolic mathematics and block diagrams. The method
dt dt
follows the procedure for deriving the state equations from
1 1
Energy:  E system = E L + EC E L = LiL2 EC = C v32g . the equation list. Start with the elemental equations for the
2 2 energy storage elements. The signals, which are derivatives
9.5  State Equations as a Time-Domain Block Diagram 543

Fig. 9.35   Block diagram of the v3g


state equations, Eq. 7.45
1
__
L
di L
v — iL iL
1
__ +
- dt 1
__
Input L - s State Variable

R iL
__
L

1
__ iL
C
dv3g
___
dt 1 v3g v3g
__
s State Variable

of the state variables, will be the input to integrator blocks. a


The integrators yield the state variables, Fig. 9.36. di L
___
v23 1
__ dt 1 iL
__
Inductor: L v23 s
___
di v23 diL L
v23 = L L →
(9.24) =
dt L dt b
Capacitor: dv
___
3g
iC 1 dt 1 v3g
dv3 g iC dv3 g __ __
 iC = C → = (9.25) C s
dt C dt iC
__
C
The block diagrams are built backward from the integrators.
We need to provide the signals fed into the integrators, in Fig. 9.36   Block diagrams yielding the state variables a the current
terms of the input, v1g, and the state variables, iL and v3g. through inductor iL, from Eq. 9.24 and b The voltage across the capaci-
Providing signal iC, in terms of the input and the state vari- tor, v3g, from Eq. 9.25
ables, is simple, since there is only one loop in the circuit,
and all the elements have the same current. The current
through the capacitor equals the current through the inductor iR − iL = 0 → v12 = RiL

iL − iC = 0 → iC = iL yields v12 in terms of state variable iL. Substitution of this re-


sult into Eq. 9.26 creates an equation for voltage v23 in terms
Providing signal v23 in terms of the input and the state vari- of the input and the state variables
ables requires a summation junction and a few blocks. Rear-
range the compatibility equation to yield voltage v23  v − v12 − v3 g = v23 → v − RiL − v3 g = v23 (9.27)

v = v12 + v23 + v3 g → v − v12 − v3 g = v23


(9.26) The block diagram of the state equations, Fig. 9.38, is creat-
ed by assembling the block diagram fragments of Figs. 9.36
The input to the system is v, and v3g is a state variable. We and 9.37.
need to create v12 in terms of the input and the state variables.
The element equation for the resistor relates voltage v12 to
current iR 9.5.3 Drawing a Block Diagram of State
v12 = RiR Equations from a System Equation

Current iR is not a state variable, but it equals the current This method defines new state variables from the higher-
through the inductor iL, which is a state variable. Using the order system equation. The block diagram of the state equa-
continuity equation for node two with the element equation tions uses the new state variables. Our preference is to use
for the resistor the energy storage variables of a system as its state variables.
544 9  Transfer Functions, Block Diagrams, and the s-Plane

v12 iL We start the recursive definitions, using output variable v3g


R as the first state variable, x1,
v12 iR
(9.28)
v3 g ≡ x1
v - v23
+
- The recursion is to define the next state variable as the de-
rivative of the previous state variable:

v3g  dv3 g dx1


= ≡ x2 (9.29)
dt dt
Fig. 9.37   Block diagram of Eq. 9.27
Substituting the state variables, x1 and x2, and their deriva-
tives into Eq. 1.49 yields two equations, either
These state variables have physical meaning. They are
 dx2 dx1
physical quantities that can be measured. The energy vari- v = LC + RC + x1
dt dt (9.30)
ables of a system do, indeed, define the energetic state of
the system. However, state variables are not unique. As we or
saw in Chap. 7, linear operations performed on a set of state dx 2
variables yield a new and valid set of state variables. Al-  v = LC + RCx2 + x1 (9.31)
dt
though alternative sets of state variables are mathematically
legitimate, they may have no direct physical meaning. Con- Which form do we use for a state equation? Equation 9.30
sequently, we will continue to emphasize that energy storage cannot be expressed as a state equation, because it has two
variables are state variables. One important application of an dx dx2
derivative terms, 1 and . A state equation can have
alternative set of state variables is to enable the expression of dt dt
a higher-order system equation in state-space. the derivative of only one state variable.
Recall from Chap. 7, that a higher-order system equation We will use Eq. 9.31 where we have made the substitution
without derivatives of the input variable can be expressed dx1
= x 2 ,to eliminate the extra state variable derivative. The
in state-space, using a recursive definition of the state vari- dt
ables, based on the output variable. Let us again use the RLC substitution necessary to eliminate the extra state variable
circuit example from Chaps. 1 and 7. The system equation, is the first state equation, Eq. 9.29. Rearranging Eq. 9.31 to
which relates input voltage v1g to voltage across capacitor express the derivative of state variable x2 as the sum of prod-
v3g, is this second-order differential equation ucts of constant terms, state variables and input, yields the
state equation:
d 2 v3 g dv3 g dx2 1 R 1
 v1g ≡ v = LC 2
+ RC + v3 g (1.48)  =− x1 − x 2 + v1g (9.32)
dt dt dt LC L LC

Fig. 9.38   Block diagram of v12 iL


the RLC circuit drawn from its R
energetic equations iR
v12
di L
___
v
+
- v23 1
__ dt 1
__ iL
- L v
___
23 s
v3g L

dv
___3g
iC 1 dt 1 v3g
__ __
C iC
__ s
C
9.6  The s-Plane 545

dx 2
___ The state equation for x 2 , Eq. 9.32, is a summation of the
dt 1
__ x2 1
__ x1 terms of input v and state variables, x1 and x 2 ,. Fig. 9.40. The
s dx1
___ s complete block diagram is formed by adding gain blocks to
dt multiply input v and state variable x1 by 1/LC, and state vari-
able x 2 by R/L, Fig. 9.41.
Fig. 9.39   Block diagram of Eqs. 9.33 and 9.34. Integrators yielding
the state variables, x1 and x2

9.6  The s-Plane


1
___ x We will plot values of the roots of the characteristic equa-
LC 1
tion, the eigenvalues of the system. Since the roots may be
dx 2
___ complex, we will use a complex plane with the real axis,
v
+
- dt
σ, and the imaginary axis, jω. Convention is to omit use of
- the subscript, “d ”, to indicate the frequency is the observed,
damped natural frequency of the physical system. It is un-
R
___ derstood that observed frequency would never be the ideal,
x
L 2 undamped, never-seen-in-nature, natural frequency, ωn.
Although we commonly refer to “the” complex plane,
Fig. 9.40   Block diagram of Eq. 9.31, the state equation for x2 it is correct to refer to “a” complex plane. Every complex
function is an input–output relationship associated with two
complex planes, one for the input and one for the output.
Reordering and summarizing the state equations, Eqs. 9.29, Complex variable s of the s-plane is the variable, introduced
9.32, and the output equation, Eq. 9.28, when we solved system equations, using the method of un-
determined coefficients. The s-plane is the plane of the roots
dx1 of the characteristic equation of a system equation. These are
= x2
dt the characteristic values or eigenvalues of the system. The
dx 2 1 R 1 symbol λ is generally used when eigenvalues are derived
=− x1 − x 2 + v (9.33) from the coefficient matrix A of the state-space representa-
 dt LC L LC
tion of a system. The symbol s is used when the eigenvalues
v3 g = x1
are calculated from the characteristic equation derived from
the differential system equation or the transfer function.
The block diagram is constructed, starting with the output of In the general case, s is a complex number, s = σ + jw .
the summation junction, which is the derivative of the sec- The values of s are the exponents of the homogeneous solu-
ond state variable. We need a cascade of integrators in series tion. If s is a purely real number, s = σ, the exponentials are
to yield the state variables, x1 and x2, Fig. 9.39. real exponentials decays. If s is a complex number, then the
homogeneous solution is underdamped and may oscillate. In
 dx 2 dx1
that case, the homogeneous solution is the product of a real
∫ dt
dt = x 2 =
dt
(9.34)
exponential and the sum of two exponentials with imaginary
exponents, jω, which represent the oscillation.
dx1
∫ dt dt = x1
(9.35)

Fig. 9.41   Block diagram of Eq. 1.48 1


— x1
LC 1 x1

1 LC
— x1
LC
1 dx2 dx1
—v — —
v 1 LC - dt 1 dt 1 x1
— + — —
Input LC - s x2 s y
Output
R
— x2
L R x2

L
546 9  Transfer Functions, Block Diagrams, and the s-Plane

We will use a second-order system with a damped, oscil- 9.6.1 Poles, Zeros, and Pole-Zero Transfer
latory homogeneous response as the example. The general Function Form
form of the characteristic equation is
A useful form of a transfer function, and the form we use
s 2 + bs + c = 0 → s 2 + 2 n s +  n2 = 0 to design control systems, is to express the numerator and
denominator of a transfer function as the product of factors,
Solving the characteristic equation or, in other words, in factored form. The factored form of the
denominator allows us to see the roots of the characteristic
−b ± b 2 − 4c −2ζw n ± (2ζw n )2 − 4w n2 equation, the system’s eigenvalues, by inspection. The roots
s1 , s2 = = of the numerator are important in the design of control sys-
2 2
tems, but a discussion of their effects requires introduction to
−2ζw n ± 4ζ 2w n2 − 4w n2 −2ζw n ± 2 ζ 2w n2 − w n2 the “root locus,” a topic in classical control theory.
s1 , s2 = = Equation 9.36 is a second-order transfer function ex-
2 2
pressed in factored form
s1 , s2 = −ζw n ± ζ 2w n2 − w n2  Output b0 s 2 + b1 s + b2 K ( s + z1 )( s + z2 )
G (s) = = 2 =
Input s + a1 s + a2 ( s + p1 )( s + p2 )
The damping ratio ζ must be less than one for the homoge-
neous response to be oscillatory. This leads to complex roots (9.36)

ζ < 1 ⇒ ζ 2w n2 < w n2 ⇒ ζ 2w n2 − w n2 < 0. The roots are the opposites of the constants, i.e.,
s = − p1 and s = − p2 are the roots of the denominator, and
Multiplying both terms by − 1 in the form of j2 reverses the s = − z1 and s = − z2 are the roots of the numerator.
order of the terms in the difference Thus far, we have used transfer functions as multiplica-
tive operators. It is also possible to evaluate them as con-
ζ 2w n2 − w n2 = j 2w n2 − j 2ζ 2w n2 ventional functions by substituting a complex number for
the variable, s = σ + jw . The result of evaluating a complex
and yields the familiar form when j 2 = j is factored out function for an argument is a complex number, which can be
of the radical expressed in Cartesian form, polar form, or as the product of
its magnitude and a complex exponential unit vector.
s1 , s2 = −ζw n ± ζ 2w n2 − w n2 = −ζw n ± j 2w n2 − j 2ζ 2w n2
G ( s ) = G (σ + jw ) = x + jy = G (σ + jw ) e jG (σ + jw )

s1 , s2 = −ζw n ± j w n2 − ζ 2w n2 = −ζw n ± jw n 1 − ζ 2 (The terminology of evaluating a transfer function is awk-


ward, because the angle of a complex number is its “argu-
Matching real and imaginary terms ment,” the term used for the value operated on by a function.)
The magnitude of the transfer function, G( s), evaluated
s1 , s2 = σ ± j w  σ = −ζw n
 → for s = − z1 or s = − z2 is zero, since the magnitude of the
s1 , s2 = −ζw n ± jw n 1 − ζ 
2
j w = jw n 1 − ζ 2 numerator is zero.

Note that the real component σ is negative, because the G ( − z1 ) = G ( − z2 ) = 0


damping ratio ζ and the ideal, undamped natural frequency
ωn are both positive. The negative value of the real com- However, the magnitude of the transfer function, G( s), eval-
ponent  σ results from our assumed of a damped, or decay- uated for s = − p1 or s = − p2 , is infinite, since the magnitude
ing, homogeneous response. The imaginary term is not of the denominator is zero.
the natural frequency. It is the damped natural frequency,
which is the frequency one observes. The name, “natural fre- G ( − p1 ) = G ( − p2 ) = ∞
quency,” is unfortunate, because the frequency it refers to is
completely undamped, which is not natural. A better term The values, s = − z1 or s = − z2 , are called “zeros” of the
would be the “ideal” or “undamped” frequency. However, transfer function, G( s), since they yield zero magnitude.
the term, natural frequency, is the nomenclature. For systems The values, s = − p1 and s = − p2 , are called “poles” of G( s).
with damping ratio of ζ ≤ 0.1, the difference between the ob- The term, “pole,” comes from a three dimensional plot, with
served, damped frequency and the ideal, undamped, natural the components of s = σ + jw forming the two axes of a
frequency is 1 % or less. complex plane with the magnitude of the transfer function,
9.6  The s-Plane 547

G ( s ) , serving as the axis normal to that plane, Fig. 9.53. The s-plane jω


magnitude increases sharply, as the value of s approaches j6

s = − p1 or s = − p2 , from any direction in the s-plane. That UNSTABLE


j4
Stable
shape is typically described as a “spike,” but the Laplace σ>0
variable’s symbol is s, so a different term, “pole,” is used. σ<0 j2

Example  The transfer function, G( s), expressed as the ratio


-10 -8 -6 -4 -2 2 4 6 8 10 σ
-j2
of two polynomials and in pole-zero (factored) form
-j4
Output 352 352
G (s) = = 3 = -j6
Input s + 14 s 2 + 56 s + 64 ( s + 2)( s + 4)( s + 8)

Fig. 9.42   The right-half of the s-plane represents unstable systems, be-
Evaluate G( s) for s1 = − 2 + j2, s2 = − 2 + j, s3 = − 2 + j0.5, cause the eigenvalues have a positive real component σ, which leads to
s4 = − 2 + j0.25, s5 = − 2 + j0.1, and s6 = − 2 + j0.05. Present the unbounded growth of the homogeneous, or natural, response
result in Cartesian and complex exponential form.
Evaluating a transfer function in the manner of a con-
ventional function by substituting for the “argument” s, the nitude for values of s along the line σ = −2. Figure 9.53b
value of a complex number can be done on an engineering shows the magnitude for values of s along the line jω  = 0,
calculator which is the real axis.

G ( s1 ) = G ( −2 + j 2)
9.6.2 Stability
352
G ( −2 + j 2) =
( −2 + j 2 + 2) ( −2 + j 2 + 4) ( −2 + j 2 + 8) The homogeneous response of a system is the “natural”
response, which represents how energy flows within the
system, whenever the system is disturbed from a dynamic
352
G ( −2 + j 2) = equilibrium by a change to the system’s input. Specifically,
( j 2) ( 2 + j 2) ( 6 + j 2) we are interested in how quickly energy is dissipated. If the
system oscillates due to internal energy flows, what is the
G ( −2 + j 2) = −8.80 − j 4.40 = 9.84e − j153
o
frequency of the oscillation?
The natural response is known as the “unforced” re-
sponse, which is somewhat of a misnomer, because the nat-
Tabulating the results ural response is produced, when the forcing function (also
known as driving function, or input) is turned on or off, or
Output 352 352 when it changes form, such as from a step to a ramp. Both the
G ( s) = = 3 2
= impulse and step inputs reveal the homogeneous response of
Input s + 14 s + 56 s + 64 ( s + 2)( s + 4)( s + 8)
 a system, and are used as test inputs. The impulse response is
G ( −2 + j 2) = −8.80 − j 4.40 = 9.84e − j153 the natural or homogeneous response, since the particular so-

G ( −2 + j ) = −15.2 − j 20.9 = 25.9e − j126 lution for an impulse response is zero. Although the impulse
 input–impulse response relationship has great theoretical
G ( −2 + j 0.5) = −18.3 − j 53.7 = 56.7e − j109

utility, since any input can be represented by the integration
G ( −2 + j 0.25) = −19.2 − j115 = 116e − j 99.5 of impulses, in practice, the step response is easier to work
G ( −2 + j 0.1) = −19.5 − j 292 = 293e − j 93.8

with due to power limitations. The steady-state response to
 the step input must be subtracted, if it is non-zero, to yield
G ( −2 + j 0.05) = −19.5 − j 586 = 586e − j 91.9 the homogeneous response.
The engineering adage, “all passive systems are stable,”
Note the increase in the magnitude of the result from 9.8 to 586, means that all systems without feedback control are inher-
as the value of s approaches the pole at s = − 2 in the negative ently stable. This is not completely true, since it is possible to
imaginary direction, from s1 = −2 + j 2 to s6 = −2 + j 0.05, construct an unstable passive system, but they destroy them-
forms a spike in magnitude at the pole. It is easier to inter- selves the first time they are operated. Thus the adage applies
pret these results graphically, rather than as a list of complex to all systems which have been used once.
numbers. Figure 9.53a shows the transfer function’s mag- Unfortunately, it is not only possible but common for a
stable, passive system to become de-stabilized, when placed
548 9  Transfer Functions, Block Diagrams, and the s-Plane

A
____ jω s-plane
s+a
Faster Decay
1 A
____ j2

Input s+a Output j

Fig. 9.43   Block diagram of a unit impulse input to a first-order system, -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 1 σ
and the resulting unit impulse response of a first-order system -j

-j2

under closed-loop feedback control. Consequently, the most Fig. 9.44   Purely real poles, s = σ. The distance from the imaginary axis
important design criterion for a control system is stability, is the decay rate, σ
which is assessed by the position of the poles of the closed-
loop system on the s-plane. 1
e− a =
Positive values of the real component σ produce unstable ea
responses. Hence, the roots of a closed-loop system’s charac-
teristic equation must be in the left-half of the s-plane for the Consequently, when the real component σ is negative, such
response to be stable. The imaginary axis is the boundary be- as σ = −ζw n , the real exponential decays with increasing
tween the stable left side and the unstable right side, Fig. 9.42. time t
The reason that the stability of a system is determined by 1
the real component σ is clear from the solution of the sys- e − n t = n t
e
tem equation. Recall the homogeneous solution of a second-
order system is whereas, if the exponent is real, then the real exponential
increases with increasing time.
yh (t ) = A1e 1 + A 2 e 2 = A1e(σ + j w )t + A 2 e(σ − j w )t Physically, a system becomes unstable, because energy ac-
s t s t

cumulates in the system over time. A stable system dissipates


energy. If the homogeneous response of the system is oscilla-
(
yh (t ) = A1eσ t e j wt + A 2 eσ t e − j wt = eσ t A1e j wt + A 2 e − j wt ) tory, the accumulation of energy in the system increases the
amplitude of the oscillations, as the energy flows between
where the undetermined coefficients, A1 and A2, depend on independent, internal energy storage elements. Eventually,
both the system and the initial conditions if the power supplied to the system is sufficiently large, the
amount of energy stored in the system exceeds the capacity
dy (0+ ) or strength of the system, thereby destroying it. In reality,
y (0+ ) = I .C.1 and = I .C.2
dt many systems of interest to mechanical engineers have lim-
ited power supplies. These systems will reach a bounded un-
Expressing the sum of the complex exponentials, stable state, in which their unstable responses reach a finite
( A1e jt + A2 e − jt ), as the trigonometric function limiting value.
To use the s-plane to design controllers, we must be able
A1e jt + A2 e − jt = A sin(t +  ) to associate the position of a closed-loop real pole or complex
conjugate pair of poles on the s-plane with the contribution
by using one of Euler’s equations that pole (or those complex conjugate poles) make(s) to the
time-domain response of a system. A convenient method for
e j θ + e− j θ e j θ − e− j θ doing so is to plot the impulse responses. Recall the Laplace
cos (θ ) = or sin (θ ) =
2 2j transformation of a unit impulse is unity

yields  f (t ) = δ (t ) = ∞ for t = 0

 f (t ) = δ (t ) = 0 for t ≠ 0
( )
eσ t A1e j wt + A 2 e − j wt = eσ t A sin (w t + φ )
  
 F (s) = 1

Sinusoid with
amplitude A Consequently, in the Laplace-domain, the impulse response
and the transfer function have the same form, although they
Amplitude A of the sinusoid is constant. However, the purely are different physical quantities, since the response is a vari-
real exponential multiplying the sinusoid is time-varying. able (or signal), and the transfer function is an input–output
Hence, the real exponential factor governs whether the am- relationship (or dynamic operator), Fig. 9.43. It is essential
plitude of the response grows or decays. that the history of a Laplace-domain calculation is known.
Recall a negative exponent represents a fractional and a The resulting ratio of polynomials in s cannot be assigned
sign inversion: meaning, if the history is unknown.
9.6  The s-Plane 549

1.0 A
_____________
Impulse Response s + 2ζωns + ω2n
2
0.8 1 A
_____________
-10t -5t -2t -t
0.6
e e e e Input s 2+ 2ζωns + ω2n Output

0.4 Fig. 9.46   Block diagram of a unit impulse input to an underdamped


second-order system, and the resulting unit impulse response
0.2

0.0
0 1 2 3 4 5
t, sec s-plane jω
j11
Fig. 9.45   Impulse responses of first-order systems with poles, s =  − 1,
j10
s =  − 2, s = − 5, and s =  − 10

Increasing Frequency
j9

j8
9.6.3  Real Component σ, the Decay Rate j7

j6
Restricting our discussion to the left-half of the s-plane, the
j5
real exponential controls the rate of decay of the homoge-
nous response. A “fast” response decays quickly. Again, the j4

negative real exponential is interpreted as j3

j2
1
e− a = j
ea
Therefore, the more negative (i.e., further left on the nega- -5 -4 -3 -2 -1 1 σ
tive real axis) the real component σ of s = σ + jw , the faster is
decay of the response Figs. 9.44 and 9.45. The time-domain Fig. 9.47   Complex eigenvalues with decay rate, σ = − 2, and increasing
response, corresponding to a purely real value of s, σ is non- frequency ω. Only the complex pole in the upper half-plane is shown,
oscillatory. The relationship between the value of σ and the of the conjugate pairs of poles
“speed” (or decay rate) of the homogeneous response is il-
lustrated with their impulse responses. Note that all of the
responses are first-order decays with the same shape, but response, which is the damped natural frequency, Fig. 9.47.
scaled in the time dimension. If a system is stable, its homo- For convenience, we do not use the subscript, “d,” to indicate
geneous response must decay with time. Hence that the ω of the imaginary component, jω, is the damped
frequency, jωd. It is understood, because all physically sys-
t
− 1 tems have damping due to either to mechanical friction, fluid
eσ t = e τ
and σ = where τ ≡ time constant
τ viscosity, or electrical resistance.
Varying the imaginary component, jω, of a closed-loop
The actual amplitude contributed by a single real pole in a pole, while holding the real component, σ, constant pro-
higher-order transfer function would depend on the coeffi- duces homogeneous responses of varying shapes, as shown
cient determined by the partial fraction expansion, known below. The constant decay rate, σ, creates a fixed duration
as the “residue,” which, in turn, is governed by the transfer for the various responses. Increasing the frequency, jω, in-
function evaluated for the magnitude of the pole, s = σ , after creases the number of oscillations which changes the shape
the factor, ( s + ( −σ )), was removed. This fact will be illus- of the response. We illustrate this with impulse responses in
trated below, when we consider the effect of relative position Figs. 9.48 and 9.49.
on the magnitude of contribution of each of three poles to the Note the peak amplitude of response increases with in-
response of a system. creasing frequency, since the sine function reaches its first
maximum earlier at higher frequency.

9.6.4 Imaginary Component ω, the Observed,


Damped Frequency 9.6.5  Damping Ratio ζ

The imaginary component of the closed-loop pole, The ideal, undamped natural frequency derives from the
s = σ + jw , is the observed frequency of the homogenous standard form of an oscillatory second-order transfer func-
tion factor
550 9  Transfer Functions, Block Diagrams, and the s-Plane

0.8 s-plane
e-2t sin(10t) jω
Impulse Response 0.6
e-2t sin(5t) s = σ + jω jωd = jωn 1-ζ 2
0.4 e-2t sin(2t) ω
e-2t sin(t) n
0.2
θ
0.0

-0.2 −ζωn σ

-0.4 Fig. 9.50   The s-plane, the complex plane used to plot the eigenvalues
0.0 0.5 1.0 1.5 2.0
of dynamic systems. One of a complex conjugate pair of eigenvalues
t, sec is shown

Fig. 9.48   Underdamped second-order impulse responses


Table 9.1  Factor 1−ζ 2 ζ 1−ζ 2
1.0 reduces ωn to  d =  n 1 − ζ 2 0.707 0.707
e -2t
0.5 0.866
Impulse Response

e-2t sin(5t)
0.5 0.4 0.917
0.3 0.954
0.0 0.2 0.980
0.1 0.995
-0.5 0.05 0.999
e-2t sin(50t)
-1.0 quency, ω or ωd, and the ideal, undamped, natural frequency
0.0 0.5 1.0 1.5 2.0
varies with the damping ratio, ζ.
t, sec
Table  9.1 illustrates the difference between the actual,
Fig. 9.49   Underdamped impulse responses and their decay envelop observed, damped frequency, ω or ωd, and the ideal, un-
for systems with s = −2 + j 5 and s = −2 + j 50 damped natural frequency, ωn, is certainly negligible, when
the damping ratio ζ ≤ 0.1.
Closed-loop poles which fall on a ray from the origin of
N (s) the s-plane have the same damping ratio, ζ. This is estab-
G ( s) = lished using Pythagorean’s theorem, to determine the dis-
s + 2ζw n s + w n2
2

tance from the origin to the closed-loop pole, which is the


ideal, undamped natural frequency, ζ ≤ 1, Fig. 9.50.
This form is useful, because the damping ratio, ζ, and period
of a damped oscillation, T, can be determined from an im-
( ) ( )
2

pulse or step response of a second-order system. The damp- (ζw )n


2
+ wn 1 − ζ 2 = ζ 2w n2 + w n2 1 − ζ 2
ing ratio is calculated using the “log decrement” formula.
The damped period yields the damped frequency, ωd
ζ 2w n2 + w n2 −w n2ζ 2 = w n

2π  radians 
 w ≡ wd = = (9.37) The inverse cosine of the angle θ, which the ray from the
T  second  origin makes with the negative real axis to find the damping
ratio, ζ, is shown in the s-plane construction, Figs. 9.50 and
The damped frequency yields the natural frequency, ωn 9.51.
ζw n
d cos (θ ) = =ζ (9.39)
d = n 1 −  2
→ n = (9.38) wn
2
 1−  
Homogeneous oscillatory responses with the same damping
What we call the “natural” frequency, ωn, is, in fact, not nat- ratio, ζ, have similar shapes, Fig. 9.52, but these are com-
ural but an ideal, which would occur in the absence of energy pressed or extended along the time axis, because the damp-
dissipation. The discrepancy between the actual, damped fre- ing ratio, ζ, is a factor in the decay rate, σ = −ζw n, and the
damped frequency, w d = w n 1 − ζ 2 .
9.6  The s-Plane 551

s-plane jω 0.6
j11

Impulse Response
Co
j10 0.4 e-5t sin(10t)

nst
j9 e-3t sin(6t)

an
0.2 e-2t sin(4t)

tD
j8
e-t sin(t)

am
j7
0.0

pin
j6

gR
j5
-0.2
atio j4 0.0 1.0 2.0 3.0 4.0
θ j3 t, sec
j2
Fig. 9.52   Impulse responses of underdamped second-order systems
j with the same damping ratio, ζ. Notice the shape of the curves is the
same, except for the temporal scaling
-6 -5 -4 -3 -2 -1 1 σ
Fig. 9.51   Complex eigenvalues with constant damping ratio ζ fall on form. The standard form has no coefficient multiplying the
a ray from the origin of the s-plane. The angle of the ray is  = 63.4°. cos(63.4° ) = 0.45 = 
Using Eq.
 = 9.39,
63.4°. cos(63.4° ) = 0.45 = 
highest power of s of the denominator polynomial. The nu-
merator and denominator are then factored.

Example One: Express the transfer function, G1( s), in stan-


9.6.6 s-Plane Plots of Transfer Function Poles dard form and plot its poles and zeros (if any) on the s-plane,
and Zeros Fig. 9.54.

A plot of a transfer function’s poles and zeros conveys most 704 352
G1 ( s ) = = 3
of the information of the transfer function in graphical form. 2 s + 28s + 112 s + 128 s + 14 s 2 + 56 s + 64
3 2

Recall that poles and zeros derive their names from the mag- 
nitude of the transfer function evaluated as a conventional 352
G1 ( s ) = (9.40)
function for a value of s. The magnitude spikes when the ( s + 2)( s + 4)( s + 8)
value of s approaches a pole, Fig. 9.53, and is zero when
the value of s corresponds to a zero. In general, there is a Example Two: Express the transfer function, G2( s), in
constant, gain K, which is a factor of the numerator. The gain standard form and plot its poles and zeros (if any) on the
cannot zero the numerator, except in the trivial case of set- s-plane, Fig. 9.55.
ting K = 0. Consequently, the gain is not a zero. The transfer
function’s gain, K, is not represented in a plot of the poles
 G (s) = 816 272
and zeros since it is not a function of s. The converse is true, 3 2
= 3 2
3s + 48s + 252 s + 480 s + 16 s + 84 s + 160
as well. A transfer function can be constructed in factored
form from its pole-zero plot, except for the unknown gain, K.
272
The s-plane provides a graphical design tool for feed- G2 ( s ) = (9.41)
back control system design. For the time being, we will use ( s + 4 − j 2)( s + 4 + j 2)( s + 8)
it to plot the roots of the denominator and the numerator of
transfer functions, after they have been factored into pole- Example Three: Express the transfer function, G3( s), in
zero form. The conventional symbols are Xs for poles, roots standard form, and plot its poles and zeros (if any) on the
of the denominator, and Os for zeros, roots of the numerator. s-plane, Fig. 9.56.

14 s 2 + 168s + 728  14  s 2 + 12 s + 52
G3 ( s ) = =  
9.6.7  Example Pole-Zero Plots 0.5s 3 + 5s 2 + 9s + 8  0.5  s 3 + 10s 2 + 18s + 16

Poles and zeros are the roots of the denominator and numera- 28( s + 6 − j 4)( s + 6 + j 4)
G3 ( s ) = (9.42)
tor polynomials, respectively, of a transfer function. The first ( s + 1 − j )( s + 1 + j )( s + 8)
step in plotting the poles and zeros of a transfer function in
polynomial form is to write the transfer function in standard
552 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.53   Magnitude


of the transfer function,
a 600
G ( s ) = 352 ( s + 2) ( s + 4)( s + 8),
as the value of s approaches the
400
pole at s = − 2. a Section along
line σ =  − 2. b Section along real |G(s)|
axis, ω = 0
200

0
-j2 -j1.5 -j -j0.5 0 j0.5 j j1.5 j2
Imaginary Component ω of s = -2 + jω

b 3,000

2,000
|G(s)|

1,000

0
-3 -2.75 -2.5 -2.25 -2 -1.75 -1.5
Real Component σ of s = σ + j0

s-plane jω s-plane jω
j3 j5

j2 j4

j j3

j2
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 1 σ j
-j

-j2 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 1 σ
-j3 -j

-j2
Fig. 9.54   Pole-zero plot of G1( s), Eq. 9.40 -j3

-j4

-j5
s-plane jω
j4
Fig. 9.56   Pole-zero plot of G3( s), Eq. 9.42
j3

j2

j
Summary
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 1 σ Linear Operators
-j

-j2 The important properties of linear operators are that they can
-j3 be distributed and their output scales with their input. These
-j4
properties allow them to be used with superposition and as
operators in block diagrams. All blocks in a block diagram
Fig. 9.55   Pole-zero plot of G2( s), Eq. 9.41 are a single input–single output relationship, except summa-
tion junctions which can have any number of inputs but only
one output.
Problems 553

Time-Domain Block Diagrams L


Integration and differentiation, inverse operations, are more
conveniently represented by their Laplace-domain symbols
than by their time-domain expressions. Higher-order differ- v(t) C R
ential equations can be expressed as time-domain block dia-
grams in which a chain of integrator blocks yields progres-
sively lower-order derivatives of the output variable. Block
Fig. P9.1   RLC circuit
diagrams with this structure reveal a technique for writing a
higher-order differential equation without differentiation of
the input in state-space. Time-domain block diagram repre-
senting system equations with differentiation of the input in Problems
state-space are more complicated due to the need to redefine
the state variables so as to eliminate the differentiation of Problem 9.1 An electric circuit consisting of a voltage
the input. source, a resistor, R = 1 Ω, an inductor, L = 2 H, and a ca-
pacitor, C = 3 F, is shown in the schematic, Fig. P9.1. The
circuit is de-energized for time, t < 0.
Transfer Functions and Laplace-Domain Block 9.1.a Formulate the state equations for this system and
Diagrams express them in standard vector-matrix form.
9.1.b Determine the eigenvalues, the ideal, undamped natu-
Transfer functions created from differential system equa- ral frequency ωn, the damped natural frequency ωd,
tions are used as operators in block diagrams. A block dia- and the damping ratio ζ.
gram with a single input and a single output can be reduced 9.1.c Use the Laplace transformation to solve the state
to a single transfer function. The properties of transfer func- equations for a vector of the transfer functions of the
tions derived by reducing block diagrams are the same as state variables.
those derived directly from differential system equations. 9.1.d Plot the poles and zeros (if any) of the transfer func-
Specifically, the denominator of the transfer function is tions for the voltage across the capacitor and the cur-
the characteristic function of the system represented by the rent through the inductor on the s-plane.
block diagram and, when set equal to zero, is the character- 9.1.e Assuming the circuit is initially de-energized, use
istic equation. Closed-loop feedback control is described by Laplace transforms and partial fraction expan-
a block diagram in which the output variable is fed back and sion, if necessary, to determine the voltage across
subtracted from the input to yield the error signal. The roots the capacitor and the current through the induc-
of the denominator of a closed-loop system transfer function tor, when the input voltage applied to the circuit is
are the systems eigenvalues. v(t ) = 24 VDC us (t ) + 48 VDC us (t )(1 − e −0.25t ) .
9.1.f Plot the input and the responses using Mathcad or
MATLAB.
s-plane
Problem 9.2 A translational mechanical system consisting
The roots of the numerator and denominator of a transfer of a mass M sliding on a lubricating fluid film with damping
function are called the zeros and poles of the system, respec- b is shown in Fig. P9.2. A spring K is attached between the
tively. The magnitude of the transfer function is zero when mass and ground. The system is at rest and relaxed, before
evaluated for a value of s which is a root of the numerator. the mass is acted upon by an applied force F( t). Two sets of
Roots of the denominator, i.e., the eigenvalues, yield an in- parameter values are tabulated in Table P9.2.
finite value of the magnitude of the transfer function. Plots
of the eigenvalues on the s-plane show their position relative
Table P9.2   Parameter values
to the real and imaginary axes and to one another. Since the
M K b
eigenvalues determine the homogeneous or natural response
of a system, the s-plane is used as a graphical design tool. Case I 3 2 1
Case II 1 2 3
554 9  Transfer Functions, Block Diagrams, and the s-Plane

x,v
C
F(t) K R1 I
M P(t)
vent to Pump
atmosphere R2
Fluid
Lubricating fluid Reservoir

Damping b
Fig. P9.4   A fluid system
Fig. P9.2   Force source acting on a spring-mass-damper system
9.3.a Derive the state equations of this system.
9.3.b Calculate the system’s eigenvalues for Cases I and II
Hydrodynamic
Shaft rigidly attached using the parameter values in Table P9.3. If the sys-
bearing damping b2 tem is underdamped, determine the ideal, undamped
Flywheel natural frequency ωn, the damped natural frequency
rotational inertia J
ωd, and the damping ratio ζ.
Compliant shaft 9.3.c Use the Laplace transform to solve the state equations
torsion spring K
to yield the transfer functions for the state variables
Fluid coupling
damping b1 X 1 (s) X (s)
and 2 .
U (s) U (s)
Angular velocity 9.3.d Use Laplace transforms and partial fraction expan-
input Ω(t)
sion, if necessary, to determine the responses of
the state variables to the angular velocity input
Ω (t ) = 200us (t ) + 160(1 − e − 0.25t ).
Fig. P9.3   A rotational mechanical system
9.3.e Plot the input and the responses using Mathcad or
MATLAB.
9.2.a Derive the state equations of this system and express
them in vector-matrix form. Problem 9.4 A schematic of a hydraulic system is shown in
9.2.b Calculate the system’s eigenvalues for Cases I and Fig. P9.4. The pump, modeled as a pressure source p( t), dis-
II. If the system is underdamped, determine the ideal, charges fluid into a hydraulic circuit consisting of two fluid
undamped natural frequency ωn, the damped natural resistances, R1 and R 2 , a long run of pipe with inertance, I ,
frequency ωd, and the damping ratio ζ. and a fluid accumulator with capacitance, C.
9.2.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables Table P9.4   Parameter values
X 1 (s) X (s) R1 R2 I C
and 2 .
U (s) U (s) Case I 0.5 10 1 2
9.2.d Use Laplace transforms and partial fraction expansion,
Case II 2 10 3 4
if necessary, to determine the responses of the state
variables to the force input F (t ) = 10us (t ) + 8e − 0.5t .
9.2.e Plot the input and the responses using Mathcad or 9.4.a Derive the state equations of this system and express
MATLAB. them in vector-matrix form.
9.4.b Calculate the system’s eigenvalues for Cases I and II
Problem 9.3 The rotational mechanical system shown in using the parameter values in Table P9.4. If the sys-
Fig. P9.3 consists of an angular velocity Ω(t) which drives tem is underdamped, determine the ideal, undamped
a compliant shaft with torsion spring constant K through a natural frequency ωn, the damped natural frequency
fluid coupling with damping b1. The shaft drives a flywheel ωd, and the damping ratio ζ.
with inertia J. The hydrodynamic bearing of the flywheel are 9.4.c Use the Laplace transform to solve the state equations
modeled as a drag cup with damping b2. to yield the transfer functions for the state variables
X 1 (s) X 2 (s)
Table P9.3   Parameter values and .
U (s) U (s)
J K b1 b2 9.4.d Use Laplace transforms and partial fraction expan-
Case I 3 20 50 0.2 sion, if necessary, to determine the responses
Case II 1 200 5 0.5 of the state variables to the pressure input
p (t ) = 1, 000us (t ) + 1, 600(1 − e − 0.25t ).
Problems 555

Fig. P9.5   Translational mechani- Mass M


cal system

Lubricating Film Spring K2


Damping b 2

Pivot Spring K1
Dashpot b1 L3

L2

Force Source F(t)


L1

Fig. P9.6   Rotational mechanical Flywheel


system Compliant Shaft Rotational Inertia J1
Spring Constant K
Flywheel
Pulley Inertia J 2
Diameter D1 Pulleys Diameters
D2 and D3

Angular
Velocity
Input Ω(t)

Pulley Fluid Coupling


Diameter D4 Damping b

9.4.e Plot the input and the responses using Mathcad or 9.5.b Calculate the system’s eigenvalues using the parame-
MATLAB. ter values in Table P9.5. If the system is underdamped,
determine the ideal, undamped natural frequency ωn,
Problem 9.5 A translational mechanical system is shown the damped natural frequency ωd, and the damping
schematically in Fig. P9.5. The system consists of a force ratio ζ.
source F( t) acting on a lever, to which are attached a damper 9.5.c Use the Laplace transform to solve the state equations
b1 and two springs, K1 and K2. Spring K2 is attached to mass to yield the transfer functions for the state variables
M which slides on a rail with a lubricating film, damping X 1 (s) X 2 (s) X (s)
, and 3 .
b2. The parameter values are tabulated. The system is de- U (s) U (s) U (s)
energized before the input force act on the system. 9.5.d Use Laplace transforms and partial fraction expan-
sion, if necessary, to determine the responses of the
Table P9.5   Parameter values state variables to the force input, F (t ) = 400 us (t ).
L1 L2 L3 b1 b2 K1 K2 M 9.5.e Plot the input and the responses using Mathcad or
MATLAB.
1 2 1 1 0.5 1,000 1,500 10

Problem 9.6 A rotational mechanical system is shown sche-


9.5.a Derive the state equations of this system. matically in Fig. P9.6. The system consists of an angular
velocity source which acts on the input shaft of a belt drive
556 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. P9.7   Hybrid electrome- Electric Motor


chanical system Rotational Inertia J1
Transduction α
Bearings Damping b1

Voltage Source v(t)


Compliant Shaft
Spring Constant K

Electrical Resistance R

Rotational Inertia J2
Bearings Damping b2

system with two flywheels with inertias J1 and J2, a com- Table P9.7   Parameter values
pliant shaft modeled as torsion spring K, and fluid coupling R α J1 b1 K J2 b2
with damping b. Note that pulleys D2 and D3 are fixed to the 2 8 3 0.6 1,000 7 0.5
same shaft. The system is de-energized before the input force
act on the system.
9.7.a Derive the state equations of this system.
Table P9.6   Parameter values 9.7.b Calculate the system’s eigenvalues using the parame-
D1 D2 D3 D4 K J1 J2 b ter values in Table P9.7. If the system is underdamped,
2 4 8 6 100 7 5 30 determine the ideal, undamped natural frequency ωn,
the damped natural frequency ωd, and the damping
ratio ζ.
9.6.a Derive the state equations of this system. 9.7.c Use the Laplace transform to solve the state equations
9.6.b Calculate the system’s eigenvalues using the parame- to yield the transfer functions for the state variables
ter values in Table P9.6. If the system is underdamped, X 1 (s) X 2 (s) X (s)
, and 3 .
determine the ideal, undamped natural frequency ωn, U (s) U (s) U (s)
the damped natural frequency ωd, and the damping 9.7.d Use Laplace transforms and partial fraction expan-
ratio ζ. sion, if necessary, to determine the responses of the
9.6.c Use the Laplace transform to solve the state equations state variables to the voltage input, v(t ) = 96 us (t )
to yield the transfer functions for the state variables 9.7.e Plot the responses using Mathcad or MATLAB.
X 1 (s) X 2 (s) X (s)
, and 3 .
U (s) U (s) U (s) Problem 9.8 Draw a block diagram which represents the
9.6.d Use Laplace transforms and partial fraction expan- state equations and output equations below.
sion, if necessary, to determine the responses of
the state variables to the angular velocity input, d  x1   a11 a12   x1  b1 
Ω (t ) = 200 us (t ).  x  = a + u
dt  2   21 0   x2   0 
9.6.e Plot the responses using Mathcad or MATLAB.

Problem 9.7 A hybrid electromechanical system is shown  y1   c11 0   x1 


in Fig. P9.7. The input to the system is a voltage which  y  = c  
 2   21 c22   x2 
drives a DC motor through wires with electrical resistance
R. The relationship between the motor’s current and torque
is TM = iM . Flywheel J1 is attached to the motor’s shaft and Problem 9.9 A system equation where u is the input variable
is supported by bearings with damping b1. Coupled to the and y is the output variable is given below.
motor’s shaft is a compliant shaft modeled as torsion spring
K. The compliant shaft drives flywheel J2 which is supported d 3y d 2y dy
by bearings with damping b2. 8u = 3 3
+ 33 2
+ 114 + 210 y
dt dt dt
Problems 557

9.9.a Express the solution of the system equation as a block 100


diagram.
9.9.b Derive the state equations from the block diagram 50
drawn for part a.
9.9.c Solve the state equations for the input u( t) shown in u(t) 0
Fig. P9.9. 0 0.5 1.0 1.5 2.0
9.9.d Plot the output, y( t), using Mathcad or MATLAB.
t, sec
-50

Problem 9.10  A system equation where u is the input vari-


-100
able, and y is the output variable is given below.
Fig. P9.9   Input u( t)
d 3y d 2y dy
12u = 2 3 + 29 2 + 273 + 620 y
dt dt dt
100
9.10.a Express the solution of the system equation as a
block diagram. 50
9.10.b Derive the state equations from the block diagram
drawn for part a. u(t) 0
0 0.5 1.0 1.5 2.0
9.10.c Solve the state equations for the input u( t) shown in t, sec
Fig. P9.10. -50
9.10.d Plot the output, y( t), using Mathcad or MATLAB.
-100
Problem 9.11 Determine the state equations and output
equations of the system expressed as a block diagram in Fig. P9.10   Input u( t)
Fig. P9.11.

Problem 9.12 For the system shown in the block diagram in Problem 9.13 A transfer function Y( s)/U( s) is shown as a
Fig. P9.12. block diagram in Fig. P9.13.
9.12.a Determine the transfer function Y( s)/U( s). Express the transfer function Y( s)/U( s) as a time-domain,
9.12.b Determine the poles and zeros of the transfer func- state-space block diagram, using the recursive algorithm of
tion and plot them on the s-plane. Sect. 9.4.2 to eliminate differentiation of the input variable.
9.12.c Use Laplace transforms and partial fraction expan-
sion to determine the response of this system to the Problem 9.14 A feedforward loop and feedback loop are
step input u (t ) = 4us (t ). shown as block diagrams in Fig. P9.14.
9.12.d Plot the response of the system in Mathcad or MAT- 9.14.a Derive the transfer functions, C( s)/R( s). Express the
LAB. transfer functions as proper ratios.

Fig. P9.11   Block diagram of u1


state and output equations α y1
Input γ
Output
u2 - __
β +
+ +
-
1 η
Input s y2
+
+
δ Output
κ

θ
ε +
y3
+
Output
- __
1 ρ
- s

φ
558 9  Transfer Functions, Block Diagrams, and the s-Plane

Fig. P9.12   Block diagram of a


system 1.5
u(t) + 1
__ 1
__ 1
__ y(t)
0.5s +_ _ _
s s s

28

32

Problem 9.17 The block diagram of a closed-loop system is


U(s) s+3
__________________ Y(s)
shown in Fig. P9.17.
2s 3 + 18s 2 + 56s + 64 9.17.a Determine the closed-loop transfer function
C( s)/R( s). Express the transfer functions as proper
Fig. P9.13   A transfer function
ratios.
9.17.b Use inverse Laplace transforms and partial fraction
expansion, if necessary, to derive the response of the
9.14.b Use inverse Laplace transforms to derive the responses systems to the step input, r (t ) = 8us (t ) .
of the systems to the step input, r (t ) = 12us (t ). 9.17.c Plot the responses in Mathcad or MATLAB.
9.14.c Plot the responses in Mathcad or MATLAB.
Problem 9.18 The closed-loop feedback control system,
Problem 9.15 A block diagram is shown in Fig. P9.15. shown in the block diagram Fig. P9.18, controls the angu-
9.15.a Determine the transfer function, C( s)/R( s). lar velocity of a flywheel, J = 2  kg  m2, supported on bear-
9.15.b Use inverse Laplace transforms and partial fraction ings with damping b = 0.05 N m rad/sec. In controls jargon, a
expansion to determine response of the system to the dynamic system placed under closed-loop control is known
step input, r (t ) = 18us (t ). as a “plant.” The transfer function, GP( s), in the block dia-
9.15.c Plot the response in Mathcad or MATLAB. gram is the transfer function of the flywheel system, where
the input is the applied torque, T( t), and the output is the an-
Problem 9.16 A closed-loop feedback system is shown as a gular velocity of the flywheel. Note that the angular velocity
block diagram in Fig. P9.16. of the flywheel is the output variable, Ωout( s), of the feedback
9.16.a Determine the closed-loop transfer function loop system.
C( s)/R( s). Express the transfer functions as proper 9.18.a Derive the transfer function that relates input force
ratios. T( s) to the angular velocity of the flywheel Ωout( s).
9.16.b Use inverse Laplace transforms and partial fraction The transfer function Ωout( s)/T( s) is the “plant” trans-
expansion, if necessary, to derive the response of the fer function, GP( s), in the block diagram.
systems to the step input, r (t ) = 12us (t ) . 9.18.b Determine the closed-loop transfer function which
9.16.c Plot the responses in Mathcad or MATLAB. relates the input, the desired or reference value of
angular velocity of the flywheel, Ωin( s), to the output,

Fig. P9.14   a Feedforward loop,


b Feedback loop
a b
R(s) 9
_____ C(s) R(s) 9
_____ C(s)
+ +
s + 12 + - s + 12

3
_____ 3
_____
s+7 s+7

Fig. P9.15   A cascade of two


blocks
C s+30
_________ 192
______________ R
2
s +4s +29 2
12s +168s +576
Problems 559

Flywheel
R 30
_____________ C Hydraulic Motor
+- 7 Inertia J
3
4s +20s2 +60s High pressure line
Fluid Resistance R

1
____
s+8
Shaft Bearings
Fig. P9.16   A closed loop feedback system Damping b

the actual or observed angular velocity of the fly- Low pressure return
wheel, Ωout( s).
9.18.c The flywheel was at rest before the system was given Fluid power unit
modeled as a
the input, Ω in (t ) = 200 rad/sec us (t ). Determine the pressure source
response of the closed-loop system
9.18.d Plot the response of the system using either Mathcad Fig. P9.19a   The “plant” to be placed under feedback control
or MATLAB.

Problem 9.19 The hybrid system shown in Fig. P9.19a is the 9.19.c Determine the response of the de-energized system
“plant” to be placed under closed loop control per the block to the input Ω in (t ) = 300 rad/sec us (t ).
diagram in Fig. P9.19b. The plant consists of a fluid power 9.19.d Plot the response using either Mathcad or MATLAB.
unit modeled as a pressure source, p( t), hydraulic lines with
fluid resistance R = 5 GPa sec/m3 on the high pressure line, a Problem 9.20 An electromechanical system is shown in
hydraulic motor with displacement vol = 24 cm3 per revolu- Fig.  P9.20a. The system consists of a voltage source v( t)
tion and a flywheel J = 10 kg m2 supported on bearing with driving a DC motor with electrical resistance R = 3 Ω and
damping b = 0.05 Nm rad/sec. The output variable is the an- torque constant KT = 7 N ⋅ m / amp. The motor’s shaft drives
gular velocity of the flywheel. a flywheel, J = 1.5 kg ⋅ m 2 , and is coupled to a power screw
9.19.a Derive the plant transfer function which relates the with four threads per inch. The nut of the power screw
input pressure P( s) to the angular velocity of the fly- moves mass M  = 50 kg which slides on a lubricating film
wheel Ωout( s). If there is a negative (sign inversion) with damping b = 6 N ⋅ sec/m. The position of mass M is to
between the input and output variables, it results be placed under closed-loop, feedback control per the block
from the linear graph transducer sign convention and diagram in Fig. P9.20b.
must be removed. A positive input to the plant must 9.20.a Derive the plant transfer function which relates the
yield a positive output. input, the voltage applied to the DC motor to output,
9.19.b Determine the closed-loop transfer function, the velocity of mass M. If there is a negative (sign
Ωout( s)/Ωin( s). inversion) between the input and output variables,

Fig. P9.17   A closed-loop feed-


back system
R +-
4s+6
________________ C
s3 +11s 2+124s+300

20

Fig. P9.18  a Block diagram of


the closed-loop feedback control
a b Flywheel
Rotational Inertia J
system. b The flywheel which
is the “plant” with its angular “Plant”
velocity under feedback control Ω in (s) Ω out (s)
+
- 8 GP(s) Bearings
Damping b

___
7 Input Torque
s+2 T(t)
560 9  Transfer Functions, Block Diagrams, and the s-Plane

“Plant” The procedure is as follows. First, express the denominator


Ω in (s) P(s) Ω out (s)
+
- 4(s+10) GP(s) polynomial in standard form. Clear the highest power of s of
its coefficient
8 1
N (s)
a0
Fig. P9.19b   Block diagram of the closed-loop feedback control system Output ( s ) =
a1 a2 a3 a
s 4 + s3 + s 2 + s + 4
a0 a0 a0 a0
it results from the linear graph transducer sign con-
vention and must be removed. A positive input to the Now, factor the denominator polynomial.
plant must yield a positive output. If there is an oscillatory second-order factor, it is gener-
9.20.b Determine the closed-loop transfer function, ally best to keep it as a second-order polynomial. Conse-
xMass( s)/xin( s). quently, a fourth-order characteristic function yields one of
9.20.c Determine the response of the system to the input, three possible forms
xin (t ) = 0.5 m us (t ).
9.20.d Plot the response using either Mathcad or MATLAB. 1
N (s)
a0
a1 3 a 2 2 a 3 a
Chapter 9 Appendix s4 + s + s + s+ 4
a0 a0 a0 a0

Inverse Laplace Transformation Using Manual 1


N (s)
Partial Fraction Expansion a0
=
The process of partial fraction expansion requires knowl- ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
edge of what order polynomial to assume for the numera-
tors of expansion terms, and a fair amount of patience while or
expanding polynomials, collecting like terms, and forming a 1
set of simultaneous algebraic equations, to solve for the coef- N (s)
a0
ficients of the numerator polynomials.
To find the time-domain variable, output( t), described by a1 3 a 2 2 a 3 a
s4 + s + s + s+ 4
the Laplace-domain signal, Output( s) a0 a0 a0 a0
1
N (s) N ( s)
Output ( s ) = a0
a 0 s + a 1 s + a 2 s 2 + a 3 s + a4
4 3
=
( s + p1 )( s + p2 ) ( s 2 + 2ζw n s + w n2 )

Fig. P9.20a   The electromechan- DC Motor Flywheel, Rotational Inertia J


ical “plant” to be placed under Torque constant K T
feedback control

Electrical Resistance R Power Screw


Mass M

Voltage Source v(t) Lubricating Film


Damping b

Fig. P9.20b   The block diagram Controller Integrator


of the feedback control system “Plant”
x in (s) vMotor(s) vMass(s) 1 x Mass(s)
__
+
- 4(s+10) GP(s)
s

8
Chapter 9 Appendix 561

or set of equations is expressed in matrix notation. The vector


of unknown constants is found by using linear algebra. As
1
N (s) illustrated below, most of the effort in partial fraction ex-
a0
pansion lies in expanding polynomials, which can be done
a1 3 a 2 2 a 3 a mechanically. A calculator or computer is needed to evaluate
s4 + s + s + s+ 4
a0 a0 a0 a0 the linear algebra.
1 The general case will be illustrated, assuming all of the
N (s) poles are first order and none are repeated.
a0
= 1. Factor the denominator in to first- and second-order
(s 2
)(
+ 2ζ1w1n s + w12n s 2 + 2ζ 2w 2 n s + w 22n ) polynomials, where the second-order factors have
complex conjugate roots
The next step is to express the Laplace transformed variable
(or signal) as the sum of terms, with the individual factors 1
N (s)
determined in step one as denominators. A polynomial is as- a0
sumed for the numerator, which is one order lower than the a1 3 a 2 2 a 3 a4
s4 + s + s + s+
denominator. If the denominator is a first-order polynomial, a0 a0 a0 a0
then assume a constant for the numerator
1
N (s)
1 a0
N (s) =
a0 ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
2. Equate the factored form with the partial fraction expan-
A B C D
= + + + sion, where the numerator terms are polynomials in s of
s + p1 s + p2 s + p3 s + p4 one order less than the denominator, i.e.,
or
1 D ( s ) = s 2 + 2n s + 2n → N ( s ) = As + B
N (s)
a0
and
( s + p1 )( s + p2 ) ( s 2 + 2ζw n s + w n2 )
A B Cs + D D (s) = s + a → N (s) = A
= + + 2
(
( s + p1 ) ( s + p2 ) s + 2ζw n s + w n2 )
In this example,
or
1
1 N (s)
N (s) a0
a0
a1 3 a 2 2 a 3 a4
(s 2
)(
+ 2ζ1w1n s + w12n s 2 + 2ζ 2w 2 n s + w 22n ) s4 +
a0
s +
a0
s +
a0
s+
a0
As + B Cs + D A B C D
= +
( s 2
+ 2ζ1w1n s + w12n ) ( s + 2ζ 2w 2 n s + w 22n
2
) = + + +
s + p1 s + p2 s + p3 s + p4

The simplest way to determine the unknown constants, A, 3. Multiply both sides by the denominator of the left side to
B, C, and D, is to multiply both sides of the equation by the clear fractions. Recall that
denominator of the left side, clearing fractions and yielding
a1 a2 a3 a4
a polynomial in the Laplace variable, s. Equating the coef- D (s) = s4 + s3 + s2 + s+
ficient terms of like powers of s produces a set of equations a0 a0 a0 a0
equal in number to the number of unknown constants. The = ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )

N (s)
( s + p1 )( s + p2 )( s + p3 )( s + p4 )
a0  A B C D 
= + + + ( s + p1 )( s + p2 )( s + p3 )( s + p4 )
( s + p1 )( s + p2 )( s + p3 )( s + p4 )  s + p1 s + p2 s + p3 s + p4 
562 9  Transfer Functions, Block Diagrams, and the s-Plane

N (s)  A  4. Expand the products of the factors into polynomials. This


=  ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 ) is the tedious part. It saves time to work one expansion
a0  s + p1  using generalized variables. Then use the result as a tem-
 B  plate
+  ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
 s + p2  ( s + a ) ( s + b) ( s + c) = s3 + (a + b + c) s 2
 C  + (a ⋅ b + b ⋅ c + a ⋅ c) s + (a ⋅ b ⋅ c)
+  ( s + p1 )( s + p2 ) ( s + p3 ) ( s + p4 )
 s + p3 
N (s)
 D  = A( s + p2 )( s + p3 )( s + p4 ) + B ( s + p1 )( s + p3 )( s + p4 )
a0
+  ( s + p1 )( s + p2 ) ( s + p3 ) ( s + p4 )
 s + p4  + C ( s + p1 )( s + p2 )( s + p4 ) + D ( s + p1 )( s + p2 )( s + p3 )

N (s)
a0
(
= A s 3 + ( p2 + p3 + p4 ) s 2 + ( p2 ⋅ p3 + p3 ⋅ p4 + p2 ⋅ p4 ) s + ( p2 ⋅ p3 ⋅ p4 ) )
(
+ B s 3 + ( p1 + p3 + p4 ) s 2 + ( p1 ⋅ p3 + p3 ⋅ p4 + p1 ⋅ p4 ) s + ( p1 ⋅ p3 ⋅ p4 ) )
(
+ C s + ( p1 + p2 + p4 ) s + ( p1 ⋅ p2 + p2 ⋅ p4 + p1 ⋅ p4 ) s + ( p1 ⋅ p2 ⋅ p4 )
3 2
)
(
+ D s + ( p1 + p2 + p3 ) s + ( p1 ⋅ p2 + p2 ⋅ p3 + p1 ⋅ p3 ) s + ( p1 ⋅ p2 ⋅ p3 )
3 2
)
N (s)
= As 3 + A ( p2 + p3 + p4 ) s 2 + A ( p2 ⋅ p3 + p3 ⋅ p4 + p2 ⋅ p4 ) s + A ( p2 ⋅ p3 ⋅ p4 )
a0
+ Bs 3 + B ( p1 + p3 + p4 ) s 2 + B ( p1 ⋅ p3 + p3 ⋅ p4 + p1 ⋅ p4 ) s + B ( p1 ⋅ p3 ⋅ p4 )
+ Cs 3 + C ( p1 + p2 + p4 ) s 2 + C ( p1 ⋅ p2 + p2 ⋅ p4 + p1 ⋅ p4 ) s + C ( p1 ⋅ p2 ⋅ p4 )
+ Ds 3 + D ( p1 + p2 + p3 ) s 2 + D ( p1 ⋅ p2 + p2 ⋅ p3 + p1 ⋅ p3 ) s + D ( p1 ⋅ p2 ⋅ p3 )

5. Collect like terms of s

N (s)
a0
(
= ( A + B + C + D ) s 3 + A ( p2 + p3 + p4 ) + B ( p1 + p3 + p4 ) + C ( p1 + p2 + p4 ) + D ( p1 + p2 + p3 ) s 2 )
(
+ A ( p2 ⋅ p3 + p3 ⋅ p4 + p2 ⋅ p4 ) + B ( p1 ⋅ p3 + p3 ⋅ p4 + p1 ⋅ p4 ) + C ( p1 ⋅ p2 + p2 ⋅ p4 + p1 ⋅ p4 )
+ D ( p1 ⋅ p2 + p2 ⋅ p3 + p1 ⋅ p3 ) s )
(
+ A ( p2 ⋅ p3 ⋅ p4 ) + B ( p1 ⋅ p3 ⋅ p4 ) + C ( p1 ⋅ p2 ⋅ p4 ) + D ( p1 ⋅ p2 ⋅ p3 ) )

N (s)
= A( s + p2 )( s + p3 )( s + p4 ) + B ( s + p1 )( s + p3 )( s + p4 ) 6. Create a set of simultaneous equations by equating like
a0
N ( s ) b0 3 b1 2 b2 b3
+ C ( s + p1 )( s + p2 )( s + p4 ) + D ( s + p1 )( s + p2 )( s + p3 ) terms of s, where = s + s + s+ .
a0 a0 a0 a0 a0

b0
s3 = A+ B +C + D
a0
b1
s2 = ( p2 + p3 + p4 ) A + ( p1 + p3 + p4 ) B + ( p1 + p2 + p4 )C + ( p1 + p2 + p3 ) D
a0
Chapter 9 Appendix 563

b2
s1 = ( p2 ⋅ p3 + p3 ⋅ p4 + p2 ⋅ p4 ) A + ( p1 ⋅ p3 + p3 ⋅ p4 + p1 ⋅ p4 ) B + ( p1 ⋅ p2 + p2 ⋅ p4 + p1 ⋅ p4 ) C
a0
+ ( p1 ⋅ p2 + p2 ⋅ p3 + p1 ⋅ p3 ) D
b3
s0 = ( p2 ⋅ p3 ⋅ p4 ) A + ( p1 ⋅ p3 ⋅ p4 ) B + ( p1 ⋅ p2 ⋅ p4 ) C + ( p1 ⋅ p2 ⋅ p3 ) D
a0

7. Express in matrix notation

 b0 
a 
 0  1 1 1 1 
 b1  
   ( p2 + p3 + p4 ) ( p1 + p2 + p4 ) ( p1 + p3 + p4 ) ( p1 + p2 + p3 )   A 
 a0   B
  =  p2 ⋅ p3 + p3 ⋅ p4   p1 ⋅ p3 + p3 ⋅ p4   p1 ⋅ p2 + p2 ⋅ p4   p1 ⋅ p2 + p2 ⋅ p3    
  
 b2   + p ⋅ p   + p ⋅ p   + p ⋅ p   + p ⋅ p 
 
C
2 4 1 4 1 4 1 3
 a0   D
   ( p2 ⋅ p3 ⋅ p4 ) ( p1 ⋅ p3 ⋅ p4 ) ( p1 ⋅ p2 ⋅ p4 ) ( p1 ⋅ p2 ⋅ p3 )   
 b3 
a 
 0

8. Solve by premultiplying both sides by the inverse of the


coefficient matrix

 b0 
 
a
−1  0 
 1 1 1 1   
b
 p +p +p
 ( 2 3 4) ( p1 + p2 + p4 ) ( p1 + p3 + p4 ) ( p1 + p2 + p3 )   a 1   A 
 0 B
 p2 ⋅ p3 + p3 ⋅ p4   p1 ⋅ p3 + p3 ⋅ p4   p1 ⋅ p2 + p2 ⋅ p4   p1 ⋅ p2 + p2 ⋅ p3     =  
   + p ⋅ p   + p ⋅ p   + p ⋅ p 
  2  b  C 
 + p2 ⋅ p4 1 4 1 4 1 3     
a D

 ( p2 ⋅ p3 ⋅ p4 ) ( p1 ⋅ p3 ⋅ p4 ) ( p1 ⋅ p2 ⋅ p4 ) ( p1 ⋅ p2 ⋅ p3 )   0   
b3
 
 a 0 

The bad-old-days method can be quicker in some cases, but, factor, and then evaluating the equation for the value of s,
in general, is more involved, and only useful for those who which is the pole of that factor
stubbornly refuse to accept that their beloved middle-school
calculator is obsolete. The method uses algebra rather than 1
N (s)
linear algebra. The constants for first-order factors are de- a0
termined by multiplying both sides of the equation by the ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
A B C D
= + + +
s + p1 s + p2 s + p3 s + p4

1
( s + p1 ) a N (s)
( s + p1 ) A
0
( s + p1 ) B
0
( s + p1 ) C
0
( s + p1 ) D
0
= + + +
( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 ) s + p1
s = − p1
s + p2
s = − p1
s + p3
s = − p1
s + p4
s = − p1
s = − p1

1
N ( − p1 )
a0
=A
( − p1 + p2 )( − p1 + p3 )( − p1 + p4 )
564 9  Transfer Functions, Block Diagrams, and the s-Plane

The process is repeated for to determine the constants, B, C, s1 0 = A + B 0  1 1  A


and D. →  3 =  2 1  B 
s0 3 = 2 A + B     
The bad-old-days procedure to determine the constants
for a second-order factor is similar, except it is necessary −1 −1
to differentiate both sides of the equation with respect to s, 1 1 0 1 1 1 1  A
 2 1  3 =  2 1  2 1  B 
to create two equations, in order to determine the two un-         
knowns.
The special cases for either procedure involve repeated  −1 1  0  A  3   A
roots. The general case of repeated roots is not a practical  2 −1  3 = I  B  →  −3 =  B 
concern, since repeated roots require two physical subsys-         
tems to be identical, or for a second-order factor to be exact-
ly critically damped. Phenomena which require parameters 3 −3 3
G ( s) = = +
to have “exact” values are extremely unlikely. s 2 + 3s + 2 ( s + 1) ( s + 2)
However, there is one common situation which generates
repeated roots. Such systems have poles at the origin, be- Alternative Method
1
cause a factor, , in a transfer function represents integra-
s 1
tion. The Laplace transform of a unit step input is also . 3 3 A B
s G (s) = = = +
s 2 + 3s + 2 ( s + 1)( s + 2) ( s + 1) ( s + 2)
The Laplace transform of a unit ramp input is the integral of
a unit step. Hence, the unit ramp input is 3 A B
= +
( s + 1)( s + 2) ( s + 1) ( s + 2)
11 1
=
s s s2
3( s + 1) A( s + 1) B ( s +11)
= +
Factors which produce repeated poles at the origin are han- ( s + 1)( s + 2) ( s + 1) ( s + 2)
dled as second-order factors
3 ( s + 1) A ( s + 1) B( s + 1)
1 As + B As + B A s B 1 A B = +
= → = 2 + 2 → 2 = + 2 ( s + 1) ( s + 2) ( s + 1) ( s + 2)
2 2 2
s s s s s s s s
3 B( s + 1) 3
Partial fraction expansion will be illustrated with five ex- = A+ → =3= A
amples. ( s + 2) s = −1 ( s + 2) s = −1 ( −1 + 2)
Example One: Use partial fraction expansion to express the
3 3 A B
second-order transfer function, Output ( s ) = 2 , as = +
s + 3s + 2 ( s + 1)( s + 2) ( s + 1) ( s + 2)
the sum of two first-order transfer functions.
Linear Algebra Method 3( s + 2) A( s + 2) B( s + 2)
= +
( s + 1)( s + 2) ( s + 1) ( s + 2)
3 3 A B
G (s) = = = +
s 2 + 3s + 2 ( s + 1) ( s + 2) ( s + 1) ( s + 2)
3 ( s + 2) A( s + 2) B ( s + 2)
= +
3 ( s + 1) ( s + 2)  A B  ( s + 1) ( s + 2) ( s + 1) ( s + 2)
= +  ( s + 1) ( s + 2)
s 2 + 3s + 2  ( s + 1) ( s + 2) 

3 A( s + 2) 3
3 ( s + 1)( s + 2) A ( s + 1) ( s + 2) B( s + 1) ( s + 2) = +B → = −3 = B
= + ( s + 1) s = −2 ( s + 1) s = −2 ( −2 + 1)
s 2 + 3s + 2 ( s + 1) ( s + 2)
3 3 −3
G (s) = = +
( s + 1)( s + 2) ( s + 1) ( s + 2)
3 = A( s + 2) + B( s + 1) → 3 = ( As + 2 A) + ( Bs + B)
Example Two: The system shown in the block diagram
1
3 = ( A + B) s + (2 A + B) Fig. A9.1 is given a unit step input, U ( s ) = . Use partial
s
Chapter 9 Appendix 565

Fig. A9.1   Block diagram of sys- 1


tem showing a unit step input —
s 1 3s + 2 Y(s)
_____ _________
U(s) s+2 s 2 + 7s +12

fraction expansion and the Laplace Transform tables to de-


termine the time-domain response. (3s + 2) ( s ( s + 2)( s + 3)( s + 4))
The s-domain output signal is: s 4 + 9 s 3 + 26 s 2 + 24 s
A s ( s + 2)( s + 3)( s + 4)  Bs ( s + 2) ( s + 3)( s + 4) 
3s + 2 = + 
Y (s) = s  s+2 
(9.43) s s + 2 s 2 + 7 s + 12 ( )( )
 Cs ( s + 2) ( s + 3) ( s + 4)   Ds ( s + 2)( s + 3) ( s + 4) 
The inverse Laplace transform of this signal is not among +  + 
 s+3   s+4 
our Laplace transform pairs. We will use partial fraction ex-
pansion to expand this signal to match the table entries. We
have many possible combinations. Recall the numerator of a 3s + 2 = A ( s + 2) ( s + 3) ( s + 4) + Bs ( s + 3) ( s + 4)
partial expansion term must be a polynomial in s of one order + Cs ( s + 2) ( s + 4) + Ds ( s + 2) ( s + 3)
lower than the denominator. One partial fraction expansion
is
( ) (
3s + 2 = A s 3 + 9 s 2 + 26 s + 24 + B s 3 + 7 s 2 + 12 s )
Y (s) =
3s + 2
(
s ( s + 2) s + 7 s + 12
2
=
As + B
)
+ 2
Cs + D
s ( s + 2) s + 7 s + 12
( ) (
+ C s 3 + 6 s 2 + 8s + D s 3 + 5s 2 + 6 s )
( ) (
3s + 2 = As 3 + A9 s 2 + A26 s + A24 + Bs 3 + B7 s 2 + B12 s )
Another is
( 3
+ Cs + C 6 s 2
+ C 8s ) + ( Ds 3
+ D5s 2
+ D6s )
3s + 2 A B C D
Y (s) = = + + +
(
s ( s + 2) s + 7 s + 12
2
) s s+2 s+3 s+4 3s + 2 = ( A + B + C + D ) s 3 + ( A9 + B7 + C 6 + D5) s 2
+ ( A26 + B12 + C 8 + D6) s + A24

s3 0 = A + B + C + D  0  1 1 1 1  A 
s 2
0 = 9 A + 7 B + 6C + 5 D 0  9 7 6 5  B 
→  =
s1 3 = 26 A + 12 B + 8C + 6 D  3  26 12 8 6  C 
    
s0 2 = 24 A  2  24 0 0 0  D 

Because the roots of the second-order factor are real, we will 1 1 1 1


−1
 0
use the latter. 9 7 6
 5  0
 
Linear Algebra Method
 26 12 8 6  3
3s + 2    
Y (s) =  24 0 0 0  2
(
s ( s + 2) s 2 + 7 s + 12 ) 1 1 1 1
−1
1 1 1 1  A 
9 7 6 5 9 7 6 5  B 
3s + 2 A B C D = 
Y (s) = = + + +
s 4 + 9 s 3 + 26 s 2 + 24 s s s + 2 s + 3 s + 4  26 12 8 6  26 12 8 6  C 
    
 24 0 0 0  24 0 0 0  D 
(3s + 2) ( s ( s + 2) ( s + 3) ( s + 4))
s 4 + 9 s 3 + 26 s 2 + 24 s
A B C D 
= + +
 s s + 2 s + 3 s + 4 
+ ( s ( s + 2) ( s + 3) ( s + 4))
566 9  Transfer Functions, Block Diagrams, and the s-Plane

0 0 0 0.042   0 1 0 0 0  A  (3) ( −3) + 2 = −2.333 = C


 2 −1 0.5 −0.25   0 0
  =  1 0 0  B  ( −3) ( −3 + 2) ( −3 + 4)

 −9 3 −1 0.333   3 0 0 1 0  C 
     
 8 −2 0.5 −0.125  2 0 0 0 1  D  Find D. Multiply both sides by ( s + 4), and evaluate for
s = −4.
 0.083   A 
 1  B ( s + 4 ) ( 3s + 2 )
 =  =
( s + 4) A
 −2.333  C  s ( s + 2) ( s + 3) ( s + 4) s
    s = −4 s = −4
 1.25   D 
( s + 4) B ( s + 4) C ( s + 4) D
+ + +
3s + 2 0.083 1 2.333 1.25 s+2 s+3 s+4
Y (s) = = + − +
(
s ( s + 2) s 2 + 7 s + 12 ) s s+2 s+3 s+4
s = −4 s = −4 s = −4

(3) ( −4) + 2 = 1.25 = D


Alternative Method
Find A. Multiply both sides by s, and evaluate for s = 0.
( −4) ( −4 + 2) ( −4 + 3)
3s + 2 0.083 1 2.333 1.25
s ( 3s + 2 ) Y (s) = = + − +
s ( s + 2) ( s + 3) ( s + 4) s s+2 s+3 s+4
s ( s + 2) ( s + 3) ( s + 4)
s=0

sA sB sC sD Perform the inverse Laplace transformation to create the


= + + + time-domain function.
s s=0 s + 2 s=0
s+3 s=0
s+4 s=0

 0.083   1 
( 2) = 0.083 = A L {Y ( s )} = L
−1 −1

 s 
+ L −1
 
 s + 2
( 2)(12)
− 1  2.333   1.25 
− L 
 s+3 
+ L −1
 
 s + 4
Find B. Multiply both sides by ( s + 2), and evaluate for
s = −2.
1 −1  1 
L {Y ( s )} = 0.083L
−1 −1
 +
s
L  
 s + 2
( s + 2 ) ( 3s + 2 ) ( s + 2) A
= −1  1  −1  1 
s ( s + 2) ( s + 3) ( s + 4)
s = −2
s
s = −2
− 2.333 L 
 s + 3
 +1.25

L 
 s + 4
( s + 2) B ( s + 2) C ( s + 2) D
+ + + y (t ) = 0.083 + e −2t − 2.333e −3t + 1.25e −4t
(9.44)
s+2 s+3 s+4
s = −2 s = −2 s = −2

Example Three: The system shown in the block diagram


(3) ( −2) + 2 =1= B Fig. A9.3 consists of an integrator followed by an over-
( ) ( −2 + 3) ( −2 + 4)
−2 damped second-order system. The system is given a unit
1
step input of U ( s ) = . Use partial fraction expansion and
Find C. Multiply both sides by ( s + 3), and evaluate for s
s = −3. the Laplace Transform table to determine the time-domain
response.
The s-domain output signal is
( s + 3) (3s + 2) ( s + 3) A
=
s ( s + 2) ( s + 3) ( s + 4) s  1 1 
Y ( s) =  2  
 s   ( s + a ) ( s + b ) 
s = −3 s = −3

( s + 3) B ( s + 3) C ( s + 3) D
+ + +
s+2 s+3 s+4
s = −3 s = −3 s = −3
Chapter 9 Appendix 567

0.15 Equate like powers of s

s3 : 0 = B+C+ D
0.10
s2 : 0 = A + 4 B + 3C + D
y(t) s1 : 0 = 4 A + 3B
0.05 s0 : 1 = 3 A.
Express the system of equations in vector-matrix form
0.00
0 1 2 3 4 5 0  0 1 1 1  A 
t, sec 0  1 4 3 1   B 
 =
Fig. A9.2   Unit step response Eq. 9.44 0  4 3 0 0  C 
    
1   3 0 0 0  D 

We will assign the values, a = 1 and b = 3. The inverse La- Solve by pre-multiplying both sides of the equation by the
place transform of this signal is not in our table. Use partial inverse of the coefficient matrix
fraction expansion to expand this signal to match the table −1 −1
entries 0 1 1 1  0  0 1 1 1 0 1 1 1  A 
1 4 3 1   0  1 4 3 1  1 4 3 1   B 
 1 1  A B C D   = 
Y ( s) =  2    = 2+ + + 4 3 0 0  0  4 3 0 0 4 3 0 0  C 
 s   ( s + a ) ( s + b) s s s+a s+b         
3 0 0 0 1   3 0 0 0 3 0 0 0  D 

Linear Algebra Method


First, clear fractions by multiplying both sides by the denom- 0 1 1 1
−1
 0 1 0 0 0  A   A 
inator of the left side. 1
 4 3 1   0  0
 = 1 0 0  B   B 
=
4 3 0 0  0  0 0 1 0  C   C 
s 2 ( s + 1) ( s + 3)         
Y ( s) = = 3 0 0 0 1   0 0 0 1  D   D 
s 2 ( s + 1) ( s + 3)
 0.333   A 
A s 2 ( s + 1) ( s + 3) Bs 2 ( s + 1) ( s + 3)  −0.444  B 
+  = 
s2 s
 0.5   C 
Cs 2 ( s + 1) ( s + 3) Ds 2 ( s + 1) ( s + 3)    
 −0.056  D 
+ +
s +1 s+3
Conventional Algebra Method (The Hard Way)
Expand and collect like powers of s Find A. Multiply both sides by s2, and evaluate using s = 0.

 1 1   A B C D 
1 = A( s + 1) ( s + 3) + Bs ( s + 1) ( s + 3) + Cs 2 ( s + 3) + Ds 2 ( s + 1) s2  2    = s2  2 + + + 
 s   ( s + 1) ( s + 3)  s s s + 1 s + 3

( ) (
1 = A s 2 + A 4 s + A 3 + Bs 3 + B 4 s 2 + B3s )  1  1 
+ (Cs + Cs 3) + ( Ds s2  2  
3 2 3
+ Ds 2 )  s   ( s + 1) ( s + 3) 
s=0
=

1 = ( B + C + D ) s 3 + ( A + 4 B + 3C + D ) s 2  B  C   D 
A + s2   + s2  + s2 
 s  s + 1  s + 3 
+ ( 4 A + 3 B ) s + ( 3 A) 0 s=0 0 s=0 0 s=0

Fig. A9.3   Block diagram of an 1


overdamped second-order system —
s 1 Y(s)
with an integrator. The input is a
1
__ _________
unit step U(s) s (s+a)(s+b)
568 9  Transfer Functions, Block Diagrams, and the s-Plane

1 2.0
= 0.333 = A
3
1.5
Find B. The technique used to find A does not work to find
1 y(t) 1.0
B. Multiplying by s leaves on the left side, and an s in the
s
denominator of the A term
0.5

 1 1 
0
s 2  
 s   ( s + 1) ( s + 3)  0 1 2 3 4 5 6
t, sec
 A  B  C   D 
= s 2  + s  + s + s
s   s  s + 1  s + 3  Fig. A9.4   Response function Eq. 9.45

This expression cannot be evaluated using s = 0 to find B,


1 1   A B C D 
because two terms blow up to infinity ( s + 1)  2 
= ( s + 1)  2 + + + 
s   ( s + 1) ( s + 3)  s s s + 1 s + 3

 1  1 
  
s  ( s + 1)( s + 3)   1 1   A  B  D 
 2    = ( s + 1)  2  + ( s + 1)   + C + ( s + 1)  
s=0 s s + 3 s s s + 3

 A  C   D 
=   + B + s  + s   1  1   A  B
s  ( s + 1)   ( s + 3)   2    = ( s + 1)  2  + ( s + 1)  
s=0 0 s=0 0 s=0
s s + 3 s = −1
s   s
0 s = −1 0 s = −1

The resolution to this dilemma is to multiply both sides by s , 2


 D 
and then differentiate with respect to s + C + ( s + 1) 
 s + 3 
0 s = −1
 1
2 1   A B C D 
s  2 = s2  2 + + + 
 s   ( s + 1) ( s + 3)  s s s + 1 s + 3  1  1  1
 2   = = 0.5 = C
 ( −1)   −1 + 3  2
1 Cs 2 Ds 2
= A + Bs + +
( s + 1) ( s + 3) s +1 s + 3
Find D. Multiply both sides by ( s + 3), and evaluate using
s = −3
d  1  d  Cs 2 Ds 2 
  =  A + Bs + +
ds  ( s + 1) ( s + 3)  ds  s + 1 s + 3  1 1   A B C D 
( s + 3)  2 
= ( s + 3)  2 + + + 
s   ( s + 1) ( s + 3)  s s s + 1 s + 3
du dv
v −u
d  u ds ds and ( s + 1)( s + 3) = s 2 + 4 s + 3
using  =  1 1   A  B  C 
ds  v  v2
 2    = ( s + 3)  2  + ( s + 3)   + ( s + 3)  +D
s s + 1 s s s + 1
d  1  d  Cs 2 Ds 2 
  = A + Bs + +
ds  s 2 + 4 s + 3  ds  s + 1 s + 3   1  1   A
= ( s + 3)  2 
 B
+ ( s + 3)  
 2    s   s
s s + 1 s = −3 0 s = −3 0 s = −3
− ( 2 s + 4) ( s + 1) 2Cs − Cs 2
( s + 3) 2 Ds − Ds 2

= B+ +  C 
+ ( s + 3)  +D
(s ) ( s + 1)2 ( s + 3)2
2
 s + a 
2
+ 4s + 3
0 s = −3

Now evaluate both sides using s = 0  1  1  1


 2   = − = −0.056 = D
− (1 + 3) 4  ( −3)   −3 + 1 18
=− = −0.444 = B
(3) 2
9
 1 1  0.333 0.444 0.5 0.056
Y ( s) =  2   = 2 − + −
Find C. Multiply both sides by ( s +1) , and evaluate using  s   ( s + 1) ( s + 3)  s s s +1 s + 3
s = −1
Chapter 9 Appendix 569

Fig. A9.5   Block diagram of 1


underdamped second-order sys- — Y(s)
tem with an integrator. The input
s 1
__ ω2n
_____________
is a unit step U(s) s s 2+ 2ζωns + ω2n

The output signal is transformed to the time-domain by the The inverse Laplace transform is not in our transform pair
inverse Laplace transform table. Use partial fraction expansion to expand this to match
1
the table entries. The repeated root of 2 requires two terms,
 0.333 0.444 0.5 0.056  A B s
y (t ) = L −1
{Y ( s)} = L −1
 2 − + − 
s 2
and . The numerator of the oscillatory second-order fac-
s
 s s s +1 s + 3 
tor must be a first-order polynomial, Cs + D.

The Laplace transform is a linear operator, so it has the dis-  1  n2  A B Cs + D


tributive property Y (s) =  2   2 = + +
 s   s + 2 n s + w n2  s 2 s s 2 + 2 n s +  n2

 0.333   0.444  −1  0.5 


y (t ) = L −1  2  − L −1   +L   This example will be easier to follow if we assign values to ζ
 s   s   s + 1
and  n. Using  = 0.3 and  n = 4,
 0..056 
− L −1  
 s+3   1 16  A B Cs + D
Y ( s) =  2   2 = + + .
 s   s + 2.4 s + 16  s 2 s s 2 + 2.4 s + 16
Its linearity allows coefficients to be factored out
This partial fraction expansion is easiest to work using linear
1 1  1  algebra. We will first create a set of simultaneous algebraic
y (t ) = 0.333L −1  2  − 0.444L −1   + 0.5L −1   equations for A, B, C, and D, which we will solve using lin-
s
  s
   s + 1
ear algebra. We will then rework the problem finding un-
 1 
− 0.056L −1   knowns individually to illustrate the difficulties created by
 s + 3 1
the repeated root, 2 .
s
The inverse Laplace transform of these functions of s are Linear Algebra
found in the transform pair table First, clear fractions by multiplying both sides by the denom-
inator of the left-hand side
 y (t ) = 0.333uramp (t ) − 0.444ustep (t ) + 0.5e − t − 0.056e −3t

(9.45)
(
16 s 2 s 2 + 2.4 s + 16 ) (
A s 2 s 2 + 2.4 s + 16 )
=
2
( 2
s s + 2.4 s + 16 ) s 2

Example Four: The system shown in the block diagram


Fig. A9.5 consists of an integrator followed by an oscilla-
+
(
Bs 2 s 2 + 2.4 s + 16 ) + (Cs + D ) s ( s 2 2
+ 2.4 s + 16 )
tory second-order system. Use partial fraction expansion and s 2
s + 2.4 s + 16
the Laplace transform pairs to determine the time-domain
response.
16 = A( s 2 + 2.4 s + 16) + B( s 3 + 2.4 s 2 + 16 s ) + (Cs + D) s 2
The algebra in this reduction entails considerably more
effort than in Example Three, because the roots of an oscil-
latory second-order factor are complex conjugates. Partial 16 = ( As 2 + 2.4 As + 16 A) + ( Bs 3 + 2.4 Bs 2 + 16 Bs )
fraction coefficients will be real numbers, except if the input + (Cs 3 + Ds 2 )
is a sinusoid. In this example, the input is a unit step. Conse-
quently, A, B, C, and D will be real. 16 = ( B + C ) s 3 + ( A + 2.4 B + D) s 2 + (2.4 A + 16 B) s + 16 A
The Laplace-domain output signal is

 1  n2  Form a set of four simultaneous algebraic equations, by


Y (s) =  2   2 2
.
 s   s + 2 n s +  n  equating coefficients of like powers of s.
570 9  Transfer Functions, Block Diagrams, and the s-Plane

s3 : 0 = B+C  16 
 2 
s2 : 0 = A + 2.4 B + D s + 2.4 s + 16  s=0
s1 : 0 = 2.4 A + 16 B  B  Cs + D 
= A + s2   + s2  2
s0 : 16 = 16 A  s  s + 2.4 s + 16 
0 s=0 0 s=0

Express this system of equations in vector-matrix form


16
=1= A
0  0 1 1 0  A  16
 0   1 2.4 0 1   B 
 =   Find B. The technique used to find A does not work to find
 0   2.4 16 0 0  C  1
     B. Multiplication by s leaves on the left side, and an s in
16  16 0 0 0  D  s
the denominator of the “A” term.
Pre-multiply both sides by the inverse of the coefficient ma-
trix to solve  1    A  B  Cs + D 
s 2   2  = s 2  + s  + s 2
 s   s + 2.4 s + 16  s   s  s + 2.4 s + 16 
−1
 0 1 1 0 0
 1 2.4 0 1  0 This expression cannot be evaluated using s = 0 to find B,
   because two terms blow up to infinity
 2.4 16 0 0 0
   
 16 0 0 0 16 ∞
 1  16 
 0 1 1 0
−1
 0 1 1 0  A     2 
s s + 2.4 s + 16 
 1 2.4 1   1 2.4 1   B 
s=0
0 0
=  ∞
 2.4 16 0 0  2.4 16 0 0  C   A  Cs + D 
=   + B + s 2
    
0  D  s  s + 2.4 s + 16 
 16 0 0 0  16 0 0 s=0 0 s=0

−1
0  A  The resolution to this dilemma is to multiply both sides by s2,
 0 1 1 0  0  1 0 0
 1 2.4 0 1   0  0 and then differentiate with respect to s
   = 1 0 0  B 
 2.4 16 0 0  0  0 0 1 0  C   1 16   A B Cs + D 
s2  2   2 = s2  2 + + 2
      
1  D   s   s + 2.4 s + 16  s 
s s + 2.4 s + 16 
 16 0 0 0 16 0 0 0

 1   A  A 16 Cs 3 + Ds 2
 −0.15     = A + Bs + 2
 = I B = B
2
 s + 2.4 s + 16 s + 2.4 s + 16
 0.15  C  C 
     
 −0.64  D  D d  16  d  Cs 3 + Ds 2 
 2  =  A + Bs + 2
ds s + 2.4 s + 16 ds  s + 2.4 s + 16 
Conventional Algebra
Find A. Multiply both sides by s2, and evaluate for s = 0
du dv
−u v
 1 16  d  u ds ds and keeping in mind that if K is
s2  2   2 using  =
 s   s + 2.4 s + 16  ds  v  v2
dK
 A  B  Cs + D  a real or complex constant, then =0
= s2  2  + s2   + s2  2 ds
s   s  s + 2.4 s + 16 

(
d s 2 + 2.4 s + 16 )
0 − 16
ds =
−16 ( 2 s + 2.4)
= B+
(3Cs 2
)( )
+ 2 Ds s 2 + 2.4 s + 16 − ( 2 s + 2.4) Cs 3 + Ds 2 ( )
( ) (s ) (s )
2 2 2 2 2 2
s + 2.4 s + 16 + 2.4 s + 16 + 2.4 s + 16

−16(2 s + 2.4) s (3Cs + 2 D)( s 2 + 2.4 s + 16) − (2 s + 2.4)(Cs 2 + Ds) 


= B +
( s 2 + 2.4 s + 16) 2 ( s 2 + 2.4 s + 16) 2
Chapter 9 Appendix 571

−16 ( 2 s + 2.4) (
s (3Cs + 2 D ) s 2 + 2.4 s + 16 − ( 2 s + 2.4) Cs 2 + Ds 
= B+   ) ( )
(s ) ( )
2 2
2 2
+ 2.4 s + 16 s + 2.4 s + 16
s=0 0 s= 0

− ( 2.4)(16)
= −0.15 = B
(16)2

Find C and D. Multiply both sides of the expansion by Determine C. It is easier to evaluate the expressions, s12 , s22 ,
s 2 + 2.4 s + 16 , and evaluate twice, using both roots, and ( s1 − s2 ) , first. Then substitute them into the expression
s1 = −1.20 + j 3.82 and s2 = −1.20 − j 3.82. for C
Note: It is always best to wait to substitute values for
complex constants, until the last possible point in a reduc-
s12 = ( −1.20 + j 3.82) = −13.15 − j 9.17
2

tion, to save effort and reduce the risk of algebraic errors and
sign errors. We will use the variable names, s1 and s2, until
we wish to evaluate the result.

 1 16  2 (
A s 2 + 2.4 s + 16 B s 2 + 2.4 s + 16 (Cs + D ) s 2 + 2.4 s + 16 ) ( ) ( )
 2   2
s

s + 2.4 s + 16 
s + 2.4(s + 16 =
s2
+ ) s
+
s 2 + 2.4 s + 16

16
=
(
A s2 + α s + β ) +
(
B s2 + α s + β ) + Cs s = s + D
2 2
s s = s1 s 0
s 0
1

s = s1 s = s1

and

16
=
(
A s 2 + 2.4 s + 16 ) +
(
B s 2 + 2.4 s + 16 ) + Cs s = s + D
s 2 s = s2 s2 0 s 0
2

s = s2 s = s2

leading to and
16 16
= Cs1 + D and = Cs2 + D
s22 = ( −1.20 − j 3.82) = −13.15 + j 9.17
2 2
s1 s22

These two equations with two unknowns yield Notice s12 and s22 are complex conjugates

16 16
− = Cs1 + D − (Cs2 + D ) = C ( s1 − s2 ) s1 − s2 = −1.20 + j 3.82 − ( −1.20 − j 3.82) = j 7.64
s12 s22

 s2 − s2 
 1 1 1 C = 16  2 22 1 
16  2 − 2  =C  s1 s2 ( s1 − s2 ) 
 s1 s2  ( s1 − s2 )

 −13.15 + j 9.17 − ( −13.15 − j 9.17 ) 


 s2 − s2  16 C = 16  
16  2 22 1  = C and − Cs1 = D  ( −13.15 − j 9.17 ) ( −13.15 + j 9.17 ) j 7.64 
 s1 s2 ( s1 − s2 )  s12
C = 0.15
572 9  Transfer Functions, Block Diagrams, and the s-Plane

Determine D  f (t ) = ur (t ) = t for t ≥ 0

16 16 Unit Ramp  1
− Cs1 = − ( 0.15) ( −1.20 + j 3.82) = D  F ( s ) = s 2
s12 −13.15 − j 9.17

In order to add the two terms, multiply the first term by


the ratio of the complex conjugate of the denominator,

Oscillatory  f (t ) =
 1
wn
− ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t )
Unit Impulse 
 −13.15 + j 9.17  w n2
 −13.15 + j 9.17  , to create the ratio of a complex number Response  F ( s ) = 2
 s + 2ζw n s + w n2

over a real number. Then divide real and imaginary compo-
nents of the complex number by the real denominator 

16  −13.15 + j 9.17  −210.4 + j146.7



Oscillatory  f (t ) = −
1
1−ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t − φ )
= Unit Impulse 
−13.15 − j 9.17  −13.15 + j 9.17  257.0  s
Response  F ( s ) = 2
= −0.819 + j 0.571  s + 2ζw n s + w n2
with
  2 
16 16 Phase Shift  where φ = tan −1  w d  = tan −1 1 − ζ
D= − Cs1 = − ( 0.15) ( −1.20 + j 3.82)   σ   
   ζ 
2
s1 −13.15 − j 9.17 
D = ( −0.82 + j 0.57 ) − ( −0.18 + j 57 ) = −0.82 + 0.18
1  0.15 
y (t ) = L −1  2  − L −1  
D = −0.64 s
   s 
 0.15s  −1  0.64 
− L −1  2  −L  2 
We now have the coefficients of the partial fraction expan-  s + 2.4 s + 16   s + 2.4 s + 16 
sion
The inverse Laplace transformation is a linear operator, allow-
 1  16  A B Cs + D ing us to factor in or out coefficients, as needed. We will need
Y (s) =  2   2  = 2+ + 2
 s   s + 2.4 s + 16  s s s + 2.4 s + 16 to factor in and factor out to match the transform table pairs

A B Cs D  0.15  1
Y ( s) = 2
+ + + L −1

 s 
L
 = 0.15
−1
 
s
s s s 2 + 2.4 s + 16 s 2 + 2.4 s + 16

 0.15s   s 
Y ( s) =
1

0.15
+ 2
0.15s
− 2
0.64 L −1
 2
 s + 2 .4 s + 16
 = 0.15

L −1
 2
 s + 2 .4 s + 16


s2 s s + 2.44 s + 16 s + 2.4 s + 16

L {Y (s)} = y(t )
−1
L −1 
 2
0.64   16  −1 
 =  L  2
0.64 

 s + 2 .4 s + 16   16   s + 2 .4 s + 16 
 1 0.15 0.15s 0.64 
y (t ) = L −1  2 − − 2 − 2   0.64   0.64  −1  16 
s s s + 2.4 s + 16 s + 2.4s + 16  L −1
 2 =
  L  2 
 s + 2. 4 s + 16  16  s + 2. 4 s + 16 
In order to perform the inverse transformation, we need to
match the two oscillatory terms exactly to the forms of the We can now use the Laplace transform pairs with  = 0.3
Laplace transform pairs and  n = 4
1 1
y (t ) = L −1  2  − 0.15L −1  
 f (t ) = us (t ) = 1 for t > 0 s
  s

Unit Step  1 s
F (s) = s
 
− 0.15L −1  2 
  s + 2.4 s + 16 
 16 
− 0.04L −1  2 
 s + 2 .4 s + 16 
   w 
y (t ) = ur (t ) − 0.15us (t ) − 0.15  −

wn
1−ζ 2
( )
e −ζw n t sin w n 1 − ζ 2 t − φ  − 0.04 

n

 1−ζ
2
(
e −ζw n t sin w n 1 − ζ 2 t 

)
Chapter 9 Appendix 573

2.0 form of the second-order factor we need for the partial frac-
tion expansion:
1.5
 30   2   2 s 2 + 20 s + 18 
y(t) 1.0 Y (s) =   
 s ( s + 3)   s   3s + 7.5s + 24 
2

0.5
 30   2   0.67 s 2 + 6.67 s + 6 
Y (s) =   
 s ( s + 3)   s   s + 2.5s + 8 
2
0
0 0.5 1.0 1.5 2.0
t, sec
−2.5 ± ( 2.5)2 + ( 4)(8)
s1 , s2 = → s1 , s2 = −1.25 ± j 2.54.
Fig. A9.6   Response function, Eq. 9.46 2

y (t ) = ur (t ) − 0.15us (t )
   
1 −(0.3)( 4)t   4 −(0.3)( 4)t  2 
sin  4 1 − ( 0.3) t − 1.27 − 0.04 sin  4 1 − ( 0.3) t 
 2 
− 0.15 − e e
     
1 − ( 0.3)  1 − ( 0.3)
2 2
  

 (
y (t ) = t − 0.15 + 0.16e −1.2t sin (3.82t − 1.27 ) − 0.04 0.17e −1.2t sin (3.82t ) ) (9.46)

Example Five: Laplace transforms, transfer functions, and Complex conjugate eigenvalues. Therefore, keep the charac-
partial fraction expansion are the tools, which allow us to teristic function as a second-order polynomial
determine the response of systems to arbitrary inputs with-
out resorting to a numerical solver, such as the Runge–Kutta  30   2   0.67 s 2 + 6.67 s + 6 
Y (s) =    
method. Our only restriction is that the inputs we apply to  s ( s + 3)   s   s 2 + 2.5s + 8 
the system’s transfer function have a Laplace transform. (It
is possible to approximate almost any function as a sum of 40 s 2 + 400 s + 360
Laplace transforms, but then we may as well use a numerical Y (s) =
solver.) We will use a first-order stable exponential growth as (
s 2 ( s + 3) s 2 + 2.5s + 8 )
the input to a system. This is a common input, which occurs
when the dynamic response of the power supply prevents Note the s 2 factor in the denominator. Equate the signal to its
us from reasonably approximating the input as a Heaviside partial fraction expansion with unknown numerator constants
step input.
40 s 2 + 400 s + 360
The system shown in the block diagram Fig. A9.7 con- Y (s) =
sists of an integrator, followed by an oscillatory second-or- (
s 2 ( s + 3) s 2 + 2.5s + 8 )
der system. Use partial fraction expansion and the Laplace
Transform tables to determine the time-domain response. A B C Ds + E
Y (s) = + + + 2
First, note the second-order transfer function is not in s 2
s s + 3 s + 2.5s + 8
standard form. Put it into standard form. Then calculate the
eigenvalues of the characteristic function to determine the Multiply both sides by the denominator of the left side. Can-
cel common factors

Fig. A9.7   Response function 30


Eq. 9.45
______
s(s + 3) 2
__ 2s 2 + 20s +18
____________ Y(s)
U(s) s 3s2 + 7.5s +12
574 9  Transfer Functions, Block Diagrams, and the s-Plane

( 40s 2
)
+ 400 s + 360 s 2 ( s + 3) s 2 + 2.5s + 8 ( ) Express in matrix form

s2 ( s + 3) ( s 2
+ 2.5s + 8 )  0   0 1 1 1 0  A 
 0   1 5.5 2.5 3 1   B 
A
= 2
(
s 2 ( s + 3) s 2 + 2.5s + 8 )   
 40  =  5.5 15.5 8 0 3  C 
s
    
B 2  400 15.5 24 0 0 0  D 
+
s
s ( s + 3) s 2 + 2.5s + 8 ( ) 360   24 0 0 0 0  E 
C 2
+
s+3
s ( s + 3) s 2 + 2.5s + 8 ( ) Solve

Ds + E
+ 2 s 2 ( s + 3) s 2 + 2.5s + 8 ( )  0 1 1 1 0  0 
−1
s + 2.5s + 8
 1 5.5 2.5 3 1   0 

40 s 2 + 400 s + 360 = A ( s + 3) s 2 + 2.5s + 8 ( )  5.5 15.5 6 0 3  40 
   
15.5 24 0 0 0  400
+ Bs ( s + 3) s + 2.5s + 8 ( 2
)  24 0 0 0 0 360 
(
+ Cs 2 s 2 + 2.5s + 8 )  0 1 1 1 0  0 1
−1
1 1 0  A 
 1  
5.5 2.5 3 1   1 5.5 2.5 3 1   B 
+ ( Ds + E ) s 2 ( s + 3) 
=  5.5 15.5 8 0 3  5.5 15.5 8 0 3  C 
    
15.5 24 0 0 0 15.5 24 0 0 0  D 
Expand polynomials  24 0 0 0 0  24 0 0 0 0  E 
 s 3 + 2.5s 2 + 8s + 3s 2 + 7.5s + 24 
40 s 2 + 400 s + 360 = A  
 s 3 + 5.5s 2 + 15.5s + 24  0 1 1 1 0
−1
 0   A
 s + 2.5s + 8s + 3s + 7.5s + 24 s 
4 3 2 3 2  1 5.5 2.5 3 1  0  B
+ B     
 s 4 + 5.5s 3 + 15.5s 2 + 24 s   5.5 15.5 8 0 3  40  =  C 
     
(
+ C s 4 + 2.5s 3 + 8s 2 ) 15.5 24 0 0 0  400  D 
 24 0 0 0 0 360   E 
+ D (s 4
+ 3s 3
)
+ E (s 3
+ 3s 2
)  A   15 
 B   6.98 
   
(
40 s 2 + 400 s + 360 = A s 3 + 5.5s 2 + 15.5s + 24 )  C  =  −5.61
   
 D   −1.37 
( 4
+ B s + 5.5s + 15.5s + 24 s 3 2
)  E   −35.3
+ C (s 4
+ 2.5s + 8s 3 2
)
Substitute into the partial fraction expansion
+ D (s 4
+ 3s 3
)
+ E (s 3
+ 3s 2 ) Y (s) =
A B
+ +
C
+ 2
Ds + E
2
s s s + 3 s + 2.5s + 8

Collect like powers of s 15 6.98 5.61 1.37 s + 35.3


Y (s) = + − −
s2 s s + 3 s 2 + 2.5s + 8
s4 : 0 = B+C + D
s3 : 0 = A + 5.5B + 2.5C + 3D + E The inverse Laplace transforms must be an exact match to
s2 : 40 = 5.5 A + 15.5 B + 8C + 3E the transform pairs. The relevant pairs are:
s1 : 400 = 15.5 A + 24 B  f (t ) = us (t ) = 1 for t > 0

s0 : 360 = 24 A. Unit Step  1
F (s) = s

Chapter 9 Appendix 575

40
 f (t ) = ur (t ) = t for t ≥ 0

Unit Ramp  1
 F ( s ) = s 2 30

 f (t ) = e − at
y(t) 20

Unit Exponential Decay  1
F (s) = 10
 s+a
0
0 0.5 1.0 1.5 2.0

Oscillatory  f (t ) =
 1
wn
− ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t ) t, sec
Unit Impulse  Fig. A9.8   Response function Eq. 9.47
w n2
Response  F ( s ) = 2
 s + 2ζw n s + w n2

Unfortunately, there is a real coefficient in the time-domain


Oscillatory  f (t ) = −
1
1−ζ 2
(
e −ζw n t sin w n 1 − ζ 2 t − φ ) functions, which is written in terms of ωn and ζ. We could
solve for them using eigenvalues, but it is easier to calculate
Unit Impulse  them by equating like coefficients of the denominator, and
 s
Response  F ( s ) = 2 the standard form of the Laplace transform table
 s + 2ζw n s + wn
2

with
  2   n = 8 = 2.83
Phase Shift  where φ = tan −1  w d  = tan −1 1 − ζ
  σ    2 2 2
s + 2 n s +  = s + 2.5s + 8 → 2.5
   ζ 
n
 = = 0.44
2 n

The first step is to express the final term as the sum of two
terms, rather than a sum over a common denominator Comparing the two oscillatory terms to their corresponding
Laplace transform table entries, we see that we can match the
15 6.98 5.61 1.37 s + 35.3 first term by factoring out its numerator coefficient
Y (s) = + − −
s2 s s + 3 s 2 + 2.5s + 8
15 6.98 5.61 1.37 s 35.3
Y (s) = 2 + 1.37 s
s s
− − 2 − 2
s + 3 s + 2.5s + 8 s + 2.5s + 8 y4 ( t ) = − L −1  s 2 + 2.5s + 8 
The first three terms will have the proper form, after the con-
s
stants in the numerators are factored out y4 (t ) = −1.37 L −1  s 2 + 2.5s + 8 
y1 (t ) + y2 (t ) + y3 (t )
The second oscillatory term requires  n2 in the numerator to
15  − 1  6.98  − 1  5.61 
= L −1
 2+
s 
L 
 s 
− L  
 s + 3
match the table entry. We provide needed value by multiply-
ing both the numerator and denominator by  n2 . This allows
−1  1  −1  1  −1  1  us to factor out the constant, 35.3, and factor in  n2 with no
= 15 L  2  + 6.98
s 
L   − 5.61
s
L  
 s + 3 net change

y1 (t ) + y2 (t ) + y3 (t ) = 15 ur (t ) + 6.98 us (t ) − 5.61 e −3t 35.3


 
y5 (t ) = − L −1
 2 
The time-domain functions corresponding to the two sec-  s + 2.5s + 8 
ond-order oscillatory factors are expressed in terms of the
 8  35.3 
y5 (t ) = −   L −1
natural frequency, ωn, and damping ratio, ζ. We have already  2 
 8  s + 2.5s + 8 
calculated the real exponential decay factor and the damped
natural frequency, since these correspond to the real and
 35.3   8 
y5 (t ) = −  L −1
imaginary terms of the eigenvalues  2 
 8   s + 2.5s + 8 
s =  + j = −ζ  n ±  n 1 −  2
576 9  Transfer Functions, Block Diagrams, and the s-Plane

We are now ready to evaluate the inverse Laplace transforms Note there is no unit ramp function in Mathcad. The unit
of the oscillatory terms ramp, for time, t > 0, is t. Multiplying t by the unit step
function prevents it from being evaluated for negative time.
 s  Using Mathcad’s notation, Φ( t) for the Heaviside unit step
y4 (t ) + y5 (t ) = −1.37 L −1
 2 
 s + 2 . 5 s + 8  function, the unit ramp function is
 35.3  −1  8 
−
 8   L  2 
 s + 2.5s + 8  ur (t ) = t Φ (t )

y4 (t ) + y5 (t )
2.83
= −1.37 e −1.25t sin ( 2.54t )
1 − ( 0.44)
2 References and Suggested Reading
 35.3  −1 Ogata K (2003) System Dynamics, 4th edn. Prentice-Hall, Englewood
− e −1.25t sin ( 2.54t − φ )
 8  Cliffs
1 − ( 0.44)
2
Ogata K (2009) Modern Control Engineering, 5th edn. Prentice-Hall,
Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
   2.54 
where  = tan −1  d  = tan −1  = 1.11 Prentice-Hall, Upper Saddle River
    1.25 

y4 (t ) + y5 (t ) = −4.32e −1.25t sin ( 2.54t )


+ 4.91e −1.25t sin ( 2.54t − 1.11)

The time-domain response is the sum of the five terms.

y (t ) = y1 (t ) + y2 (t ) + y3 (t ) + y4 (t ) + y5 (t )

 y (t ) = 15ur (t ) + 6.98us (t ) − 5.61e − 4.32e sin ( 2.54t )
−3t −1.25t

+ 4.91e −1.25t sin ( 2.54t − 1.11) (9.47)


Frequency Response
10

Abstract
Sinusoidal excitation of dynamic systems is common. Sinusoidal excitation arises from
rotation of elements within the system or environment, and wave phenomena, internal or
external to the system. It is cyclic or periodic. The sinusoid period and its inverse frequency
affect the system’s dynamic response. The characteristics of sinusoids allow them to be
superposed, to approximate arbitrarily shaped periodic inputs, including square waves or
pulse trains. The steady-state response of a linear system to a sinusoidal input is sinusoidal.
In general, the steady-state output sinusoid’s magnitude or amplitude differs from that of
the input sinusoid. Input and output sinusoids peak at different times. The steady-state re-
sponse of a system to sinusoidal inputs across a range of frequencies is called the system’s
frequency response, which is represented graphically by Bode and Nyquist plots.

10.1 Overview of Sinusoidal Excitation The systems are shown in Figs. 3.13 and 3.31. The system
and Frequency Response equations for the velocity of the mass in those systems are
Eqs. 3.26 and 3.34. All are reproduced below for reference.
In the previous chapters, we solved first- and second-order
linear differential system equations for step inputs. We used  F (t ) M dv1g F (t ) dv1g b
= + v1g → = + v1g (3.26)
those superposition step inputs and step responses to create b b dt M dt M
pulse inputs and their corresponding pulse responses. We ob- 2
served if pulse inputs are reduced in duration, then any arbi-  1 dF (t ) d v1g b dv1g K
= + + v1g (3.34)
trary input function can be approximated, by superimposing M dt dt 2 M dt M
time shifted pulses and the system’s response, and calculated
by superimposing the corresponding pulse responses. Forming the transfer functions, starting with the mass-
Sinusoids are the second fundamental type of input we must damper system:
investigate. Cyclic or periodically varying excitation of physical
systems occurs naturally by wind, waves, and tides. Mechani-  F (t )   dv1g  b 
L  =L   + L  v1g 
cal occurrences can be attributed to rotating and reciprocating  M   dt   M 
machine elements. Electrical occurrences arise from alternating dv
electrical current used to power electrical machines and circuits.
1
L { F (t )} = L  1g  + b L v1g { }
M  dt  M
These inputs are either sinusoidal, or they may be reasonably
modeled as sinusoidal. Further, just as scaled and time shifted 1 b
F ( s ) = sV1g ( s ) + V1g ( s )
steps or pulses are superimposed to create an arbitrary non- M M
periodic input, scaled and phase-shifted sinusoids of different fre- 1  b
F ( s ) =  s +  V1g ( s )
quencies can be superimposed to create an arbitrary period input. M  M
We will use the force source mass-damper system and the 1
force source spring-mass-damper system from Chap. 3 as V1g ( s ) M
= (10.1)
the example systems in our investigation of the response of  F (s) b
first and second-order energetic systems to sinusoidal inputs. s+
M
K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_10, © Springer Science+Business Media New York 2014 577
578 10  Frequency Response

Fig. 3.13  a Translational me-


chanical mass-damper system
a b 1
driven by a force source, x,v
b Linear graph of the mass-
damper system driven by a force
source
Mass F(t) b M
F(t) Lubricating
g M fluid film
1 Damping b
g g

Fig. 3.31  a Force source-spring-


mass-damper system of Fig. 3.23,
a b 1
b Linear graph, Fig. 3.24b
x,v

1 F(t) b K M
F(t) K
M g
g

g g
Lubricating fluid
Damping b

Now, the spring-mass-damper system: The step responses of the initially de-energized mass-
damper and the spring-mass-damper systems to a step input
 d v1g b dv1g K 
2
 1 dF (t )  of F( t) = 10  N us( t) are shown in Fig. 10.1. The response of
L  = L  2 + + v1g  the initially de-energized, first-order, mass-damper system
 M dt   dt M dt M 
to the sinusoidal input, F (t ) = 10 N sin(w t ) for the frequen-
 d v1g  b
2
1  dF (t )   dv1g  K
M
L 
 dt 
 = L  2 + L   + L v1g{ } cies of 2 rad/sec and 10 rad/sec, are shown in Fig. 10.2. Fig-
ure  10.3 shows the response of the initially de-energized,
 dt  M  dt  M
second-order, spring-mass-damper system to the same si-
1 b K
sF ( s ) = s 2V1g ( s ) + sV1g ( s ) + V1g ( s ) nusoidal force inputs.
M M M
Inspection of the responses of the first-order, mass-damp-
1  b K er system and the second-order, spring-mass-damper system
 sF ( s ) =  s 2 + s +  V1g ( s )
M  M M to the same magnitude force as both step input and as a si-
1 nusoidal input, with the latter at the frequencies of ω = 2 and
V1g ( s ) s
M ω = 10 rad/sec, yields the following observations.
=
F (s) b K • The transient period of each system appears to be inde-
s2 + s+ (10.2)
M M pendent of the sinusoidal excitation frequency. It is of the
same duration as the transient response of the systems to
Evaluation of the transfer functions, Eqs. 1.1 and 1.2, for the the step input, Figs. 10.1a and 10.2a and b, and Figs. 10.1b
parameters, M = 2 kg, b = 2 N × sec/m, and K = 200  N/m, and 10.3a and b. The transient period is approximately
yields: 8 to 10 sec.
1 • The peak values of the sinusoidal response of the first-
V1g ( s ) M V1g ( s ) 0.5 order, mass-damper system decreased during the tran-
 = → =
F (s) b F (s) s + 1 (10.3)
sient period, with the mean value of the response equal to
s+ zero in steady-state.
M
• The amplitude of the sinusoidal response of the first-order,
and mass-damper system decreased, when the input force fre-
quency was increased from ω = 2  rad/sec to ω = 10  rad/
1
V1g ( s ) s V1g ( s ) sec, Figs. 10.2a and b.
M 0.5s
= → = 2 • The amplitude of the sinusoidal response of the second-
 F (s) b K F ( s ) s + s + 100 (10.4)
s2 + s+ order, spring-mass-damper system increased, when the
M M
10.1  Overview of Sinusoidal Excitation and Frequency Response 579

Fig. 10.1   Responses of the


initially de-energized a first- a 6
b
0.5
order mass-damper system, and M = 2 b = 2 F0 = 10 M = 2 b = 2 K = 200 F0 = 10
b the second-order spring-mass- 5
0.25
damper system to the step input v1g 4 v1g
of 10 N
m
___ 3 m
___ 0
sec 2 sec
-0.25
1
0 -0.5
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t, sec t, sec

Fig. 10.2   Responses of the initially


de-energized first-order mass-damper a 4
b
rad
___
1.0 rad
system to a sinusoidal force of 10 N M = 2 b = 2 F0= 10 ω = 2 sec M = 2 b = 2 F0= 10 ω = 10 ___
sec
with a the frequency of 2 rad/sec, and 2 0.5
b the frequency of 10 rad/sec v1g v1g
0 0
m
___ m
___
sec sec
-2 -0.5

-4 -1.0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t, sec t, sec

Fig. 10.3   Responses of the ini-


tially de-energized second-order
a b
0.2 8
mass-spring-damper system to a rad
M = 2 b = 2 K = 200 F0= 10 ω = 2 ___ rad
___
sec M = 2 b = 2 K = 200 F0= 10 ω = 10 sec
sinusoidal force of 10 N with a 4
0.1
the frequency of 2 rad/sec, and b v1g v1g
the frequency of 10 rad/sec
m 0
___ m 0
___
sec sec
-0.1 -4

-0.2 -8
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
t, sec t, sec

input force frequency was increased from ω = 2  rad/sec to Fig. 10.4, and of the spring-mass-damper system as Fig. 10.5.
ω = 10  rad/sec, Figs.  10.3a and b. Added to these plots is the sinusoidal input force, scaled on
• An exponential decay envelope bounds the first-order, the right side.
mass-damper system’s responses at both input frequen- Inspection of the steady-state plots of the input force and
cies and the second-order, spring-mass-damper system’s the output velocity yields additional observations.
response to the input force at ω = 2  rad/sec, Figs. 10.2a • The first-order, mass-damper system’s output, the velocity
and b and 10.3a. of the mass, peaks after (later than) the force input at both
• An exponential growth envelope bounds the second- ω = 2 and ω = 10 rad/sec, Figs. 10.4a and b.
order, spring-mass-damper’s response to the input force • The second-order, spring-mass-damper system’s velocity
at ω = 10  rad/sec, Fig.  10.3b. output peaks after (later than) the force input at ω = 2 but
• The transient oscillations of the second-order, spring- simultaneously at ω = 10  rad/sec, Figs.  10.5a and b.
mass-damper system’s step response and sinusoidal All of the observations above can be understood through in-
response at ω = 2 rad/sec appear to be the same frequency, vestigation of the transient portion of the sinusoidal response
Figs. 10.1b and 10.3a. The latter transient oscillations are of the two systems. We shall find that, as with step inputs,
superposed on the larger sinusoid, which remains in the the sinusoid response of first-order systems has less varia-
system’s steady-state. tion than the response of second-order systems. We shall also
The steady-state portions from 10 to 14 second of the si- find that working with sinusoids in the time-domain requires
nusoid responses of the mass-damper system are plotted as a great deal of effort which can be eliminated, when we are
580 10  Frequency Response

Fig. 10.4   Steady-state portion


of the responses of the first-order a rad
___
M = 2 b = 2 F0 = 10 ω = 2 sec
b rad
M = 2 b = 2 F0= 10 ω = 10 ___
sec
mass-damper system to a sinu- 4 12 0.6 12
F(t) F(t) v1g
soidal force of 10 N with a the
frequency of 2 rad/sec, and b the 2 6 0.3 6
frequency of 10 rad/sec v1g
v1g F(t) v1g F(t)
m 0 0 0
___ N ___m 0 N
sec sec
2 -6 -0.3 -6

4 -12 -0.6 -12


10 11 12 13 14 10 11 12 13 14
t, sec t, sec

Fig. 10.5   Steady-state portion of


the responses of the second-order a rad
M = 2 b = 2 K = 200 F0 = 10 ω = 2 ___
sec
b rad
___
M = 2 b = 2 K = 200 F0= 10 ω = 10 sec
mass-spring-damper system to a 0.2 12
12 8 v1g
sinusoidal force of 10 N with a F(t) F(t)
the frequency of 2 rad/sec, and b
0.1 v1g 6 4 6
the frequency of 10 rad/sec
v1g F(t) v1g F(t)
m 0
___ 0
N ___m 0 0
sec N
sec
-0.1 -6 -4 -6

-0.2 -12 -8 -12


10 11 12 13 14 10 11 12 13 14
t, sec t, sec

only interested in the steady-state response, by working in case, we would also use a computer to calculate and plot
Laplace-domain, using what is termed the Frequency Re- the time history. In contrast, if we were concerned that the
sponse Equation, which we will develop in Sect. 10.2. steady-state response of a linear model to a sinusoidal input
“Frequency response” refers to the steady-state response contained the critical condition, we could easily calculate the
of a system to a sinusoidal input. The system reaches steady- magnitude and phase shift manually, obviating any need to
state, when the natural (homogeneous or unforced) portion plot the steady-state response to find the critical value.
of the response has decayed to insignificance after four or From our physical experience, we expect the response of
five time constants, leaving just the forced (particular, excit- a physical system to a sinusoidal input to be periodic, but
ed, or driven) response. Steady-state has the same meaning, not necessarily sinusoidal. Further, we expect the magnitude,
whether the input to the system is a sinusoid or a step. The frequency, and phase shift of the periodic output to depend
two steady-state responses, however, are very different. Since on (1) the amplitude of input, (2) the frequency of the ex-
the forced (particular) response of a linear system must have citation, and, in some way, on (3) the nature of the system.
the same form as the input, when a system subjected to a step When we work with linear models of physical systems, our
input is in steady-state, every power variable in the system expectations can be more definite. The response of a linear
is constant. In contrast, when the same system subjected to model to a sinusoidal input will be sinusoidal. Further, the
a sinusoidal input reaches steady-state, every power variable amplitude of the input sinusoid scales the amplitude of the
in the system is a sinusoid. Likewise, when a step response output variable as a direct proportion. We must determine
reaches steady-state, we see that there is no further change how the frequency of the input affects the amplitude of the
with time. When the same system reaches steady-state in re- output, as well as the “phase” relationship between the input
sponse to a sinusoidal input, we see that each cycle is identi- and output sinusoids.
cal to the cycles which preceded it and which will follow it.
If we were concerned that the initial application of a si-
nusoidal load to a linear model would create a critical condi- 10.1.1 Calculating Magnitude and Phase Angle
tion for machine design, we would need to investigate the from a Response
entire duration of the transient response. We can formulate
the transient response of a system to a sinusoidal input, either Physical experience leads us to expect that the magnitude of
by using transfer functions and Laplace transforms, or a set the output variable of a system driven by a sinusoidal input
of first-order differential equations in state-space. In either depends on three variables: (1) the nature of the physical
10.1  Overview of Sinusoidal Excitation and Frequency Response 581

Input Period T Shift ∆t


max in

A 2A|G(jω)|
2A max out
A|G(jω)|
0 0
Output
min out
Input Output
min in

Fig. 10.6   Oscilloscope trace measurements used to calculate the mag- Fig. 10.7   Oscilloscope trace measurements used to calculate period
nitude ratio. The factor of two cancels in the ratio of output/input T and frequency ω, of the two signals and phase shift φ between them

system, (2) the amplitude of the input sinusoid, F0 in our ex- and rad/sec may be presented as 1/sec. The symbol, f, is used
ample system, and (3) the frequency of the input, ω. Further, for frequency in Hz. The symbol, ω, represents angular fre-
we expect the output to have the same frequency as the input, quency in rad/sec.
when the system is steady-state. We do not expect the two If there is a phase difference between the input and out-
sinusoids to necessarily peak at the same time. One may be put signals, measure the smaller time difference, Δt, between
shifted, relative to the other. We have seen these phenomena peaks of the two signals. It is impossible to determine from
in Figs. 10.2 through 10.5. a single pair of input–output traces, whether the time shift
The “magnitude ratio” is the ratio of the amplitude of the between input and output is the larger or smaller of the two
output sinusoid over the amplitude of the input sinusoid. The possible time shifts. In practice, one determines which is the
factor, which represents the effect of the system on the mag- true time shift, by starting with a low input frequency, and
nitude of the output variable, equals the magnitude of the observing the shift as the frequency is increased. Alternative-
transfer function evaluated as a convention function, where ly, if the transfer function of the system is available, phase
the argument is the input frequency multiplied by the com- angle can be calculated, as we shall see in Sect. 10.3.
plex number, j, |G( jω)|, as we shall see when we develop the
Frequency Response Equation If we had measurements of ∆t
= phase shift as a fraction of a cycle
the input and output sinusoids on a two-channel oscilloscope T
or data acquisition system, we can easily calculate the mag- 2π rad ∆t
= G ( jw ) ≡ φ
nitude ratio and, thus, |G( jω)|. cycle T
The time difference between the peaks of the input and
output variables is normalized by the period of the cycle, T, The ratio, ∆ t /T , is the phase shift as a fraction of a cycle.
to yield the “phase angle,” φ. In other words, the phase angle Multiplying the phase shift as a fraction of a cycle (or circle)
φ is the fraction of a cycle by which the input and output by 2π rad yields the phase shift in radians, Fig. 10.7.
sinusoids are shifted in time, with the fraction of a cycle (or
circle) expressed as an angle. 2π rad ∆ t
= G ( jw ) ≡ φ in radians
The magnitude ratio is calculated from traces of the cycle T
input and output sinusoids, by measuring the peak to
peak amplitudes, rather than the mathematical amplitude, Likewise, to calculate the phase angle in degrees, multiply
A, Input (t ) = A sin(w t ), which is from zero to the peak, the fraction of a cycle by 360°. We will routinely use degrees
Fig. 10.6. The rationale for measuring twice the amplitude is for geometric constructions, since we can easily visualize an
the ease of the measurement. angle in degrees. We will routinely use radians for calcula-
tions, since radians are the natural angular measures.
output (t ) peak to peak 2 A G ( jw )
Magnitude Ratio = = = G ( jw )
input (t ) peak to peak 2A 360o ∆t
= G ( jw ) ≡ φ in degrees
cycle T
The duration of sinusoid cycle is its period, T. The periods 2π rad
of the input and output sinusoid are equal in steady-state. In- T ≡ period =
rad
struments report data and are set, by using frequency in units w
sec
of Hz, or cycles/sec. However, system dynamics calculation
∆t
must be made by using frequency in units of rad/sec. Beware φ= 2π rad ∆T ≡ time shift and φ ≡ phase shift
T
that all angular measures are dimensionless, so that both Hz
582 10  Frequency Response

5τ Positive Impulse
Transient Steady-State
F0

v1g F0 sin(ω0t) 0
0
m
___
sec -F0

-2 π __
__ T 2π
__ 3π
__
0 ω0 = 2 ω0 = T ω0
ω=2 ω=4 ω=8
t, periods
0 5 10 15
t, sec Fig. 10.9   Sinusoidal force, F (t ) = F0 sin(w t ). Cross-hatching indi-
cates the half cycle of positive impulse
Fig. 10.8   Response of the mass-damper system to sinusoidal force
input, F (t ) = 10 N sin(w t ), with frequencies of 2, 4, and 8 rad/sec

Positive Impulse
Unfortunately, the time shift and period data often contain
measurement error, due to the difficulty of identifying iden- F0
tical locations in sinusoidal cycles. If the sinusoids were
clean, then centering them both about zero amplitude would
be feasible. The zero crossing would provide the location in F0 sin(2ω0t) 0
the cycle needed for the measurements. In practice, electron-
ic noise is superposed on the signals, limiting the accuracy
of peak-finding algorithms. The difficulty is typically com- -F0
pounded by a degree of non-linearity, which deforms the out- π
___ π
__ 2π
__ 3π
__
0
2ω0 ω0 ω0 ω0
put sinusoid. The human eye is often a more accurate judge
of the peak value or midpoint of a fuzzy trace than are algo- t, periods
rithms. Filtering data to remove noise must be done with care.
Fig. 10.10   Doubling the frequency of the input force to 2w 0 halves the
As we will see later, a noise filter imparts its own dynamic duration of the positive impulse
characteristics to the filtered signal. The safest noise reduc-
tion method provides a simple running average that elimi-
nates outlying values, which an algorithm may misidentify as the transient period, with the responses centered on zero in
“peaks,” without introducing dynamic characteristics. steady-state.

10.1.2.1  Magnitude as a Function of Frequency


10.1.2 Transient and Steady-State Response The decrease in amplitude of the output with increased fre-
of a First-Order System quency of the input can be understood as an instance of im-
pulse–momentum theorem. Recall impulse is the time inte-
The parameter values of M = 2 and b = 2 were used to create gral of force, Fig. 10.9. Increasing the frequency of the input
the step response plot in 10.1a and the sinusoidal response force decreases the half period of the sinusoid. The positive
plots of Figs. 10.2 and 10.4. The eigenvalue and the time impulse delivered to the system, before the force reverses
constant of the system is one second, per Eq. 1.3. The dura- sign, is reduced as the frequency is increased, Fig. 10.10,
tion of the transient response is defined by the engineer as thereby reducing the maximum positive velocity. The same
a multiple of the time constant. We will use five time con- analysis applies to the negative impulse and the maximum
stants as the end of the transient period, which corresponds negative velocity.
to 99.3 % progression to steady-state. We must also consider the damping in this system. We
Plots of the output variable, the velocity of mass must deduct the portion of the applied force, which is bal-
v1g (t ), for three different input force frequencies, anced by the shear of the fluid film, to yield the force acting
w = 2 rad/sec , 4 rad/sec, and 8 rad/sec, Fig. 10.8, show the to accelerate the mass, FM. The force acting to accelerate the
amplitude of v1g (t )   decreasing with increasing frequency mass is plotted with input force F( t) and the output, the veloc-
of the input force, F0 sin(w t ). This is always the case, if ity of the mass, v1g, in Fig. 10.11. The positive and negative
the system is first order, and the input is not differentiated. impulse of FM is cross-hatched. Also shown is the time shift
The response plots show a “sag” of the peak values during Δt between input force F( t) and output mass velocity, v1g.
10.1  Overview of Sinusoidal Excitation and Frequency Response 583

5τ 4
a Transient
Steady-State
6 10 3
F(t) v1g
3 5 F(t)
x 1g 2
v1g ∆T m
N
0 0
m
___ 1
sec FM
-3 -5 N ω=2
FM 0
ω = 0.5 0 5 10 15 20 25
-6 -10
t, sec
0 5 10 15 20 25
t, sec
Fig. 10.12   Displacement of the mass, x1g, for an input force frequency

b Transient Steady-State of ω = 2 rad/sec, calculated by trapezoidal integration of the velocity, v1g
6 10
F(t) With reference to Figs. 10.11b and c, the positive impulse
v1g
3 5
of FM, that is, the area under the first peak, ends when FM
F(t)
swings negative, while the input force is still positive, be-
v1g ∆T N
0 0 cause the mass has gained sufficient velocity for the damp-
m
___ ing force to exceed the input force at that moment. The fol-
sec FM
FM lowing negative impulse of FM is larger than the initial pos-
-3 -5 N
itive impulse, decelerating the mass. That, in turn, leads to
-6
ω=1 -10 a larger positive impulse of FM, because the velocity swings
0 5 10 15 positive, later in that half cycle. At the end of the transient
t, sec
period, the positive and negative impulses of FM are equal,
5τ with the velocity of the mass centered at zero. Do not ex-
c Transient Steady-State
pect the displacement of the mass to be centered about zero.
6 10
F(t)
The initial asymmetry of the mass velocity leads to a non-
zero mean displacement, as shown in Fig. 10.11c, for an
3 v1g 5 F(t) input frequency of ω = 2  rad/sec.
v1g N
0 0
m
___ ∆T 10.1.2.2  Phase Angle as a Function of Frequency
sec FM
-3 -5
The relationship between input frequency and the phase shift
N
and the input and output variables is more difficult to visu-
FM
ω=2 alize or calculate. We have an intuitive awareness that the
-6 -10
0 2.5 5 7.5 10 magnitude of a vibration or oscillation is dependent on the
t, sec frequency of the input, because we can feel or hear the speed
of an engine, a vehicle, or of rotating machinery with its re-
Fig. 10.11   v1g, F( t), and the force acting to accelerate the mass, FM, sulting vibration. We do not have a similar intuitive sense for
are plotted. The cross-hatch area is the impulse accelerating the mass phase shift. We cannot demarcate the time difference in peak
during the transient period. a Input force frequency, w = 0.5rad/sec.
Time shift, Δt = −0.93 sec. b Input force frequency, w = 1rad/sec. Time amplitudes of the input and output, as a fraction of the period
shift, Δt = −0.78 sec. c Input force frequency, w = 2 rad/sec. Time shift, of the oscillation.
Δt = −0.56 sec Phase shift data reveal a great deal about a dynamic sys-
tem. In practice, however, phase shift data tend to be less
The migration of the mean value of the velocity of mass, accurate than amplitude data. When we view the input and
v1g, from its initial positive mean to a mean of zero, can be output sinusoids on a two-channel oscilloscope, measure-
understood by comparison with the force acting to accelerate ment of the time shift between the two wave forms requires
mass, FM, as the difference between applied sinusoidal force human judgment, which leads to error. Use of automatic
F( t) and the damping force acting to shear the lubricating measurement does not improve the accuracy of phase-shift
fluid film, Fb. data. Peak values identified by an instrument’s algorithm are
often noise. In any event, judicious parsing to remove spuri-
FM = F (t ) − Fb → FM = F (t ) − bv1g ous data is sufficient to yield useful results.
584 10  Frequency Response

The steady-state phase angle, φ, of the mass-damper sys- when we discuss Bode plots in Sect. 10.5. Determining the
tem is calculated below, by using the time shift data reported center of the peak of noisy and misshapen sinusoids can
in Fig. 10.11. The negative sign indicates the velocity of the be more prone to error, than the numerical error found in
mass peaks after, or “lags,” the input force. An output which the evaluation of period and phase shifts from actual data.
peaks before the input sinusoid is described as “leading.” Phase-angle data are very important in the process of deter-
A leading output might seem to violate cause and effect. mining the transfer function of a system from its frequency
Shouldn’t the effect follow the cause? There is no violation response. Phase-angle data are useful, even though their
of causality. The lead of an output does not occur on the first uncertainty is much greater than the corresponding mag-
cycle during the transient period. It develops during the tran- nitude data.
sient period, and is established by steady-state, as will be When input and output signals of a system driven by a si-
illustrated with the spring-mass-damper system. nusoidal input are viewed together on an oscilloscope, there
The period of the sinusoid is calculated as T = 2π /w . is nothing inherent in the image to distinguish a phase lag
rad from a phase lead. A shift in time between peaks of the two
Input Frequency w = 0.5
sec sinusoids is all that is apparent. Is the shift a large lead or a
smaller lag? To identify the phase shift as a lead or a lag, one
∆T − 0.93sec
= = − 0.074 decimal fraction of a cycle must have more information, either the phase angles at other
T 12.6 sec input frequencies or the transient response. When frequency
φ = − 0.074 ·2π rad = − 0.46 rad response data are collected manually, begin with the low-
est input frequency feasible. Increase the input frequency to
φ = − 0.074 · 360o = − 26.5o see the progression of the phase shift. An understanding of
theory is needed to interpret phase-angle data.
Input Frequency ω = 1

∆T − 0.78 sec 10.1.3 Transient and Steady-State Response


= = − 0.124 decimal fraction of a cycle
T 6.28 sec of a Second-Order System
φ = − 0.124 · 2π rad = − 0.78 rad Figures 10.2 and 10.3 show the sinusoidal response of sec-
φ = −0.124 · 360o = − 44.7o ond-order systems to be more varied than that of first-order
systems. The basis of the variation in sinusoidal response is
Input Frequency ω = 2 the internal energy transfers, or power flows, which occur
in a second-order system. The homogeneous, or natural,
∆T − 0.56 sec response of a second-order system, apparent in its step re-
= = − 0.18 decimal fraction of a cycle
T 3.14 sec sponse, is characterized by the system’s eigenvalues. Two
real eigenvalues indicate an overdamped, non-oscillatory
−0.18 ⋅ 2π rad = φ = −1.13 rad homogeneous response. A complex conjugate pair of eigen-
values indicates an underdamped response. The same homo-
−0.18 ⋅ 360o = φ = −64.8o geneous, or natural, response of the second-order system is
superposed on its particular or steady-state response, which
Phase angle φ between the input force and the output veloc- is sinusoidal due to the sinusoidal input.
ity increases with increasing input frequency. We shall see Overdamped second-order systems’ sinusoidal response
that the theoretical limit for a first-order system is 90o of is analyzed as the superposition of the responses of two
phase shift φ. At very low frequency, the input force and the first-order factors. The sinusoidal response of under-
resulting velocity of the mass move together with an imper- damped second-order systems is distinctly different. The
ceptible phase shift. As the input frequency increases, the power flow between the two independent energy storage
momentum and kinetic energy of the mass prevent it from elements, which produces oscillations in a step response,
reversing direction as quickly as the input force, leading to a leads to “resonance” in a sinusoidal response. Resonance
phase shift between the input and output. occurs when the input frequency approaches the natural
There is a slight error in the phase angle, φ at the input frequency of the system, due to enhanced power flow from
frequency, ω = 1 rad/sec, due to limited numerical precision. the sinusoidal source into the oscillating system. The maxi-
The input frequency equal to one over the time constant of mum power flow into the system occurs, when the source
a first-order system, w = 1 / τ rad/sec, is known as the sys- frequency matches the natural oscillation of the system.
tem’s “corner” frequency, and equals − 45°, as we shall see, The source pushes when the mass is naturally moving
10.1  Overview of Sinusoidal Excitation and Frequency Response 585

ζ = 0.7 5τ Steady-State
1.0 a Transient
3 σt
0.8 F0 |G(jw)|e + F0 |G(jw)|
2
ωd 0.6
___ v1g 1
ω n 0.4
m 0
___
0.2 sec
-1

0 -2
σt
0 0.2 0.4 0.6 0.8 1.0 -F0 |G(jw)|e - F0 |G(jw)| ω=8
-3
Damping Ratio ζ 0 2 4 6 8
t, sec
Fig. 10.13   Ratio of the observed, damped frequency, wd, over the ideal,
5τ Steady-
completely undamped, natural frequency wn plotted against damping b Transient
State
6
F0 |G(jw)|(1-e σt )
away, and pulls when the mass is naturally approaching. A 4
person pushing a playground swing is synchronized with
the motion of the swing, pushing when the swing is mov- v1g 2
ing away. Second-order systems with damping ratios ζ less
m 0
___
than 0.7 exhibit a “resonant peak,” a local maximum in the
sec -2
magnitude of their output variable, when the input frequen-
cy is nearly equal to their natural frequency. -4
We will first look at the envelope, which bounds the re- ω = 10
sponse of an underdamped second-order system, during -6
0 1 2 3 4 5 6
the transient period and into the steady-state period. Let’s t, sec
use the parameters values, M = 2 kg, b = 2 N × sec/m, and
K = 200 N/m. The spring-mass-damper’s transfer function is
c 5τ Steady-State
Transient
3 σt
1 F0 |G(jw)|e + F0 |G(jw)|
V1g ( s ) s V1g ( s ) 2
M 0.5s
= → = 2
F (s) b K F ( s ) s + s + 100 1
s2 + s+ v1g
M M
m 0
___
The eigenvalues, s1 , s2 = σ ± j w , are s1 , s2 = − 0.5 ± j 9.99. sec -1
The ideal, undamped, natural frequency is
-2
w n = K / M = 200 / 2 = 10. The damping ratio is σt
-F0 |G(jw)|e - F0 |G(jw)| ω = 12
ζ = 0.05. The resonant frequency of an underdamped system -3
0 2 4 6 8
is the actual, damped, frequency, ωd. The ideal, undamped,
t, sec
natural frequency, ωn, is used for convenience. The differ-
ence between the damped frequency and the undamped natu- Fig. 10.14   Velocity of the mass of the spring-mass-damper system
ral frequency is less than 2% for the damping ratio ζ < 0.2, plotted with the envelopes, which bound the response. The parameter
and is one-tenth of 1% when the damping ratio ζ = 0.05, values used were, b = 2 N · sec/m, M = 2 kg and K = 200  N/m. The
natural frequency, ωn, of the system is 10 rad/sec. The frequency of
Fig. 10.13.
the input force was ω = 8, ω = 10, and ω = 12 in a, b and c, respectively.
Figure 10.14 shows the sinusoidal response of the spring- Note that the envelope in b is a growth to the steady-state value, while
mass-damper system at the input frequencies of ω = 8, the envelopes in a and c are decays to the steady-state values
ω = 10, and ω = 12 rad/sec. The bounding envelope for the
responses to the input frequencies of ω = 8 and ω = 12 rad/sec,
Figs. 10.14a and c, which are below and above the natural fre- where the final value is the magnitude of the steady-state
quency, is a real exponential decay to a non-zero final value: response. The coefficient of the real exponential is the real
component of the eigenvalues, σ.
 envelope(t ) = ± ( F0 G ( jw ) eσ t + F0 G ( jw ) ) (10.5)
586 10  Frequency Response

2 5τ
a Transient Steady-State
0.5
1 -F0 |G(jw)|(1-e σt
)
sin(10t)
+sin(11t)
0 v1g
0
m
___
-1
sec

-2
0 2 4 6 8 -0.5 σt
t, sec -F0 |G(jw)|e - F0 |G(jw)| ω = 20
0 2 4 6 8 10
Fig. 10.15   A beat results from the sum of two sinusoids of close to t, sec
equal frequencies. This beat is the sum of sin(10t) plus sin(11t)

b Transient
Steady-State

At the resonant frequency, Fig. 10.14b, the bounding en- F0|G(jw)|(1-e σt)


velope is a stable exponential growth: 0.2

envelopew n (t ) = ± F0 G ( jw ) (1 − eσ t )
(10.6)
v1g
0
The product in the bounding envelope equations, F0 G ( j w ) , m
___
is the steady-state magnitude of the output variable, the ve- sec
locity of the mass, v1g. The steady-state amplitude of the -0.2
sinusoidal response of the system is due to the amplitude -F0 |G(jw)|(1-e σt ) ω = 30
of the input, F0, and its frequency, ω, and to the nature of
0 1 2 3 4 5 6 7
the system. The transfer function for the output variable,
t, sec
G( s), is a mathematical description of the nature of the sys-
tem. As it happens, the factor that the system contributes 5τ
c Transient
to the magnitude of steady-state sinusoidal response is the
magnitude of its transfer function, evaluated as a conven- -F0 |G(jw)|(1-e σt ) ω = 40
0.2
tional function by substituting for the Laplace variable, s,
the product of imaginary number j and the input frequen-
cy, ω, s = jω. Evaluating G( jω) yields a complex number, v1g
which can be expressed in Cartesian form or in polar form m 0
___
by its magnitude, G ( j w ) , and angle, G ( j w ) . The Fre- sec
quency Response Equation is based on this mathematical
fact. It greatly reduces computational effort, when the only -0.2
σt
-F0 |G(jw)|e - F0 |G(jw)|
aspect of the sinusoidal response of interest is the steady-
state. We will review calculations with complex numbers 0 1 2 3 4 5
in Sect. 10.2, and derive the Frequency Response Equation t, sec
in Sect. 10.3. Fig. 10.16   The transient portion of the spring-mass-damper sys-
The input frequencies, ω = 8 and ω = 12 rad/sec, are close tem’s sinusoidal response at input frequencies, w = 2w n , w = 3w n , and
enough to the system’s natural frequency, ωn = 10  rad/sec, w = 4w n , the first three harmonics
that a “beat” forms between 2 and 4 sec into responses. A
beat results from the sum (superposition) of two sinusoids A second interesting phenomenon is the behavior of the
of nearly equal frequencies. The resulting waveform’s am- system, when the input frequency is an integer multiple of
plitude is bounded by a sinusoid of frequency equal to the its natural frequency. In the terminology of vibrations, the
difference between the two sinusoids. The amplitude is natural frequency is the “fundamental” frequency, and its
equal to the sum of the amplitudes of the two sinusoids, integer multiples are its harmonics. The response of the
Fig. 10.15. We often hear beats, when rotating machinery is spring-mass-damper system to the first three harmonics of
running at approximately, but not precisely, the same speed. its resonant frequency, ω = 2ωn, ω = 3ωn, and ω = 4ωn, are
Air travelers hear them, and may feel them, from the turbo- shown in Fig. 10.16. The transient portion of the response
fan jet engines. at ω = 2ωn = 20 rad/sec is fit on the negative side by one
10.1  Overview of Sinusoidal Excitation and Frequency Response 587

bounding envelope, for both the resonant and non-resonant a 5τ Steady


cases. The positive side is almost flat. The responses at the Transient
-State

second and third harmonics are symmetric about zero. 2


FM v1g
35
∆T
As with the first-order, mass-damper system, the magni- F(t)
tude and phase shift in the steady-state response of the sec- 1 20
F(t)
ond-order spring-mass-damper system to a sinusoidal input v1g N
force can be understood by impulse acting to accelerate the 0 0
mass, the area under the trace of FM. m
___ FM
sec
The period of the sinusoid is calculated as T = 2π /w . -1 N
-20
rad 2π
Input Frequency: w = 8 → T= = 0.785 sec ω=8
sec w -2 -35
The output, v1g, peaks before the input, F( t). The output 0 1 2 3 4 5 6
leads the input. The phase angle is positive. t, sec
b 5τ Steady
∆T + 0.180 sec -State
= = + 0.23 decimal fraction of a cycle Transient
T 0.785 sec 6 v1g ∆T=0 100
FM F(t)
o
φ = + 0.23 · 2π rad = +1.45 rad , φ = + 0.23 ·360 = + 82.8o 4
50
2
F(t)
rad 2π v1g N
Input Frequency: w = w n = 10 → T= = 0.628 sec 0 0
sec w m
___ FM
sec
The input is the natural frequency of the system. The out- -2
-50 N
put, v1g, and the input, F( t) peak simultaneously. There is no -4
phase shift. ω = 10
-6 -100
0 1 2 3 4 5 6
∆T 0 t, sec
= = No Phase Shift → φ = 0
T 0.628 sec c 5τ Steady
-State
rad 2π Transient
Input Frequency w = 12 → T= = 0.524 sec 2
sec w FM v1g F(t) ∆T 40
The output, v1g, peaks after the input, F( t). The output
1 F(t)
lags the input. The phase angle is negative. 20
v1g N
0
∆T − 0.112 sec m 0
___ FM
= = − 0.22 decimal fraction of a cycle sec
T 0.524 sec -20
-1 N
φ = − 0.22 · 2π rad = −1.38 rad , φ = −0.22 ·360 = − 79.2
o o
ω = 12 -40
-2
0 1 2 3 4 5 6
The positive phase shift, or lead, for input frequencies less t, sec
than the natural frequency, w < w n , is due to differentiation
Fig. 10.17   Plots of the input force, F( t), the output velocity, v1g, and
of the input force. Close inspection of Fig. 10.17a reveals the force acting to accelerate the mass, FM, of the spring-mass-damper
that the output velocity of the mass does not lead the input system. a Input force frequency w = 8rad/sec. The time shift is posi-
force on the first half cycle, preserving causality. The lead tive, ∆t = + 0.18sec. The output leads the input. b Input force frequency
of the output has started by the end of the first cycle. Differ- is the natural frequency, w = w n = 10 rad/sec. The time shift Δt = 0  sec.
c Input force frequency w = 12 rad/sec. The time shift is negative,
entiation contributes + π/2 rad or + 90° to the phase angle of ∆t = − 0.52sec. The output lags the input
the output, over the entire frequency range. Without differen-
tiation of the input, the output variable would lag the input,
i.e., have a negative phase shift, for all input frequencies. frequencies. The change in the phase angle is − π/2 rad, or
The limit of the change of a second-order system’s phase − 90°, at the second-order system’s corner frequency, which
angle φ is − π rad, or −180°, as it progresses from low to high equals its natural frequency.
588 10  Frequency Response

A sin(ωt) A|G(jω) |sin(ωt+φ) 10.2.1.1 Example One: First-Order Mass-Damper


G(s) System
where φ = G(jω) The transfer function is evaluated by substituting for the
Laplace variable, s, the product of the imaginary number,
Fig. 10.18   Block diagram of the steady-state frequency response re-
j, and the input, or excitation frequency, ω. Any form of
lationship
the transfer function can be used. The two common forms
of a first-order transfer function are the pole zero and the
time constant. The “pole-zero” form is the factored form.
10.2  Frequency Response Relationship Poles are roots of the denominator, and zeros are roots of
the numerator. Note the constants, p and z, in the factors are
The steady-state response of a linear system to a sinusoidal the opposites of the values of s, which represent the pole
input can be easily calculated from the system’s transfer func- and the zero.
tion, G( s). The relationship between the input and the output 1
sinusoids is expressed as a block diagram in Fig. 10.18. s+z V1g ( s ) M
The effect of the system on the sinusoidal output is calcu- Pole − Zero Form G ( s ) = K p − z → =
s+ p F (s) b
lated, by evaluating the system’s transfer function as a con- s+
M
ventional complex function, substituting for s the product of
the imaginary number, j, and the input frequency, ω, in radi- An alternative form is the time-constant form
ans per second, where G ( jw ) is the magnitude of the trans- 1
fer function, evaluated by substituting jω for the Laplace vari- τ s + 1 V1g ( s ) b
able, s, and G ( jw ) ≡ φ is the phase shift between the input Time Constant Form G ( s ) = Kτ N → =
τDs +1 F (s) M
and output sinusoids. It cannot be overemphasized that the s +1
b
frequency response relationship only holds in steady-state.
If the phase shift, φ, is negative, then the response follows, G( s) is a complex function. The result of evaluating G( s)
or “lags” behind, the input. The system is classified as a “lag” for the purely imaginary argument jw , will be a com-
due to its “phase lag.” If φ is positive, the response precedes, plex number with the magnitude, G ( jw ) , and the phase
or “leads,” the input. The system is classified as a “lead” due angle, G ( jw ).
to its “phase lead.” Phase leads may seem to violate the prin- Substituting in the parameters, M = 2 and b = 2, into the
ciple of cause and effect, since the effect (output) appears to pole-zero form yields:
occur before the cause (input). This is not the case, because 1 1
the frequency response equation only applies in steady-state. V1g ( s ) M 2 0.5
The input has been applied long enough for the system to G (s) = = → G (s) = → G (s) =
F (s) b 2 s +1
“get into a rhythm” (or “the swing of things,” if you prefer). s+ s+
M 2
The output is not being caused by the cycle that immediately
follows it, but by the cumulative effect of all the cycles that Evaluating the mass-damper system’s transfer function,
preceded it, during the transient portion of the response. G( s), for the product of the imaginary number, j, and the
The derivation of the frequency response equation is input force frequency of w = 0.5 yields:
presented in Sect. 10.3. We will begin with an example cal-
culation, by using the frequency response relationship. We 0.5
G ( j 0.5) =
will then apply it to calculate the responses of the first and j 0.5 + 1
second-order mechanical systems plotted above.
Handheld calculators with engineering functions can evalu-
ate this ratio, and return a complex number in Cartesian or
10.2.1 Example Frequency Response polar form, which is the efficient approach. We will express
Calculation the numerator and denominator as complex exponentials, to
illustrate the contribution of the numerator and denominator
The arithmetic of complex numbers was discussed in to the magnitude and phase angle, Fig. 10.19.
Sect. 2.7. It is reviewed briefly below. Calculations for the Recall from Sect. 2.7 the magnitude of a complex number
Frequency Response Equation use both the Cartesian and is calculated by the Pythagorean Theorem. The Pythagorean
complex exponential forms of a complex number. Calcu- theorem uses the magnitudes of the real and imaginary
lation of the angle (argument) and magnitude (modulus) components. Remove the imaginary unit vector, j, from the
of a complex number presented in Cartesian form permits imaginary component before squaring that component, to
the number to be expressed in complex exponential form. avoid error.
10.2  Frequency Response Relationship 589

j Im Im

-44+j12 φD
j10
j0.5 1+j0.5

α j6
φD Re
-50 -40 -30 -20 -10 10
0.5 1 Re -j5

Fig. 10.20   Plot of the numerator and denominator of the ratio


Fig. 10.19   Plot of the numerator and denominator of the ratio j6
0.5 G2 ( j12) =
G ( j 0.5) = − 44 + j12
j 0.5 + 1
0.5e j 0 0.5 ( j 0 − j 0.46)
G ( j 0.5) = = e
An angle calculation is prone to error due to the signs of 1.03e j 0.46
1.03
the components of the complex numbers. The angle is calcu-
lated by using the inverse tangent function. G ( j 0.5) = 0.49e − j 0.46

 opposite  10.2.1.2 Example Two: Second-Order Spring-


φ = tan −1 
 adjacent  Mass-Damper System
Evaluate the transfer function of the velocity of the mass
Determine in which quadrant of the complex plane a complex in the spring–mass–damper system for the product of the
number lies, before calculating its angle. If a complex number imaginary number, j, and the input frequency, ω = 12  rad/
falls in the second or third quadrant, then the number’s angle sec. We will again use the parameter values M = 2 kg,
must be calculated indirectly, as illustrated in Sect. 2.7. Hand- b = 2 N × sec/m, and K = 200  N/m.
held calculators’ inverse tangent function calculates the ratio
1
(opposite / adjacent ), and returns an angle in the first quad- V1g ( s ) s
rant, if the ratio is positive, and in the fourth quadrant, if the G2 ( s ) = = M
F (s) b K
ratio is negative. s2 + s+
Express the numerator and denominator as complex ex- M M
ponentials. The numerator’s value is a positive real number.
Its magnitude is its value. Its angle with the positive real axis V1g ( s ) 0.5s
G2 ( s ) = =
is zero. F (s) s 2 + s + 100

0.5 = 0.5 and 0.5 = 0 → 0.5 = 0.5e j 0 Evaluating for s = j12 :

The denominator’s value is a complex number. 0.5( j12) j6


G2 ( j12) = =
( j12) 2 + ( j12) + 100 −144 + j12 + 100
1 + j 0.5 = 12 + 0.52 = 1.03
and j6
G2 ( j12) =
− 44 + j12
 0.5 
 (1 + j 0.5) = tan −1  = 0.46 rad=26.4o
 1 
Although we could divide the numerator and denominator by
j 0.46 two, we will not.
1 + j 0.5 = 1.03e
Express the numerator and denominator as complex ex-
ponentials, Fig. 10.20. The numerator is an imaginary num-
The complex number, G( j0.5), as a ratio of complex expo- ber. Its magnitude is its value. Its angle with positive real
nentials: axis is π/2 rad, or 90°. Differentiation of the numerator adds
positive π/2 rad, or 90°, to the angle of a transfer function
0.5 0.5e j 0
G ( j 0.5) = = evaluated for any input frequency, since the factor is on the
j 0.5 + 1 1.03e j 0.46 imaginary axis. Its magnitude changes with changing input
ea frequency, but its angle is fixed at 90°.
Using the property of exponentials, = ea −b :
eb π j
π
j 6 = 6 and j 6 = rad = 90o → j 6 = 6e 2
2
590 10  Frequency Response

The denominator is a complex number. Its magnitude calcu- Transfer functions are unusual functions. So far, we
lation is straightforward. have used transfer functions as dynamic operators which
operate on the Laplace transform of the input function
−44 + j12 = 442 + 122 = 42.3 multiplicatively, to yield the Laplace transform of the out-
put function. Transfer functions can also be evaluated for
The complex number is in the second quadrant. A handheld an argument, as are ordinary functions. A transfer function
calculator would return the wrong inverse tangent for φD: is a complex function. Evaluating a transfer function for
a complex argument yields a complex number. In the fre-
 12  quency response equation, the system’s transfer function is
tan −1  ≠ −0.27 rad = −15.5o
 −44  evaluated by substituting for the Laplace variable s a purely
imaginary number, the input frequency multiplied by the
This result is clearly incorrect. The desired angle is calcu- imaginary number j. The result is a complex number which,
lated, by finding the angle, α, subtended by the complex like all complex numbers, can be expressed in terms of its
number, − 44 + j12 and the negative real axis. Subtract that magnitude (or “modulus”) and angle (or “phase angle” or
angle from π, or 180°. simply “phase”).
The derivation of the frequency response equation re-
 12 
α = tan −1  = 0.27rad = 15.5o quires the use of Laplace transforms, partial fraction ex-
 44 
pansion, and a representing complex functions as complex
φ D = π − α = π − 0.27 = 2.87rad = 164.4o exponentials. We will use a first-order system with the trans-
fer function,
 ( − 44 + j12) ≡ φ D = 2.87rad = 164.4o
Output ( s ) 1
G (s) = =
The denominator as a complex exponential is Input ( s ) s+2

− 44 + j12 = 42.3e j 2.87 and excite it with a sinusoid of amplitude A and angular fre-
quency ω, input (t ) = A sin(w t ). The Laplace transform of the
The complex number, G2( j12), as a ratio of complex expo- input is
nentials is Aw
Input ( s ) = L { A sin(w t )} = 2
π
s + w2
j π
6 e 2 j − j 2.87
G2 ( j12) = j 2.87
= 0.14e 2 We multiply the transfer function, G( s), by the Laplace
42.3 e transform of the input, to obtain the Laplace transform of
π
j π the output.
6 e 2 j − j 2.87
G2 ( j12) = = 0.14e 2
42.3 e j 2.87 Aw 1
Input ( s )G ( s ) = = Output ( s )
G2 ( j12) = 0.14e − j1.30 = 0.14e − j 74.5
o
s2 + w 2 s + 2

The phase angle of − 74.5° is the correct phase angle for We are interested in the steady-state response of the system
this ideally linear, second-order system. The phase angle is to the sinusoidal input. We need to perform the inverse La-
− 79.2°, calculated from the traces of the input and output place transform to return to the time-domain. We then take
variables in Fig. 10.17c. Although the magnitude of the out- the limit as t → ∞. The function, Output( s), is not in the
put is within 0.7 % of its steady-state at a value of 5τ, the Laplace transform table, so we will need to perform partial
phase angle is not. fraction expansion. Normally, we would keep the denomina-
tor factor, s 2 + w 2 , since sine and cosine are in Laplace trans-
form tables. However, for this derivation, we will factor the
10.3 Derivation of the Frequency Response denominator into the product of first-order factors
Equation
Aw 1
Output ( s ) =
A fortunate mathematical coincidence allows easy calcula- s2 + w 2 s + 2
tion of the steady-state response of a system to a sinusoidal Aw 1
input. The relationship, known as the frequency response Output ( s ) =
equation, is a function of the frequency and amplitude of
( s + jw )( s − jw ) s + 2
the input sinusoid, and makes use of the system’s transfer B1 B2 C
function, Fig. 10.21. Output ( s ) = + +
s + jw s − jw s + 2
10.3  Derivation of the Frequency Response Equation 591

A sin(ωt) A|G(jω) |sin(ωt+φ) Using the same procedure, B2 and C are found to be
G(s)
where φ = G(jω) Aw 1 Aw
B2 = =
Fig. 10.21   Block diagram of the steady-state frequency response re- ( jw + jw ) jw + 2 2 jw ( jw + 2)
lationship
A
B2 = G ( jw )
Determine the numerator constants by multiplying both sides 2j
by the denominator of the corresponding term and canceling
identical factors. Evaluate by using the pole of that term. To Aw
C=
find the constant, B1, 4 + w2

Aw 1 B1 B C
( s + jw ) = ( s + jw ) + 2 ( s + jw ) + ( s + jw )
( s + jw )( s − jw ) s + 2 s + jw s − jw s+2
Aw 1 B2 C
= B1 + ( s + jw ) + ( s + jw )
( s − jw ) s + 2 s − jw s +2
Aw 1 B2 C
= B1 + ( s + jw ) + ( s + jw )
( s − jw ) s + 2 s = − jw s − jw s = − jw
s+2 s = − jw

Aw 1 B2 C
= B1 + ( − jw + jw ) + ( − jw + jw )
( − jw − jw ) ( − jw + 2) − jw − jw s+2
Aw 1 B2 C
= B1 + ( − jw + jw ) 0 + ( − jw + jw )0
( − jw − jw ) ( − jw + 2) − jw − jw s+2

Hence, the constant, B1, is where C is real. The expanded output function is

Aw 1 Aw B1 B2 C
B1 = = Output ( s ) = + +
( − jw − jw ) ( − jw + 2) −2 jw ( − jw + 2) s + jw s − jw s + 2

−A which becomes more intimidating, when the complex con-


B1 =
2 j ( − jw + 2) stants are substituted in:

The transfer function is A 1


Output ( s ) = − G ( − jw )
2j s + jw
1 A 1 Aw 1
G ( s) = + G ( jw ) +
s+2 2j s − jw 4 + w s + 2
2

Therefore, B1 can be written as


That B1 and B2 are complex constants does not affect the in-
−A 1 −A verse Laplace transformation. Constants, either real or com-
B1 = = G ( − jw )
2 j ( − jw + 2) 2 j plex, can be factored out of the inverse transform, because it
is a linear operator.

 B1 B2 C 
L {Output (s)} = Output (t ) = L
−1 −1
 + + 
 s + jw s − jw s + 2 

 B   B2  −1  C 
Output (t ) = L −1  1  + L −1
  +L  
 s + jw   s − jw   s + 2

 1  −1  1  −1  1 
Output (t ) = B1L −1   + B2L   + CL  
 s + jw   s − jw   s + 2
592 10  Frequency Response

All of the inverse transforms are of the form G ( jw ) is the system’s transfer function, G ( s ), evaluated for
s = jw , where ω is the frequency of the forcing function.
 1  The relationship between G ( jw ) and G ( − jw ) can be seen,
L −1
 =e
− at
when both are expressed as complex exponential:
 s + a 

Therefore, 1 1 e j0
G ( jw ) = → G ( jw ) =
jw + 2   w
j  tan −1   
 2 
Output (t ) = B1e − jw t
+ B2 e j wt
+ Ce −2 t w 2 + 4e 

  w
Before dealing with the constants, B1, B2, and C, let us take 1 − j  tan −1   
 2 
G ( jw ) = e 
the limits of the exponentials as t → ∞. We will start with w2 + 4
the limit of e−2t.

1 1 and
lim e −2t = e −2 ∞ = = =0
t →∞ e2∞ ∞
1 1 e j0
G ( − jw ) = → G ( − jw ) =
The homogeneous response of the system decays to zero in − jw + 2   −w  
j  tan −1 
 2  
steady-state. Regardless of the order of the system, we will w2 + 4 e 

always have the same result. Only the particular, or forced,   −w  


response remains in steady-state. 1 − j  tan −1 
 2  
G ( − jw ) = e 

Now, consider the limit w2 + 4


lim e − j w t = e − j w ∞ 1
  w
j  tan −1   
t →∞  2 
G ( − jw ) = e 

w2 + 4
Does e − j w ∞ decay to zero? No. Recall that e jφ is the complex
exponential unit vector, where the angle of the unit vector is
the term in the exponent, multiplied by j. The complex ex- G ( jw ) and G ( − jw ) have the same magnitude, differing only
ponential unit vector e − jwt has a time-varying angle, θ = w t. in the sign of their angle. Let us generalize our notation by
This unit vector rotates about the origin of its complex plane defining
completing a rotation of 2π rad, as t increases by ∆t = 2π / w .
The negative sign does not indicate decay; rather, it inverts 1
the sign of the angle of the unit vector. Consequently, the G ( jw ) ≡
w2 + 4
limit, lim e − jwt does not yield a final value. In the limit, as
t →∞
time goes to infinity, the unit vector accumulates an infinite and
angle rotating about the origin of a complex plane. However, φ ≡ G ( jw ) → φ = N ( jw ) − D( jw )
we cannot distinguish between ∞ and ∞ + 2π . Therefore,
there is no final value for either e jwt or e − jwt . w w
φ = 0 − tan −1   = − tan −1  
Hence, as t → ∞, the steady-state output is  2  2

Output (t ) steady − state = B1e − jwt + B2 e jwt + Ce −2t 0 Using this notation with complex exponential unit vectors,

Substituting in the complex constants, B1 and B2: G ( jw ) = G ( jw ) e jφ and G ( − jw ) = G ( jw ) e − jφ

−A A We can now rearrange the steady-state output expression


Output (t ) steady − state = G ( − jw )e − jwt + G ( jw )e jwt
2j 2j into a form which can be simplified, by using one of Euler’s
equations
We know that the steady-state response of the system, the
particular solution, will be a sinusoid, since we are forcing −A A
Output (t ) steady − state = G ( − jw )e − jwt + G ( jw )e jwt
the system with a sine. Our objective is to eliminate the com- 2j 2j
plex terms, and yield a real trigonometric function. We will
need to use either Euler’s sine or cosine equation. −A A
Output (t ) steady − state = G ( jw ) e − jφ e − jwt + G ( jw ) e jφ e jwt
2j 2j
e jφ − e − jφ e jφ + e − jφ
sin(φ ) = or cos(φ ) =
2j 2
10.4  Fourier Series Approximation of Periodic Signals 593

A
Output (t ) steady − state =
2j
(
G ( jw ) e jφ e jwt − e − jφ e − jwt ) 1

 e j (w t + φ ) − e − j (w t + φ ) 
Output (t ) steady − state = A G ( jw )   Input(t) 0
 2j

Output (t ) steady − state = A G ( jw ) sin(w t + φ ) (10.7)


-1

0 2 4 6 8 10
Equation 10.7 is the Frequency Response Equation which t, sec
we sought.
Fig. 10.22   Unit amplitude square wave with the period, T = 2 sec. Note
the tapered appearance due to the Heaviside unit step function being
assigned a value during its transition
10.4 Fourier Series Approximation
of Periodic Signals
and
2π 
Although sinusoidal inputs are common and important to 
t + 
 1 w
consider in machine design, the frequency response equation  w
has even more general application. The response of a system
bm =
π ∫
t1
f (t ) sin(mw t )dt (10.11)

to a single sinusoid can be extended, to allow calculation of


the response of the system to any arbitrary periodic input. These integrals are intimidating. Recall how to calculate an-
The Fourier series allows any periodic signal f( t) to be con- gular frequency from the period of a signal.
structed as the weighed sum (or superposition) of sinusoids,
plus a constant. 2π
w=
∞ ∞
T

f (t ) = a 0 + ∑ an cos ( nw t ) + ∑ bm sin ( mw t ) (10.8)
n =1 m =1 This can be rearranged as

where ω is the angular frequency (rad/sec). The cosines w 1 2π


= and =T
and sines in the summation are integral multiples of the fre- 2π T w
quency, ω, of f( t), the periodic signal being replicated. The
frequency, ω, is called the fundamental frequency, even The difference between the limits of integration is the period
though it is not necessarily the natural frequency of the sys- of the fundamental frequency.
tem. The frequencies, which are integral multiples of ω, are
harmonics. If a dynamic system is linear, then the response  2π  2π
of the system to a periodic signal can be calculated, as the  t1 +  − t1 = =T
w w
weighted (or scaled) sum (or superposition) of the response
of the system to the sinusoids of the Fourier series of the The integrals are evaluated over one period T, beginning at
input. Although it takes an infinite number of sinusoids in any point in the cycle.
a Fourier series to “exactly” reproduce an arbitrary periodic In the expression for a0, Eq. 10.9, the integral of f (t )
input, it takes relatively few to create a reasonably accurate with respect to time is multiplied by w / 2π = 1/ T . Hence,
approximation. the constant, a 0 , is the time average of the periodic signal,
The coefficients, a 0 , a n, and bm, are calculated by evalu- f (t ). The constant a 0 is the vertical offset (also called a DC
ating the following integrals offset) of f (t ) from zero.
The integrals for a n and bm , Eqs. 10.10 and 10.11, yield
2π 
 
t + 
 1 
w
coefficients for the cosines and sines, inversely weighted for
w (10.9) the harmonic frequencies, nω and mω, respectively.
a0 =
2π ∫
t1
f (t )dt
Recall from Calculus the integrals:

 1
 2π 
t + 
 1 
w ∫ cos(nwt )dt = nw sin(nwt )
w
an =
π ∫
t1
f (t ) cos(nw t )dt (10.10)
594 10  Frequency Response

0.5
1 1

Three v1g F(t)


Term 0 0 0
Series ___
m
sec N
-1 -1
-0.5
0 2 4 6 8 10 0 2 4 6 8 10
t, sec t, sec
Fig. 10.23   Three term Fourier series approximation of a unit ampli- Fig. 10.25   Response of the first-order mass-damper system with the
tude square wave with a period  T = 2 sec of Fig. 10.22 parameter values, M = 2 and b = 2, to a unit square wave force input with
a period T of 2 sec

1 0.5
1
Eight
Term 0 v1g F(t)
Series 0 0
___
m
-1 sec N
-1
0 2 4 6 8 10
t, sec -0.5
0 2 4 6 8 10
t, sec
Fig. 10.24   Eight term Fourier series approximation of a unit amplitude
square wave with a period  T = 2 sec of Fig. 10.22
Fig. 10.26   Response of the first-order system mass-damper system
with the parameter values, M = 2 and b = 2, to a three-term Fourier se-
ries approximation of a unit square wave force input with a period T
and of 2 sec
1
∫ sin(mwt )dt = − mw cos(mwt ) is made reasonably large. That is not the case. The “ring-
ing” at the corners of the approximate square wave is the
Consequently, the integrals for an and bm yield increasingly “Gibbs” phenomenon, named for the same J. Willard Gibbs
small coefficients, as the integers, n and m, increase. The of Gibbs free energy, who identified the ringing as intrinsic
Fourier series can be truncated, when the magnitude of the to a Fourier series approximation of a square wave using a
coefficient is small enough, that the contribution of its term finite number of terms. Although the ringing at the corners
to the sum is insignificant. In practice, relatively few terms looks bad, in practice, it has little or no effect. The Fourier
are needed in a Fourier series. Although it takes an infi- series approximation of the square wave is accurate enough
nite number of sinusoids in a Fourier series to “exactly” for engineering applications.
reproduce an arbitrary periodic input, it takes relatively Just how much fidelity do we need for a useful approxi-
few to create a reasonably accurate approximation. Con- mation? As it happens, in most applications, not much, be-
sider a square wave with an amplitude of one, and a period cause energy storage in systems is an integral process. This
of T = 2 sec, which corresponds to a circular frequency of is referred to as “integral causality.” Integration smoothens
w = π rad/sec , Fig. 10.22. sharp corners and obliterates fine details. Figure 10.25 is the
Figures  10.23 and 10.24 are Fourier series approxima- response of the first-order mass-damper system with the pa-
tions of this square wave, in which the series was truncated at rameter values, M = 2 and b = 2, to the unit square wave with
three terms and eight terms, respectively. Adding five higher a period of T = 2 sec. The response of the same system to a
frequency terms to the series approximation will increase the three-term Fourier series approximation of the square wave
steepness of the transition, and flatten the waveform’s top is shown in Fig. 10.26.
and bottom. Note the corner of the approximate square wave. There is a slight wobble to the response to the three-term
A reasonable expectation would be for the oscillation at the Fourier series approximation. The transition is less sharp.
corners to vanish, when the number of terms in the series The differences between the response of the system to the
10.5  Bode Plots 595

0.4
A sin(ωt) A|G(jω) |sin(ωt+φ)
Square Wave G(s)
Response where φ = G(jω)
0.2
v1g
Fig. 10.28   Block diagram of the steady-state frequency response re-
___
m lationship
sec
0
Three Term
Series Response log, or in decibels versus the common logarithm of the
-0.2 input frequency ω on the abscissa (horizontal axis) and (2)
0 0.5 1 1.5 2 a plot of the phase angle G ( jw ) on the ordinate versus
t, sec common logarithm of the input frequency on the abscissa.
Note that the plot of the magnitude, G ( jw ) , is equivalent
Fig. 10.27   Comparison of the first cycle of the response of the mass- to the response of the system to a sine with unit amplitude
damper system to the unit square wave input force, dark trace, and the
three-term Fourier series approximation of the square wave force, light
since:
trace
A G ( jw ) 1 G ( jw )
square wave and to the three-term Fourier series approxima- = = G ( jw )
A 1
tion are difficult to see, unless they are superimposed and
enlarged over the first cycle, Fig. 10.27. 10.5.1  Review of the Properties of Logarithms
Although an approximate square wave constructed from
three sinusoids is certainly not an exact replica, the super- Before we work with Bode plots, we will review the prop-
position of the system’s response to the three sinusoids re- erties of logarithms. Common logarithms, base 10, and
semble the response of the system to the square wave, within natural logarithms, base e, Euler’s number have the same
the precision of most engineering modeling. properties. The properties of logarithms derive from the
properties of the products and ratio of exponentials. Cal-
culation of a logarithm and exponentiation are inverse op-
10.5  Bode Plots erations.

Except for resonance, the maximum values of the power 10log10 (a) = a and log10 10a = a( )
variables in an energetic system driven by a sinusoid input
occur during the transient responses of startup and shut- We are most familiar with exponents of base 10. When in
down. However, often our interest is the steady-state re- doubt regarding the property of an exponential or a loga-
sponse of a system to a sinusoidal input, Fig. 10.28. When rithm, check it with an example in base 10.
we are only interested in the steady-state response, the plots The logarithm of a product is the sum of the logarithms.
in the input and output sinusoids above have two shortcom-
ings. First, it is difficult to extract the information we need (
10a ·10b = 10a + b → log 10a · 10b = a + b
(10.12) )
to predict the steady-state response of the system, G ( jw )
and G ( jw ), from these plots. Secondly, the information is Similarly, the logarithm of a ratio is the difference of the
only for one specific input frequency. Bode plots, named logarithms of the numerator and denominator.
for Hendrik Bode (1905–1982), are plots of the magnitude
and phase angle of a system, over a wide range (many or- 10a  10a 
 b
= 10a − b → log  b  = a − b (10.13)
ders of magnitude) of input frequency. They are used to 10  10 
represent experimental data, as well as design closed-loop
control systems and electronic “filters” to attenuate un- The logarithm of a power equals the power times the loga-
wanted sinusoidal signals. rithm.
Bode’s solution to the problem of presenting three di-
mensions or axes of data, G ( jw ) , G ( jw ), and ω, on the
two-dimensional plane of a piece of paper was to create

(10 )
a b
= 10a ·b → log 10a (( ) ) = a · b
b
(10.14)

two two-dimensional plots which had one identical axis,


the frequency of the input sinusoid. The two Bode plots are and
(1) a plot of the magnitude of the transfer function G ( jw )
on the ordinate (vertical axis), as either common (base 10) ( )
log c d = d log ( c )
596 10  Frequency Response

The logarithm of a ratio is a restatement of the product of familiar with decibel as the “unit” of sound intensity. In fact,
logarithms. a decibel is not a unit. Sound is a pressure wave. Sound in-
tensity is reported as a dimensionless ratio of air pressures.
10a  10a 
= 10a · 10 − b → log  b  = a − b
10 b
 10  Pressure of Sound Wave
Sound Intensity =
Ambient Air Pressure
and
Sound intensity is presented in decibels, because the pres-
 ( )
log c · d −1 = log ( c ) − log ( d ) (10.15) sure ratio can vary over a very large range. It could be re-
ported as the logarithm of the pressure ratio. However, to
Neither exponentiation nor the logarithm function is a linear avoid working with decimal fractions, the logarithm of the
operator. Consequently, neither operation distributes onto a pressure ratio, a bel, is multiplied by 10 to yield a decibel
sum.

10a + b ≠ 10a + 10b and log ( c + d ) ≠ log ( c ) + log ( d ) y dB = log10 ( x ) (10.17)

We noted that logarithms and exponentiation are inverse op- Unfortunately, an exception was created for the logical defi-
erations. The antilog of a quantity is the exponentiation of nition of a decibel above. A common use of the unit decibel
that quantity to the base on which the logarithm was calcu- is in electrical power calculations. Electrical power is the
lated. The antilog of a common logarithm is calculated by product the current flowing through an element and the volt-
raising 10 to that power. age across the element. Using v = iR. electrical power can
be expressed as:
antilog ( x ) = 10 x
v2
P = iv → P = i 2 R or P =
An important logarithmic value is zero. The antilog of the R
common logarithm equal to zero is one, 100 = 1. This result
derives from the properties of logarithms. Back in the bad old days of slide rules, any step which re-
duced the effort of calculations, even only slightly, was con-
 10   10  sidered worthwhile. Consequently, since
log   = log (1) and log   = log (10) − log (10) = 0
 10   10 
 v2   v
∴ log (1) = 0
( )
log i 2 R = 2 log (iR ) and log   = 2 log  
 R  R

Logarithms only exist for real numbers greater than zero. it was decided to add a second definition for decibel when
There is no exponent for any base which yields either zero only one of the two power variables is used! When the cal-
or a negative number. Negative logarithms represent positive culation involves only one power variable, rather than the
real numbers between zero and one. They derive from the product of two power variables, a decibel is defined as
logarithm of a ratio, where the numerator equals one:
y dB = 20 log10 ( x )
(10.18)
a 0
10 1 10
b
= 10a − b → b
= b = 100 − b The dual definitions can lead to miscommunication and
10 10 10
substantial error, particularly in this case, since the prefix,
and “deci,” is Latin for 10. It is impossible to identify which of
the two definitions of decibel was used, when one is pre-
  1 sented a value in dB without a context or the history of the
log   = log (1) − log ( c ) = 0 − log ( c ) = − log ( c ) (10.16) calculation. The definition of decibel used in system dynam-
 c
ics, control theory, and in Bode plots is y dB = 20 log10 ( x).
You may be wondering, “Why use decibels at all? Why
10.5.2 Decibels not report the logarithm directly?” That is a good question.
Scaling engineering values by factors of 10 and 100 is done to
A decibel, abbreviated dB, is a scaled logarithm. The bel and eliminate the human errors in remembering and communicat-
the decibel were first used in acoustics. The bel is no longer ing decimal fractions. It is easier to remember a number be-
used, but the decibel, 10 or 20 times a bel, is still widely used tween 1 and 100, than it is to remember a number between 0.1
in acoustics and electrical engineering. You are probably and 0.001. We will use both decibels and straight logarithms.
10.5  Bode Plots 597

1 10 100 1,000 1 10 100 1,000


rad
ω, ___
sec

Fig. 10.29   Logarithmic interpolation value


10 20

Fig. 10.30   Expansion of the interval from 10 to 20. Each interval is


Example Convert −3 dB to a decimal fraction. scaled logarithmically

−3
−3
− 3dB = 20 log10 ( x ) → = log10 ( x ) → 10 20 = 10log10 ( x) LB
20
10 − 0.15
= x = 0.708 LA

Verbally, an attenuation of −3  dB is a reduction of almost


30 % (29.2 %). This conversion example illustrates the rea-
soning behind scaling decimal fractions and the use of deci- 1 10 20 100 1,000
bels. Which is easier to remember, −3 dB or the base 10 log
of −0.15 ? Fig. 10.31   Measurement of position within the decade, 10–100, to
determine the ratio, LA/LB

10.5.3  Interpolating on a Logarithmic Scale have the same relative spacing as the gridlines of a decade,
Fig. 10.30.
Interpolating on a logarithmic scale was second nature to Visualizing a decade grid in the interval from 10 to
engineers 40 years ago, prior to electronic calculators, when 20 rad/sec would yield an estimate, hopefully accurate to
calculations were performed with slide rules, where the another digit, 16, in this case. If there were need for more ac-
primary scales are logarithmic scales. (Multiplication and curacy or confidence, then measuring the lengths LA and LB,
division were performed by aligning the scales so as to ei- Fig. 10.31, using a linear scale, such as an engineer’s scale, if
ther add or subtract logarithms. “Slide-rule accuracy” meant you have one, or a millimeter scale, can be used to compute
three significant figures, where the third digit was estimated, the value as follows:
when one worked at the high end of the scale.) Fortunately, LA

slide rules are obsolete, but the use of logarithmic scales re- Interpolation = ( Lower Decade Limit ) · 10 LB

mains a practical way of presenting data which varies over


several orders of magnitude. Unless the datum one wishes to where the Lower Decade Limit refers to the decade in which
read from a logarithmic plot falls on a grid line, reading data the interpolation is made, as in this example, the decade, 10
from a logarithmic plot requires logarithmic interpolation. to 100. The ratio of the lengths LA over LB is the logarithm of
The essence of interpolating on a logarithmic scale is the the unknown in that decade.
division of a scale into logarithmic subdivisions. All loga- If the lengths are LA = 7.0 and LB = 32.5 on an arbitrary
rithmic scales have the same proportions, regardless of the linear scale, then the interpolation yields the value:
magnitude of the number, which the scale may represent.
The technique is best presented as an example. First, we  7 
Interpolation = 10 10 32.5  = 10 (1.64) = 16.4
need to define a “decade.” A decade on a logarithmic scale is  
bounded by limits which differ by a factor of 10. The main
grid lines of 1 to 10 and 10 to 100 bound decades. So do the
grid lines 2 to 20 and 40 to 400. 10.5.4  Log-Magnitude Bode Plots
Now, the example interpolation. Determine the value of
the frequency, indicated by the dashed line on the logarith- Bode plots are created by evaluating the magnitude and
mic scale in Fig. 10.29. phase angle of a transfer function over a range of frequen-
All intervals between gridlines on a logarithmic scale cies. The range chosen depends on the nature of the system.
are subdivided logarithmically. If gridlines were shown As a rule of thumb, start with a reasonable frequency range,
in the interval from 10 rad/sec to 20 rad/sec, they would or “spectrum,” by setting the lower and upper limits two
598 10  Frequency Response

Fig. 10.32   Log-magnitude Bode 1 20


plot of the velocity of the mass V1g(s)
_____ 0.5
____
of the first-order mass-damper =
system
F(s) s+1
0 0
log(0.5) = -0.30

log|G( jω)| -1 -20 |G( jω)|, dB

-2 -40

-3 -60
0.01 0.1 1 10 100
rad
Input Frequency ω, ___
sec

orders of magnitude below and above either the frequency, factor of 10 on a logarithmic scale is known as a “decade.”
1 Consequently, the slope of the trace in this frequency range is
w ≈ , for a first-order system, or an overdamped second-
τ
order system and w ≈ w d or w n . −2 − ( −1) −1 −1 −20 dB
slope = = ≡ or
 rad  10 decade decade
100
10.5.4.1 First-Order System Log-Magnitude  sec 
Bode Plot  rad 
 10 
We will begin with the log-magnitude Bode plot of the first- sec 
order, mass-damper system with the parameter values, M = 2
and b = 2, Fig. 10.32. A slope of − 1 per decade increase in input frequency or,
equivalently, −20  dB per decade when the ordinate is in
1 decibels, for frequencies greater than the corner frequency,
V1g ( s ) M V1g ( s ) 0.5
= → G1 ( s ) = = is characteristic of the log-magnitude plot of a first-order
F (s) b F (s) s +1 system.
s+
M Figure 10.33 shows straight lines overlaid on the low fre-
quency and high frequency portions of the trace, and extend-
An important reference on the log-magnitude axis is the log- ed until they intersect at the corner frequency, w = 1rad/sec.
magnitude = 0 line. What is the value of any base raised to the The corner frequency of a first-order system is equal to the
zero power, i.e., e0 or 100? One. The value of log G ( jw ) = 0 magnitude of the system’s pole
indicates that the magnitudes of the input and output variable
are equal.  1
M b
G1 ( s ) = → wc = (10.19)
log G ( jw ) = 0 → antilog ( log G ( jw ) ) = 10 log G ( jw )
s+
b M
M
100 = 1 → G ( jw ) = 1
10.5.4.2 Second-Order System Log-Magnitude
Notice that between the lower input frequency limit of Bode Plot
w = 0.01 and w = 0.2,, the log-magnitude plot is horizon- We next examine the log-magnitude plot of the second-order
tal or “flat,” with a value of log G ( jw ) = − 0.30. The trace spring-mass-damper system, Fig. 10.34. We again use the pa-
curves downward over the frequency range of approxi- rameter values, M = 2 kg, b = 2 N · sec/m, and K = 200  N/m.
mately w = 0.4 to w = 2. This is the “knee” of the trace.
We will identify the “corner” frequency in this range. The  1
V1g ( s ) s V1g ( s )
plot appears to be virtually straight over the frequency range M 0.5s
= → G2 ( s ) = = 2
above the knee to the upper frequency limit of the plot. F (s) b K F ( s ) s + s + 100
s2 + s+
Notice that the trace passes through log G ( j10) = −1 and M M
log G ( j100) = −2. Again, the increase of frequency by a (10.20)
10.5  Bode Plots 599

Fig. 10.33   Log-magnitude Bode rad


___
plot of the velocity of the mass of Corner Frequency, ω c = 1 sec
the first-order mass-damper sys-
tem, annotated with straight lines 1 20
fit to the low and high-frequency V1g(s)
_____ 0.5
____
segments of the trace. The slope =
of the high-frequency line is −1
F(s) s+1
0 0
or −20 dB per decade of increase
log(0.5) = -0.30
in the input frequency. The inter-
section of the two straight lines -1= -20 dB
defines the corner frequency, log|G( jω)| -1 -20 |G( jω)|, dB
ωc = 1  rad/sec

-2 -40

-3 -60
0.01 0.1 1 10 100
rad
Input Frequency ω, ___
sec

Fig. 10.34   Log-magnitude Bode 0 0


plot of the velocity of the mass V1g(s) _________
_____ 0.5s
of the second-order spring-mass- = 2
-1
F(s) s + s +100 -20
damper system

-2 -40
log|G( jω)| |G( jω)|, dB
-3 -60

-4 -80

-5 -100
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

The slope of the low frequency line is +1, or +20 dB, per de- The log-magnitude Bode plots for the factors, 0.5 and s,
cade of increase in the input frequency. The slope of the high are shown in Fig. 10.41. The factor, 0.5, is a constant not
frequency line is −1, or −20 dB, per decade of increase in the a function of the input frequency. The factor, s, represents
input frequency. The intersection of the straight lines super- differentiation of the input sine. The factor, log jω , equals
posed on the low and high frequency portions of the trace zero, when ω = 1, because the magnitude of the imaginary
defines the corner frequency, ωc = 10  rad/sec, Fig.  10.35. number, j, is unity. The slope of the trace is +1 or +20 dB. An
To understand the Bode plot of this transfer function, we increase in the input frequency by a factor of 10 is defined as
will take the transfer function apart by using the properties of a decade. Since log(10) = 1, and log jω is plotted against
logarithms, and then plot the individual factors. frequency ω on a log scale, the slope is + 1 per decade.
The constant factor, 0.5, is not a function of the input fre-
 a ·b  1 quency, since it is a constant. Thus, its log-magnitude Bode
log  = log(a ) + log(b) + log  
 c   c plot is a horizontal line with the value, log(0.5) = − 0.30. Al-
though multiplying a transfer function by 0.5 reduces, or at-
 0.5s  tenuates, its magnitude, constant factors are termed “gains”
log(G2 ( s )) = log  2
 s + s + 100  and given the symbol, K. A “fractional gain” is an attenu-
ation, as in this case. The effect of a gain factor is a verti-
 1  cal shift of the log-magnitude Bode plot trace. If K > 1, then
log(G2 ( s )) = log(0.5) + log( s ) + log  2
 s + s + 100 
600 10  Frequency Response

Fig. 10.35   Log-magnitude Bode rad


plot of the velocity of the mass Corner Frequency, ω c = 10 ___
sec
of the second-order spring-mass-
damper system, annotated with 0 0
straight lines fit to the low and V1g(s) _________
_____ 0.5s
high frequency segments of the = 2
-1
F(s) s + s +100 -20
trace

-2 -40
log|G( jω)| +1= +20 dB -1= -20 dB |G( jω)|, dB
-3 -60

-4 -80

-5 -100
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Fig. 10.36   Log-magnitude Bode


rad
plot of the transfer function fac- log|jω|=0 at ω = 1 ___
sec
tors, 0.5 and s. The logarithm of
the magnitude of jω has a slope 4 80
of +1, or +20 dB, per decade

3 60
log|jω| +1= +20 dB
2 40
log|G( jω)| |G( jω)|, dB
1 20
log|0.5|
0 0
log(0.5) = -0.30

-1 -20
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

log( K) is positive, and the vertical shift is upward by the Comparison of the slopes of the log-magnitude Bode plot
value of log( K). Conversely, if K < 1, then log( K) is negative, of the transfer function Eq. 10.20, Fig. 10.35, with the slopes
and the vertical shift is downward by the value of log( K). of the log-magnitude Bode plots of the three factors of that
The straight lines superposed on the low and high transfer function, Figs. 10.36 and 10.37 reveal that the slopes
frequency portions of the log-magnitude Bode plot, of the factors sum to yield the slopes of the complete transfer
1 / ( s 2 + s + 100), intersect at the corner frequency of function. The summation of the slopes of the log-magnitude
w c = 10 rad/sec. The corner frequency of an underdamped plots of individual factors of a transfer function to yield the
second-order factor is its natural frequency. There is a dis- slope of the log-magnitude plot of the entire transfer function
tinct “resonant peak” in this log-magnitude plot, due to is due to the properties of logarithms,
the factors low damping ratio, ζ = 0.05. The low-frequency
value of the log-magnitude curve equals the log(1/100).  a ·b
log  = log ( a ) + log (b ) − log ( c ) − log ( d )
The straight line superposed on the trace for frequencies  c · d 
greater than the corner frequency has the slope of −2 or
−40 dB per decade, twice the slope of the similar line of the The effect of the damping ratio ζ on the resonance of a second-
first-order transfer function, Fig. 10.33. order system is illustrated by the family of log-magnitude
10.5  Bode Plots 601

Fig. 10.37   Log-magnitude Bode rad


plot of the transfer function fac- Corner Frequency, ω c = 10 ___
sec
tor 1/ ( s 2 + s + 100). The corner
frequency is w c = 10 rad/sec. 0 0
The straight line superposed on
the trace for frequencies greater
log | 1
___________
(jω) 2 +jω+100 |
than the corner frequency has the
slope of −2 = −40 dB per decade
( (
1
log ___
100
= -2 -40

log|G( jω)| -4 -80 |G( jω)|, dB


-2= -40 dB

-6 -120

-8 -160
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Fig. 10.38   A family of log-mag- 1.5 30


nitude Bode plots showing the
effect of damping ratio ζ on the ζ = 0.02
resonant peak. The input frequen-
cy is normalized (divided) by the
natural frequency of the system. 1.0 20
ζ = 0.05
Note that the numerator of G(s)
is w n2 , giving the magnitude of
the transfer function the value ζ = 0.1
of one and the log-magnitude of 0.5 10
zero at low frequencies ζ = 0.2
log|G( jω)| ζ = 0.3 dB
ζ = 0.4
0.0 0.0
ζ = 0.5
ζ = 0.6
ζ = 0.7
ζ = 0.8
-0.5 -10
ζ = 0.9
ζ = 1.0

-1.0 -20
ωn2
_____________
G(s) =
s + 2ζωns + ωn2
2

-1.5 -30
0.1 1 10
ω
Normalized Input Frequency, __
ω n

plots of Fig. 10.38. These plots are for systems which have of a system to a sinusoidal input. Lightly damped systems
the same ideal, undamped natural frequency ωn as the spring- with resonant frequency below the frequencies excited by
mass-damper systems above but values of the damping ratio their normal operating speeds can be operated safely, but
ζ which range from ζ = 0.02 to ζ = 1.0. they must be started up deliberately, so as to pass through the
Notice that the amplification of the output for the damp- resonant frequency range quickly, relative to the duration of
ing ratio ζ = 0.05 is log G ( jw d ) = 1.0, or a factor of 10. the system’s transient response.
Remember frequency response is the steady-state response
602 10  Frequency Response

Fig. 10.39   Bode phase-angle rad


___
plot of the mass-damper system Corner Frequency, ω c = 1 sec

0o
V1g(s)
_____ 0.5
____
=
-15 o F(s) s+1
o
-30

G(jω) -45 o

-60 o

-75 o

-90 o
0.01 0.1 1 10 100
rad
Input Frequency ω, ___
sec

10.5.5  Phase-Angle Bode Plots In phase-angle Bode plots, it is common for the input fre-
quency, ω, to be expressed in rad/sec, and the phase angle,
The phase angle of a transfer function can be calculated in φ, to be expressed in degrees. The input frequency must be
two ways. The first is to evaluate the transfer function in its in radians per second, to calculate either the magnitude or
entirety. The second is to evaluate the numerator and denom- the phase angle of the transfer function. The resulting phase
inator individually, then subtract the angle of the denomina- angle is then reported in the units we choose. Degrees have
tor from the phase angle of the numerator, to find the phase two advantages over radians. It is much easier to visualize
angle of the transfer function. an angle in degrees. Visualizing an angle in radians, other
than a common fraction or an integer multiple of π, requires
G jw =   N ( jw )  = N jw − D jw ≡ φ a mental calculation to convert from radians to degrees or
( )   ( ) ( ) (10.21)
 D ( jw )  a fraction of a circle or semicircle. Secondly, frequency re-
sponse data are often used in graphical vector constructions.
Bode phase-angle plots are one of the few instances in en- Protractors are marked in degrees, not radians.
gineering, when dissimilar units for the same quantity are
used on the same plot. The “units” in this case are angular 10.5.5.1 First-Order System Phase-Angle
measures. Quotations are placed around “units,” because all Bode Plot
angular measures are, in fact, dimensionless. A radian can be We begin with phase-angle plot of the first-order mass-
seen as dimensionless from its definition: The angle which damper system, still using the parameter values, M = 2 and
subtends an arc of length, L, on a circle with radius R is b = 2, Fig.  10.39.
We evaluate the frequency response of velocity of the
arc length L mass by making the substitution, s = jw .
θ radians =
radius R
V1g ( s ) 0.5 V1g ( jw ) 0.5
= → =
 arc length L   length  F (s) s +1 F ( jw ) jw + 1
[θ radians] =  radius R  =  length  = dimensionless
   
Recall from Sect. 2.7.1.1, the difference between two vec-
It is not obvious that the angular measure of degrees is also tors, Z1 − Z2, is the vector drawn from the tip of Z2 to the tip
dimensionless, but it is. A degree is defined as the angle, of Z1, as shown in Fig. 2.16, reproduced here for reference.
which subtends an arc length equal to 1/360 of the circum- A complex number in Cartesian form is a sum. If we
ference of a circle. A degree is not an arc length. A degree is express the sum, jω + 1, as a difference by a double-sign
a dimensionless ratio of lengths, an arc length over the cir- inversion:
cumference. Revolutions and cycles are also dimensionless
ratios, again, arc length over circumference. jw + 1 = jw − ( −1)
10.5  Bode Plots 603

Fig. 2.16   Vector shortcut to con- Imaginary


struct the difference, Z3 = Z1 − Z2.
The difference Z3 is drawn from Z 3= Z 1 - Z 2
the tip of Z2 to the tip of Z1 j2
-Z2 = -(3+j)

Z 1= -2+j2
Z 3= Z 1- Z 2 j
Z 2 = 3+j
= (-2+j2)-(3+j)
= -5+j
-6 -5 -4 -3 -2 -1 1 2 3 Real

-Z2 = -(3+j) -j

repeated roots of the characteristic equation. The phase-an-


Im gle plot of a critically damped system is Fig. 10.41.
jω s = jω
10.5.5.2 Underdamped Second-Order System
-(- _

Phase-Angle Plot
jω ___
jω ___
1)

+1
_
_

The phase-angle plot of the second-order spring-mass-


φD φD damper system, with the parameter values, M = 2, b = 2, and
K = 200, is Fig.  10.42. The phase-angle plot is of backward
S-shape with an abrupt transition at the corner frequency,
p = -1 -p = 1 Re
the natural frequency, w c = w n = 10 rad/sec. The phase-an-
gle plot begins at + 90° at the lower input frequency limit.
Fig. 10.40   Construction on a complex plane showing jw − ( −1) and Differentiation of the input, represented by the s in the nu-
jw + 1. The former allows the vector, s = jω, to lie on the imaginary merator, adds + 90° of phase angle, since s = jw is a purely
axis, jw + 1 = jw − ( −1) imaginary number, and the angle of the imaginary axis is
+ 90°, measuring angles conventionally, counter clockwise
then we can draw the complex number as a vector on the from the positive real axis.
s-plane with its tail on -1 and its tip on the imaginary axis, The sharpness of the transition is due to the low damp-
Fig. 10.40. ing ratio of ζ = 0.05. In the case of an almost completely un-
It is clear from Fig. 10.40 that the maximum angle of the damped system, the transition is even sharper with almost
first-order denominator factor,  jw + 1, reaches a maxi- square corners. The response goes from having no phase
mum of π/2 = 90° as the frequency, ω, approaches infinity. shift to being − 180° out of phase with the sinusoidal input,
Because the factor is in the denominator, the angle contrib- when the input frequency crosses the system’s natural fre-
uted to the transfer function is negative. quency. You may have experienced a system like this with
a lightly damped spring-mass-damper system at a child’s
 V1g ( jw )   1   1  birthday party, a wooden paddle with the rubber ball tethered
lim    = lim    = lim    to it by an elastic band. The ball and paddle move toward
w →∞
 F ( jw )  w →∞  jw + 1 w →∞  jw − ( −1) 
each other, before the paddle strikes the ball, then away from
each other after the impact. The opposite directions of the
 V1g ( jw )  π
lim  
w →∞
 F ( jw ) 
 w →∞
( 2
)
= lim 1 − ( jw − ( −1)) = − = −90o input, the velocity of the paddle, and the output, the velocity
of the ball, is 180° of phase shift.
Figure 10.43 shows a family of phase-angle plots for un-
All first-order system phase-angle plots have the same derdamped second-order systems in the form of Eq. 10.22,
shape. First-order systems transition through a change of 90° with damping ratios varying from ζ = 0.02 to ζ = 1.0
of phase angle, as the input frequency increases from ap-
proximately two decades below the corner frequency to two w n2
G (s) =
decades above the corner frequency. They pass through 45° s 2 + 2ζw n s + w n2
of phase-angle change at their corner frequency.
A damping ratio of ζ = 1.0 is the so-called critically w n2
G (s) =
damped condition, with two identical real eigenvalues, the
( s − ζw n + jw n 1 − ζ 2 )( s − ζw n − jw n 1 − ζ 2 )

(10.22)
604 10  Frequency Response

Fig. 10.41   Bode phase-angle rad


___
plot of a critically damped Corner Frequencies, ω c = 1 sec
second-order system with corner
frequencies, w c = 1rad/sec 0
o

0.25
G(s) = __________
(s + 1)(s+1)
o
-45

G(jω) -90 o

o
-135

o
-180
0.01 0.1 1 10 100
rad
Input Frequency ω, ___
sec

Fig. 10.42   Bode phase-angle rad


plot of the spring-mass-damper Corner Frequency, ω c = 10 ___
sec
system
o
90
V1g(s) _________
_____ 0.5s
= 2
F(s) s + s +100
o
45

G(jω) 0 o

o
-45

o
-90
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

The input frequency is normalized by the system’s natural conjugate poles, which are symmetric relative to the real
frequency, which is also the corner frequency of an under- axis. When the input frequency, ω = 0, per Fig.  10.44a,
damped second-order system. The numerator’s angle is the angles made by the complex numbers, ( jw − ( − p1 ))
zero, since it is a positive real number, and angles are mea- and ( jw − ( − p2 )), are equal, opposite, and sum to zero.
sured from the positive real axis. The phase angles begin As the input frequency is increased,  ( jw − ( − p1 ) ) be-
at zero at the lower limit of the normalized frequency axis. comes less negative.  ( jw − ( − p1 ) ) equals zero, when the
All of the phase-angle curves pass through a phase angle input frequency equals the damped, natural frequency of
of − 90o, when the input frequency, ω, equals their natural the system, w = w d . The magnitude, ( jw − ( − p1 )) , is at its
frequency, ωn. minimum at w = w d . If the damping ratio of the system
The phase angle of an underdamped second-order fac- is less than ζ = 0.707, then the product of the denomina-
tor can be understood by the constructions in a complex tor factors’ magnitudes, ( jw − ( − p1 )) ( jw − ( − p2 )) , is
plane, shown in Figs. 10.44 and 10.45 of the denomina- also at its minimum when w = w d , and there is a resonant
tor of Eq. 1.21. Underdamped systems have complex peak in the magnitude plot. Figure 10.46 shows rays from
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 605

Fig. 10.43   A family of Bode 0o


phase-angle plots of under- ζ = 0.02
ζ = 1.0 ζ = 0.05
damped second-order systems ζ = 0.9
with damping ratios varying ζ = 0.1
ζ = 0.8 ζ = 0.2
from ζ = 0.02 to ζ = 1.0. The input ζ = 0.7
frequency, ω, is normalized by -45 o ζ = 0.3
ζ = 0.6 ζ = 0.4
the natural frequency, ωn ζ = 0.5

G(jω) -90 o

-135 o

ωn2
_____________
G(s) =
s 2 + 2ζωs + ω n2
-180 o
0.01 0.1 1 10 100
Normalized Input Frequency, __
ω
ωn

Fig. 10.44   Construc-


tions in a complex plane of a Im
b Im
( s + p1 )( s + p2 ), written with -p1 -p1 s-(-p1)
a double sign inversion as
jωd= jω n 1-ζ 2 s = jωd= jωn 1-ζ2
( s − ( − p1 ))( s − ( − p2 )), and evalu- φ1
ated for s = jw . a s = jw = 0.
s-

The angles are equal but oppo-

2)
(-p

-p
site: −φ1 = φ 2 , b s = jw = jw d .
1
)

s-(
The angle, φ1 = 0 s = jω = 0
σ = -ζω n Re σ = -ζω n Re
)
p
-(-
2

s φ2
φ2
-jωd = -jωn 1-ζ2 -jωd = -jωn 1-ζ2
-p2 -p2

the origin representing loci of equal damping ratio. The accurate, except near the corner frequencies of the first- and
damping ratio, ζ = 0.707, is a ray inclined at 45° from the second-order factors. Sketches of the phase-angle plot will
negative real axis. For a given damped natural frequency, lack accuracy, but will be useful.
the closer the pole is to the imaginary axis, the smaller A transfer function is a complex function. When a trans-
the magnitude, ( jw − ( − p1 )) , and the more quickly fer function is evaluated for s = jw , the result is a complex
 ( jw − ( − p1 ) ) changes, as the input frequency approach- number, or, equivalently, the ratio of two complex numbers,
es and then exceeds the damped natural frequency. Eq. 1.22.

 G ( jw ) =
(
K xN1 + jy N1 )( x N2 ) (
+ jy N2  xNm + jy Nm )
10.6 Asymptotic Approximation (x D1 + jyD1 )( x
D2 + jyD2 ) ( x Dn + jyDn )
of the Log-Magnitude Bode Plot
xN + jy N
An understanding of frequency response is enhanced by the G ( jw ) = = x + jy (10.23)
xD + jyD
ability to sketch the Bode log-magnitude and phase-angle
curves. Sketches of the log-magnitude plot will be reasonably
606 10  Frequency Response

Im Rays of Equal ζ 0.4 0.2 0.05


0.7 0.6 0.5 0.3 0.1 jω
s = jω > jωd

1)
(-p
0.8

s-
Moving a pole
φ2 toward the imaginary
-p1 axis decreases its 0.9
jωd = jωn 1-ζ2 damping ratio ζ
0.95
2)
s-(-p

σ
σ = -ζω n Re s-plane

φ2 Fig. 10.46   Rays from the origin of the s-plane of equal damping ratio, ζ

-jωd = -jωn 1-ζ2


-p2 where φ N and φ D are the angles of the numerator and de-
nominator. We can collect the magnitudes and angles as
Fig. 10.45   Construction in a complex plane of ( s + p1 )( s + p2 ),
written as ( s − ( − p1 ))( s − ( − p2 )) and evaluated for s = jw > jw d . xN + jy N e jφ N x + jy N jφ N − jφN
As the input frequency approaches infinity, the angles of both com- G ( jw ) = jφ D
→ G ( jw ) = N e e
xD + jyD e xD + jyD
plex numbers approach 90o and their sum 180°. The angle contributed
to the transfer function is negative, because these terms are in the
denominator
xN + jy N
G ( jw ) = e ( N D)
j φ −φ
 (10.25)
xD + jyD
In order to sketch Bode log-magnitude and phase-angle
plots, we will express the ratio of two complex numbers as The angles of the numerator and denominator are summed.
the product of the magnitude and complex exponential unit To create a sum of the magnitudes, we take the logarithm of
vectors: the ratio of the magnitudes.

x + jy N e ( N N )
j x + jy
 xN + jy N xN + jy N e jφN   xN + jy N 
G ( jw ) =
xD + jyD
= N
xD + jyD e j( xD + jyD )
=
xD + jyD e jφD
log 
 xD + jyD 
(
 = log xN + jy N − log ( xD + jyD ) ) (10.26)

(10.24)
We have the scheme we need. We will now apply it to a
transfer function with multiple factors in the numerator and
denominator.

G ( jw ) =
(
K xN1 + jy N1 )( x N2 ) (
+ jy N2  xNm + jy Nm )
(x D1 + jyD1 )( x
D2 + jyD2 )( x Dn + jyDn )
jφ N1 jφ N 2 jφ N m
K xN1 + jy N1 e xN2 + jy N2 e  xNm + jy Nm e
G ( jw ) = jφ D1 jφ D2
xD1 + jyD1 e xD2 + jyD2 e  xDn + jyDn e jφDm

Rearranging to separate the magnitudes and the complex ex-


ponential factors:

jφ N1 jφ N 2 jφ N m
K xN1 + jy N1 xN2 + jy N2  xNm + jy Nm e e e
G ( jw ) = jφ D1 jφ D2
xD1 + jyD1 xD2 + jyD2  xDn + jyDn e e e jφDm
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 607

We now use the fact that e a eb = e a + b :

K xN1 + jy N1 xN2 + jy N2  xNm + jy Nm e


(
j φ N1 + φ N 2 ++ φ N m )
G ( jw ) =
xD1 + jyD1 xD2 + jyD2  xDn + jyDn e
(
j φ D1 + φ D2 ++ φ Dm )

overall gain of the transfer, and usually yields tremendous


ec
And the fact, = ec − d , which yields: errors. The asymptotic forms of first and underdamped sec-
ed

K xN1 + jy N1 xN2 + jy N2  xNm + jy Nm (( )(


j φ N1 + φ N 2 ++ φ N m − φ D1 + φ D2 ++ φ Dm ))
G ( jw ) = e
xD1 + jyD1 xD2 + jyD2  xDn + jyDn

The Bode log-magnitude and phase-angle plots separate ond-order factors have a resemblance which makes them
these factors as: easier to remember; factor out a value equal to the constant

 xN + jy N xN + jy N  xN + jy N 
log magnitude = log  K 
1 1 2 2 m m

 xD1 + jyD1 xD2 + jyD2  xDn + jyDn 

( ) ( )
log magnitude = log ( K ) + log xN1 + jy N1 + log xN2 + jy N2 +  + log xNm + jy Nm ( )
( ( ) (
− log xD1 + jyD1 + log xD2 + jyD2 +  log xDn + jyDn ) ( ))
and
term, leaving unity behind.
( ) (
Phase Angle φ = φ N1 + φ N2 +  + φ Nm − φ D1 + φ D2 +  + φ Dm ) Equation 10.27 is a third-order transfer function with a
constant gain K and derivative factor in the numerator, along
with a first-order factor and underdamped second factor in
10.6.1  Asymptotic Approximation Form the denominator

We will draw what are known as “asymptotic” approxima-  Ks


G (s) = (10.27)
tions of the log-magnitude Bode plots. The straight lines, (s + a)(s 2
+ 2ζw n s + w n2 )
which were superposed on the log-magnitude plots of the
first- and second-order systems, Figs. 10.33, 10.35 and The required forms of first- and second-order factors for the
10.37, are the asymptotes we will use to approximate the asymptotic approximation have a common feature, i.e., the
log-magnitude curve. The term, asymptote, is used because constant term in the factor is unity. The asymptotic approxi-
the low-frequency and high-frequency portions of the log- mation form is created by factoring out the value of the con-
magnitude plots are not actually straight lines, although they stant term.
appear to be. Consequently, the approximation only equals First-order factor asymptotic approximation form:
the true value at the limits, in this case, the lower and upper
limits of the input frequency of the plot, where we fix the  s 
ends of the straight lines on the true curve. ( s + a ) = a  + 1 (10.28)
a 
The asymptotic approximation of the log-magnitude
Bode plot has required forms for both first and underdamped Underdamped second-order factor in asymptotic approxima-
second-order factors. First and underdamped second-order tion form:
factors must be put into the asymptotic approximation form
 s 2 2ζw n s 
within the transfer function. Failure to do so will affect the (10.29) (
s 2 + 2ζw n s + w n2 = w n2  2 +
 wn w n2
+ 1

)
608 10  Frequency Response

Expressing the transfer function G(s), Eq. 10.27, in asymp- K K K  K 1


totic form: GFirst ( s ) = → = = 
Order s+ p s+ p  s   p  s 
p +1  p + 1
Ks  p 
G (s) =
(s + a)(s 2
+ 2ζw n s + w n2 )  K 1
log GFirst ( jw ) = log  p  jw
Ks Order +1
G (s) = p
s   s 2 2ζw n s 
a  + 1 w n2  2 + + 1 1
a   wn w n2  K
log GFirst ( jw ) = log + log jw
Order p +1
K p
s
aw n2
G (s) = Consequently, the log-magnitude curve of a first-order trans-
s  s
2
2ζw n s 
 + 1  2 + + 1 fer function has a constant gain factor, log( K /p ), which,
a  wn w n2 
when summed to the first-order factor’s log-magnitude
curve, shifts the curve for the transfer function vertically.
The purpose of the asymptotic forms for first and second- The frequency at which the two asymptotes intersect is
order factors becomes clear, when we consider the log-mag- called the “corner” frequency, even though it is not a right
nitudes of the factors in the limiting cases of low and high angle. The corner frequencies of both first- and underdamped
frequencies. second-order factors have physical significance. The corner
frequency of an underdamped second-order factor is the un-
damped natural frequency, ωn. What does the corner frequency
10.6.2 Asymptotic Approximation of a first-order factor correspond to, when a first-order factor
of First-Order Factors cannot oscillate? The time constant, τ, is the quantity, which
provides the units of time in the denominator of the corner fre-
We will first consider the limiting cases of log-magnitude of quency of a first-order factor. Recall that all angular measures
 1   1  are dimensionless. Although we write the units of angular fre-
first-order factor  s  evaluated as log  j w  start- quency, ω, as radians per second, the dimensions are
 + 1  +1 
a  a 
1 1
ing with the low-frequency limit: [w ] =  sec  =  τ 
c
   
Low-frequency limit: K K  K 1
G First ( s ) = → = 
 s+ p s+ p  p  s 
1 
Order
  1 
 p + 1
lim  log  j w   = log  j 0  = log  1 = 0
w →0   +1   +1   1
  a   a  K  K 1 1
= → = wc = p
s + p  p  (τ s + 1) τ
The low-frequency limit establishes one of the two asymp-
totes. Recall the first-order transfer function for the mass- If the first-order factor is in the numerator of the transfer
damper system has the “flat,” or nearly horizontal low- function:
frequency portion of its log-magnitude curve, which corre- s 
sponds to this asymptote. Recall the log-magnitude = 0 line G First ( s ) = K ( s + z ) → K ( s + z ) = z K  + 1
Order z 
means that at the input frequency, ω, the output variable’s
sinusoid has the same amplitude as the input sinusoid. s 
K ( s + z ) = z K  + 1 → w c = z
z 
Output ( jw )
G ( jw ) = =1 A common error is to confuse the constants, p and z, with
Input ( jw )
their corresponding pole and zero. The values of s which
Output ( jw ) yield zero in the denominator and the numerator are the pole
log G ( jw ) = log = log (1) = 0 and zero, respectively:
Input ( jw )
s+z
G (s) = → s pole = − p and szero = − z
The general case of a first-order transfer function expressed s+ p
in asymptotic approximation form is:
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 609

First-order log-magnitude plots all have the same shape. The We now evaluate the high frequency limit of log-magni-
magnitude of the output is “normalized,” when it is divided tude of the first-order denominator factor:
by the magnitude of the input. This enables us to obtain mag-
 1  1
nitude data for any input magnitude, and present the result,   1
lim log j w  = log j∞ = log
as if the magnitude of the input were unity. w →∞
 + 1  + 1 j ∞
a a
Output ( jw ) Output ( jw )
G ( jw ) = = 1
Input ( jw ) Input ( jw ) log = log (1) − log ( ∞ ) = 0 − ∞ = −∞
j∞
The reverse process, that of scaling the normalized magni-
tude data to the magnitude of the input we expect, is one ap- The high frequency limit of the log-magnitude is negative infin-
plication of the log-magnitude Bode plot. Normalization and ity. We need more information to establish the high-frequency
the reverse process of scaling only work, if the system can be asymptote. We will determine the frequency at which the addi-
reasonably modeled as linear. The further the system is from tive term, one, and the reciprocal factor, 1/a, become negligible,
an ideal linear system, the greater the error. relative to jω. We evaluate the log-magnitude of the factor for
In the bad old days, templates with the knee curve of first- the input frequency equal to corner frequency, w = w c = a : v
order log-magnitude traces were sold at bookstores. Most
1 1
engineers calculated, or remembered, the attenuation of the 1
log j w c = log j a = log
true first-order log-magnitude curve at the corner frequency, +1 +1 j +1
a a
−3 dB, and then blended in a smooth curve meeting the as-
ymptotes, at frequencies of approximately 2.5 times above  1 
and below the corner frequency.
log   = log (1) − log j + 1
 j +1 
The error between the log-magnitude approximation and
the true first-order log-magnitude trace at the corner frequen- log (1) − log ( 1 + 1 ) = 0 − log ( 2 ) = − 0.1505
2 2

cy is the same in all cases, because the curve is normalized.


The approximation has no attenuation at the corner frequen- We now increase the frequency to 10 times the corner fre-
cy, whereas the true value is quency, w = 10w c :
1 1
V1g ( jw c ) 1 1
= = log j w = log j 10w c
F ( jw c ) jw c + 1 j1 + 1 a
+1
a
+1

V1g ( jw c ) 1 1  1 
1 1
F ( jw c )
=
j +1
= = = 0.707 log j 10a
+1
= log   = log (1) − log j10 + 1
12 + 12 2  j10 + 1 
a

log
V1g ( jw c )
= log ( 0.707 ) = − 0.151
log (1) − log ( 10 + 1 ) = 0 − log (10.05) = −1.0022
2 2

F ( jw c )
and now to 100 times the corner frequency, w = 100w c :
In decibels,
1 1  
1
V1g ( jw c ) log j 100w c = log j 100a = log  
20 log = 20 log ( 0.707 ) = 20 ( −0.151) = −3.01dB +1 +1  j100 + 1 
F ( jw c ) a a

The attenuation of a first-order denominator factor at its log (1) − log j100 + 1 = log (1) − log ( 100 + 1 )
2 2

corner frequency, −3 dB, provides the definition of “band-


width.” The bandwidth of a signal is the frequency range, 1
over which the signal has less than −3 dB attenuation, which log j 100w c = 0 − log (100.005) = −2.00
corresponds to an attenuation of 29.3 %. The log-magnitude +1
a
plots of a first-order numerator factors and denominator fac-
tors are mirror images. Consequently, the error between the The resulting log-magnitudes of the first-order factor,
asymptotic approximation, and the true log-magnitude curve evaluated at the frequencies of w = w c , w = 10w c , and
at the corner frequency is the inverse of the denominator fac- w = 100w c , are presented in Table 10.1.
tor. The true value of the magnitude at the corner frequency We see that the slope of the log-magnitude curve is
of a numerator factor is an amplification of +3 dB. negative one per decade, between one and two decades in
610 10  Frequency Response

Table 10.1   Change in the log-magnitude of a first-order transfer func- as that used for the first-order factor. We will evaluate the
tion in the first two decades above the corner frequency
low-frequency limit of the asymptotic approximation form,
Input frequency Log(|G( jω)|) Change of a decade
increase in frequency
determine the corner frequency, and then find the slope of the
ωc −0.1505
high-frequency asymptote. Unfortunately, because the shape
of an underdamped second-order log-magnitude curve is a
10ωc −1.0022 −0.8517
function of the damping ratio, the asymptotic approximation
100ωc −2.0000 −0.9978 is significantly less accurate than for a first-order factor. In
the bad old days, one either used a template or attempted to
frequency above the corner frequency. We can now draw the reproduce the resonant peak from a family of curves, similar
“asymptotic” approximation to the log-magnitude curve of a to Fig. 10.38. In the present day, the ability to sketch the
first-order factor once we know the corner frequency. log-magnitude curve is for understanding and design, not to
If the first-order factor is in the denominator, the high-fre- create accurate plots, which we now make using software.
quency asymptote slopes down at negative one per decade of The underdamped second-order asymptotic approxi-
frequency from the corner frequency, Fig. 10.47. If the first- mation form for a factor in the denominator of a transfer
order factor is in the numerator, the high-frequency asymp- function is
tote slopes up at positive one per decade from the corner fre-
quency. If the log-magnitude plot is scaled in decibels, then w n2 1
G (s) = = 2
the slopes are −20 dB per decade and +20 dB per decade. s + 2ζw n s + w n
2 2
s 2ζw n s
+ +1
Figure 10.48 shows the error between the asymptotic ap- w n2 w n2
proximation and the true log-magnitude curve for a first-order
denominator factor. A positive error means that the approxi- Note the resemblance to the asymptotic approximation
mation is greater than the true magnitude. Note that the error form of a first-order factor. Both asymptotic approximation
is symmetric about the corner frequency, ωc. Also note that the forms have unity as the zero-order (constant) term of the
error at the corner frequency equals the attenuation that de- polynomial.
fines the limit of a band width, log G ( jw c ) = −0.15 = −3 dB, We calculate the low frequency limit of the magnitude:
because the asymptotic approximation equals one, and its
logarithm equals zero at the corner frequency. 1 1
lim
w →0
( jw ) 2

+
2ζw n jw
+1
= ( j 0) 2

+
2ζw n j 0
+1
=1
w 2
w 2
w 2
w 2
10.6.3 Asymptotic Approximation of n n n 0 n 0

Underdamped Second-Order Factors


The low frequency limit is the same as for a first-order
The development of the asymptotic approximation for an factor. The low-frequency asymptote is the log-magnitude
underdamped second-order factor follows the same logic equals zero line.

Fig. 10.47   Asymptotic approxi- _


1
mation of the log-magnitude of Corner Frequency ω c = τ
first-order factors
3 60

2 40

log| τjω+1 | 1 +1 20
1 decade
0 0 dB
| 1
log _____
τjω+1 | -1
1 decade
-1 -20

-2 -40

-3 -60
0.1 1 10 100 1,000
ω
Normalized Input Frequency, __
ωc
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 611

0.20 4 1

0.15 3
log ( jw ) 2

+
( 2)(0.5) w n jw + 1
w 2
n w n2
Error Error
0.10 2
log|G(jω)| dB 1
0.05 1 = log j w
2 2
n ( 2)(0.5) w n j wn
+ +1
w 2
n w 2
n
0 0
0.01 0.1 1 10 100
ω 1 1
Normalized Input Frequency, __
ω log = log
c j 2 + ( 2)( 0.5) j + 1 −1 + j + 1

Fig. 10.48   Error between asymptotic approximation and true first- 1


order denominator factor log-magnitude curve. Positive error means log = log (1) − log (1) = 0 − 0 = 0
that the approximation was greater than the curve. The error curve is j
inverted for a first-order numerator factor
Success. Repeat the calculation for 10ωn:
1
We next consider the corner frequency for the asymptotic
approximation. Inspection of Fig. 10.38, reveals that a line log ( jw ) 2

+
2ζw n jw
+1
representing the natural frequency of the system, ω/ωn = 1, w 2
n w n2
passes through the resonant peak of the curves for damping
1
ratio, ζ < 0.1. We established by means of the s-plane vector
constructions, Figs. 10.44, 10.45, and 10.46, that the reso- = log ( j10w n ) 2

+
2ζ10w n jw n
+1
nant peak is a function of the actual, observed, damped fre- w n2 w n2
quency, ωd. We expect the asymptotic approximation to be in
1
error at the resonant peak of the system, in any case. Conse-
log j 2 100w n2 2ζ10w n jw n
quently, we will use the ideal, undamped, natural frequency, + +1
w n2 w n2
ωn, as the corner frequency. It is a convenient value, since it
is the divisor used to create the asymptotic approximation 1
form of underdamped second order factors.
= log j 100 w 2ζ10 w n j w n
2 2
n
We now determine the slope of the high-frequency as- + +1
ymptote as we did for the first factor, by evaluating the log- w 2
n w n2
magnitude at the corner frequencies, ωn, 10ωn and 100ωn.
1
Start with ωn: = log (1) − log  −992 + ( 20ζ ) 
2
log
−100 + j 20ζ + 1  

1 1
and for 100ωn:
log ( jw ) 2

+
2ζw n jw
+1
= log ( jw )n
2

+
2ζw n jw n
+1
w n2 w n2 w 2
n w n2 1
log ( jw ) 2

+
2ζw n jw
+1
1
1 w 2
n w n2
log j w 2ζ w n j w n
2 2
n = log 2
+ + 1 j + j 2ζ + 1 1
2
wn 2
wn
= log ( j100w ) n
2

+
2ζ100w n jw n
+1
w 2
w n2
1 1 n
log = log 1
−1 + j 2ζ + 1 j 2ζ
= log j 10000 w 2ζ100 w n j w n
2 2
n
1 + +1
= log (1) − log ( 2ζ ) = 0 − log ( 2ζ ) = − log ( 2ζ ) w w n2
2
log n
j 2ζ
 1 
= log 
We can check our calculation by referring to Fig. 10.38 for  −10000 + j 200ζ + 1 
the value of the damping ratio, ζ, whose curve passes through = log (1) − log − 9,999 + j 200ζ + 1
the log-magnitude equal zero line at the input frequency, ωn.
= 0 − log  − 9,9992 + ( 200ζ ) 
2
It is the curve with the damping ratio, ζ = 0.5.
 
612 10  Frequency Response

Table 10.2   Change in the log-magnitude of an underdamped sec- ω = 10ωn:


ond-order transfer function in the first two decades above its corner
frequency 1
( )
Input frequency Log(|G( jω)|) Change of a decade increase in
( jw ) = log (1) − log
2
frequency log 2ζw n jw −992 + 202
+ +1
ωn −0.3010 w 2
w 2
n n
10ωn −2.0043 −1.7033
= 0 − 2.0043 = −2.0043
100ωn −4.0000 −1.9957
ω = 100ωn:
We must decide how to evaluate the terms with the damping
ratio, ζ. One way to proceed is to evaluate the expressions 1
for ζ = 1, since that curve is a limiting condition farthest from log ( jw ) 2

+
2ζw n jw
+1
the high-frequency asymptote. Alternatively, we could use w n2 w n2
ζ = 0.5, which lies closest to the asymptote. Lastly, we could
decide not to bother evaluating the terms with the damping = log (1) − log  −99992 + ( 200) 
2

 
ratio, ζ, because the negative real number under the radical
is larger in magnitude, and when squared, will dominate. We = 0 − 4.0000 = −4.0000
will evaluate the log-magnitude expressions, using ζ = 1.
ω = ωn: Summarizing the results in Table 10.2, we see that for ζ = 1,
the slope of the curve equals negative two per decade by
within a decade above the corner frequency. The equivalent
1 slope is −40 dB per decade in decibels.
( jw ) = 0 − log ( 2) = −0.3010
2
log 2ζw n jw The asymptotic approximation of the log-magnitude curve
+ +1
w 2
n w 2
n
of an underdamped second-order denominator factor is plot-
ted on top of the true family of curves, Fig. 10.49. Note that
the best agreement between the asymptotic approximation
and the actual log-magnitude curves is for ζ = 0.6. The

Fig. 10.49   Asymptotic ap- 1.5 30


proximation of an underdamped
second-order denominator factor, ζ = 0.02
superposed on the curves of the
true log-magnitude for damping
ratios from ζ = 1.0 to ζ = 0.02 1.0 20
ζ = 0.05

ζ = 0.1
0.5 10
Asymptotic ζ = 0.2
Approximation
log|G( jω)| ζ = 0.3 dB
ζ = 0.4
0.0 0
ζ = 0.5
ζ = 0.6
ζ = 0.7
ζ = 0.8
-0.5 -10
ζ = 0.9
ζ = 1.0

-1.0 -20
1
_____________
G(s) = 2
__
s 2ζs
+ ___ + 1
ωn2 ωn
-1.5 -30
0.1 1 10
ω
Normalized Input Frequency, __
ω n
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 613

Fig. 10.50   Asymptotic ap- Corner Frequency ωc = ω n


proximation of an underdamped
second-order denominator factor.
4 80
The corner frequency is the un-
damped, natural frequency. The
slope after the corner frequency 3 60

| |
2
is −2, or −40 dB, for a denomina- ( jω) ____
2ζ jω
tor factor and + 2, or + 40 dB, for log ___ + ω +1 2 40
a numerator factor ω 2n n +2
1 20
1 decade

| |
0 0 dB
1
____________ 1 decade
2 -1 -20
log (___
jω) ____
2ζ jω
+ ω +1 -2
ωn 2 n
-2 -40

-3 -60

-4 -80
0.01 0.1 1 10 100
ω
Normalized Input Frequency, __
ωc

asymptotic approximations for underdamped second-order if the input frequency is increased by a factor of 10, then
factors in both the numerator and the denominator are shown the log-magnitude of the differentiator increases by posi-
as Fig. 10.50. tive one, Fig. 10.51.

log j10w = log (10w ) = log (10) + log (w ) = 1 + log (w )


10.6.4  Integrator and Differentiator Factors
Hence, the slope of a differentiator factor is positive one, or
The Laplace variable, s, as a factor in the numerator, repre- +20 dB, per decade of frequency.
senting differentiation, Integrators have the inverse relationship, since they are
factors of the denominator.
Ks
G (s) =
(
( s + a ) s 2 + 2ζw n s + w n2 ) log
1
= log (1) − log (w ) = 0 − log (w ) = − log (w )
jw
and in the denominator, representing integration (a so-called
free integrator), Increasing the input frequency by a factor of 10,
K
G (s) = 1
(
s ( s + a ) s 2 + 2ζw n s + w n2 ) log
j10w
= log (1) − log (10w ) = 0 − log (10w )

contributes to the log-magnitude Bode plot across the fre- − log (10w ) = − (log (10) + log (w )) = −1 − log (w )
quency spectrum. Neither has a corner frequency.

10.6.4.1 Log-Magnitude of Integrator decreases the log-magnitude of the integrator factor by nega-


and Differentiator Factors tive one. The slope of the integrator’s contribution to the
A differentiator contributes the log of the input frequency to log-magnitude plot is negative one, or −20 dB, per decade
the log-magnitude Bode plot increase in frequency.
Locating the straight log-magnitude line of a differentia-
log jw = log (w ) tor or integrator is easiest at either the lower limit of the input
frequency, or at the input frequency, w = 1. Note that the log-
Using the fact magnitude line of both the integrator and the differentiator
log ( a · b ) = log ( a ) + log (b ) .
614 10  Frequency Response

Fig. 10.51   Log-magnitude The Integrator and Differentiator cross the


rad
plots of an integrator and a log| jω| = 0 line at the input frequency ω = 1 ___
differentiator sec
3 60

2 40

1 20
log| jω |
0 0 dB

| |
log ─
1
jω -1 -20

-2 -40

-3 -60
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

1
jω Recall the property of exponentials, = e− a
s = jω ea
s-plane
π  1 1 1 − jπ
s= jω = __ = = e 2 (10.30)
2 s jw w

σ Factors in the denominator contribute negative angles to the


1
__ 1
__ 3π
__ transfer function’s phase angle. The angle of an integrator, a
s
=

=
2
1
__ __ π
1 - __
= =
s jω 2 multiplicative s in the denominator, is −90° = −π/2.
1 __ 1 Although the magnitude of a differentiator and an integrator
__ =
s jω is a function of the input frequency, their angles are not. The
angle contributed by a differentiator or an integrator is a con-
Fig. 10.52   s-plane plots of a differentiator, s, and an integrator, 1/s stant ± 90o = ± π /2 across the frequency range or spectrum.

cross the log-magnitude = 0 line at the input frequency, ω = 1,


since: 10.6.5  Gain Factors

log j = log (1) = 0 The term, “gain,” means “multiplicative constant,” which is
and not necessarily positive or greater than one. The term, gain,
and the symbol, K, are used for many different constants. A
1 1
log = log = log (1) − log (1) = 0 − 0 = 0 transfer function’s gain is changed, when the transfer function
j j is expressed in asymptotic approximation form, as illustrated:

10.6.4.2 Phase Angle of Integrator Ks


G (s) =
and Differentiator Factors ( s + a ) ( s 2 + 2ζw n s + w n2 )
The phase angle of a differentiator is easy to understand
from its plot on the s-plane, Fig. 10.52. A purely imaginary Ks
G (s) =
number, s = jw , lies on the imaginary axis at an angle of s   s 2 2ζw n s 
+ 90o = + π /2 to the positive real axis. a  + 1 w n2  2 + + 1
a   wn w n2 
The phase angle of an integrator is understood, once it is
expressed as a complex exponential:  K  s
G (s) =  2 
1 1 1 1 1 aw
 n s  s 2
2ζw n s 
= → = =  + 1  2 + + 1
s jw s jw ejw j
π
a  wn w n2 
we 2
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 615

Fig. 10.53   Sketch One. Indi- Corner Frequency ωc = 8


vidual factors’ log-magnitude
plots 1 20
Gain = log(4)=0.6
0.6 12 dB

0 0
___
1
1s+1
__
8
1 decade
log|G(jω)| -1 -20 dB
Integrator -1
-1
1 decade
-2 -40

-3 -60
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

K 6 Calculate the log-magnitude range, by finding the differ-


Here, the gain is the term, . ence between the sums of the log-magnitude of the factors at
aw n2
the lower and upper frequency limits.
The log-magnitude of a gain is independent of frequency, 7 Sketch Two. Starting from the lower frequency limit, sum
since a gain is a constant. Consequently, the log-magnitude the log-magnitudes, working from left to right.
contribution of a gain constant is a horizontal line, because
the factor contributes equally across the frequency spectrum. 10.6.6.1 Example One: Sketch the Log-Magnitude
The effect of a gain on the resultant log-magnitude plot of Bode Plot for
the entire transfer function is a vertical shift equal to the log
of the gain. 32
G1 ( s ) =
s ( s + 8)
10.6.6 Sketching Asymptotic Approximation
Log-Magnitude Bode Plots 1 The transfer function is already in factored form.
2 Express the factored transfer function in asymptotic ap-
Two sketches are need. One is for the log-magnitude plots proximation form.
of the individual factors. The second is for the sum of indi-
32 32 4
vidual factors of the first plot. G1 ( s ) = = → G1 ( s ) =
The process for sketching a log-magnitude Bode plot is s ( s + 8) s   s 
s 8  + 1 s  + 1
as follows: 8  8 
1 Factor the numerator and denominator of the transfer func-
tion into: (1) gain K, (2) a differentiator or an integrator, 3 The corner frequency of the first-order denominator factor
(3) first-order factors, and (4) underdamped second-order is 8 rad/sec.
factors. 4 Estimate the input frequency range. The corner frequency
2 Express the first and second-order factors of the transfer falls in the decade, 1–10 rad/sec. Adding one decade below
function in the asymptotic approximation form. and two decades above yields an input frequency range of
3 Identify the corner frequencies of the first and under- 0.1 to 1,000 rad/sec.
damped second-order factors. 5 Sketch One, Fig. 10.53. Sketch the individual factors’ con-
4 Estimate the frequency range by adding two decades tributions to the log-magnitude plot.
above the highest, and one decade below the lowest corner Calculate the log of the gain factor, Log(4) = 0.60.
frequency. 6 Calculate the range in log-magnitude of the sum of the
5 Sketch One. Sketch the individual factors’ contributions. component factors, by summing the log-magnitude at the
lower and upper frequency limits.
616 10  Frequency Response

Fig. 10.54   Asymptotic ap- Corner Frequency ωc = 8


proximation of the log-magni-
tude of the transfer function, 2.0 40
G ( s ) = 32/s ( s + 8) 1.6 32 dB
1 decade

-1
0 0

log|G(jω)| -2.0 1 decade -40 dB


-2

-4.0 -80
4
G1(s) = _________
s
s ___
8
+1( ) -120
-6.0
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

Log Magnitude at the Lower Frequency Limit: 10.6.6.2 Example Two: Sketch the Log-Magnitude
1.0 + 0.6 + 0 = 1.6 Bode Plot for
Log Magnitude at the Upper Frequency Limit: 32 ( s + 2)
G2 ( s ) =
0.6 − 2.1 − 3.0 = −4.5
( 2
s s + 30 s + 200 )
7 Sketch Two, Fig. 10.54. Sum the factors. Start at the lower
frequency limit. Work to the right, in the direction of in- 1 Although this transfer function appears to be factored,
creasing frequency, to create the composite plot, which is the check the eigenvalues of the second-order factor. If they are
asymptotic approximation of the log-magnitude plot of the purely real, then the factor is overdamped, and must be ex-
transfer function. The composite plot will consist of straight- pressed as the product of the two first-order factors. Failure
line segments. Sum the slopes of the lines of all factors, at to check the eigenvalues of second-order factors is the most
the lower frequency limit. The sum of the slopes is the slope common error sketching log-magnitude Bode plots.
of the composite log-magnitude line, at the lower frequency
limit. The composite plot can be created by summing values − 30 ± 302 − ( 4)( 200)
s 2 + 30 s + 200 = 0 → s1 , s2 =
of the log-magnitudes of the individual factors at selected 2
frequency values. However, it is easier to sum slopes and s1 = −10 and s2 = −20
extend the log-magnitude line at that slope, until there is a
change in slope. The slope of the composite line from the The second-order factor is overdamped. It must be factored.
lower frequency limit does not change until the first corner
32 ( s + 2) 32 ( s + 2)
frequency. Therefore, extend the composite log-magnitude G2 ( s ) = =
line at the slope of − 1/decade, from the lower frequency ( 2
s s + 30 s + 200 ) s ( s + 10)( s + 20)
limit to the corner of the first-order factor, w c = 8. The slope
changes at the corner frequency from negative one per de- 2 Express the factored transfer function in asymptotic ap-
cade to negative two per decade. There are no other corner proximation form.
frequencies. Extend the composite log-magnitude line at the
s 
slope of − 2/decade from the corner frequency of the first-
32 ( s + 2)
+ 1 (32)( 2) 
order factor, w c = 8, to the upper frequency limit to com- 2 
G2 ( s ) = =
plete the plot. s ( s + 10)( s + 20)  s   s 
s (10)  + 1 ( 20)  + 1
If we compare the asymptotic approximation and the  10   20 
true log-magnitude curve, Fig. 10.55, we find the only s 
error is the frequency range of the first-order factor’s cor- 0.32  + 1
2 
ner frequency. G2 ( s ) =
 s  s 
s  + 1  + 1
 10   20 
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 617

Fig. 10.55   True log-magnitude Corner Frequency ωc = 8


curve ( black trace) and the
asymptotic approximation of 2.0 40
the log-magnitude ( lighter 1.6 32 dB
Asymptotic
trace) of the transfer function, Approximation
G1 ( s ) = 32/s ( s + 8)
0 0

log|G(jω)| -2.0 True Log -40 dB


Magnitude Curve

-4.0 -80
4
G1(s) = _________
(
s
s ___
8
+1 )
-6.0 -120
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

Fig. 10.56   Asymptotic Corner Frequencies ωc = 10


approximations of the log-
ωc = 2 ωc = 20
magnitudes of the factors
of the transfer function,
G ( s ) = 32( s + 2)/( s ( s + 10)( s + 20)) 2.0 40

__
1s + 1 +1
2
Integrator
1.0 1 decade 20
1 decade

-1
___
1
1 s+1
___
log|G( jω)| 0 10 0 dB
log(0.32) = -0.5 1 decade
-10
Gain = 0.32
-1
-1.0 -1 -20
___
1
1 decade 1 s+1
___
20

-2.0 -40
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

3 Identify the corner frequencies. The three first-order factors Calculate the log of the gain factor. Log(0.32) = − 0.495 = − 0.5.
have corner frequencies of w c1 = 2, w c 2 = 10, and w c3 = 20. 6 Estimate the log-magnitude range of the summation. The
4 Estimate the frequency range. The lowest corner frequency log-magnitude value at the lower frequency limit can be
is in the decade, w = 1 to w = 10 . The highest corner fre- summed directly. The upper limit summation requires ex-
quency is in the decade, w = 10 to w = 100. The estimated trapolating the lines of the individual factors, to estimate
frequency range for the Bode log-magnitude plot is from their values for the input frequency, w = 10, 000 rad/sec. The
w = 0.1 to w = 10, 000. extrapolation is based on the slope of each line.
5 Sketch One, Fig. 10.56. Sketch the individual factors’ con- Log Magnitude at the Lower Frequency Limit:
tributions to the log-magnitude plot. 1.0 + 0 − 0.5 = 0.5
618 10  Frequency Response

Fig. 10.57   Asymptotic Corner Frequencies ωc = 10


approximation of the log- ωc = 20
magnitude Bode plot ωc = 2
of the transfer function,
G2 ( s ) = 32( s + 2)/( s ( s + 10)( s + 20)) 1.0 20
1 decade
0 -1 0
1 decade
-1.0 -1 -20

-2.0 -40

log|G( jω)| -3.0 -60 dB


1 decade
-4.0 -2 -80

-5.0 -100

-6.0 G2(s) =
0.32 ( )
s
___
2
___________
+1
-120

-7.0
s
s ___
10
+1( )( )
20
s
___
+1
-140
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Fig. 10.58   True log-magnitude Corner Frequencies ωc = 10


curve ( black trace) and the ωc = 20
asymptotic approximation of
ωc = 2
the log-magnitude ( lighter
trace) of the transfer function, 1.0 20
G ( s ) = 32( s + 2)/( s ( s + 10)( s + 20)) Asymptotic
0 0
Approximation
-1.0 -20

-2.0 -40

log|G( jω)| -3.0 True Log -60 dB


Magnitude Curve
-4.0 -80

-5.0 -100

-6.0 G2(s) =
0.32 ( )
s
___
2
___________
+1
-120

-7.0
s
s ___
10 ( )( )
+1
20
s
___
+1
-140
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Log Magnitude at the Upper Frequency Limit:


in Example One. The asymptotic approximation for a first-
3.7 − 0.5 −2.7 −3 −4 ≈ –6.5 order factor has minimum error, if there are no other corner
7 Sketch Two, Fig. 10.57: Sum the factors to create the com- frequencies in the frequency band of the knee region of the
posite log-magnitude line for the transfer function. Work true curve. In this example, the two first-order denominator
from low frequency to high frequency. Sum the slope at each factors’ knee regions overlap with each other and with knee
inflection point. of the first-order numerator factor. This leads to an accumu-
Comparison of the asymptotic approximation and the true lation of error between the asymptotic approximation and the
log-magnitude curve, Fig. 10.58, reveals greater error than true curve.
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 619

Fig. 10.59   Asymptotic approxi- Corner Frequencies


mations of the log-magnitudes of ωc = 2 ωc = 20
the factors of the transfer function,
G3 ( s ) = 132s /(( s + 2)( s 2 + 30 s + 400)) 1.0 20
Differentiator +1
1 decade 1
_________
0 s2 _____
_____ 30s 0
+ +1
log|G( jω)| 400 400 dB
Gain = 0.165 1 decade
log(0.165) = -0.78 -15.6
-1.0 -20
-2
1
____
s
__ -1
+1
2 1 decade
-2.0 -40

-3.0 -60
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

10.6.6.3 Example 3: Sketch the Log-Magnitude 3 Identify the corner frequencies. The first-order factor has
Bode Plot for the corner frequency, w c1 = 2. The corner frequency of the
second-order factor is the ideal, undamped, natural frequen-
132 s cy of the factor, w n = w c 2 = 400 = 20.
G3 ( s ) =
( s + 2) ( s 2
+ 30 s + 400 ) 4 Estimate the frequency range. The lowest corner frequency
is in the decade, w = 1 to w = 10. The highest corner fre-
1 Although this transfer function appears to be factored, quency is in the decade, w = 10 to w = 100. The estimated
check the eigenvalues of the second-order factor. frequency range for the Bode log-magnitude plot is from
w = 0.1 to w = 10, 000.
−30 ± 302 − ( 4)( 400) 5 Sketch One, Fig. 10.59. Sketch the individual factors’ con-
s 2 + 30 s + 400 = 0 → s1 , s2 =
2 tributions to the log-magnitude plot.
s1 , s2 = −15.0 ± j13.2 Calculate the log of the gain factor. Log(0.165) = − 0.783.
6 Estimate the log-magnitude range of the summation, ex-
The second-order factor has complex conjugate eigenvalues, trapolating the lines to the upper frequency limit as needed.
and is, therefore, underdamped. Keep the factor as a second- Log Magnitude at the Lower Frequency Limit:
order polynomial. 0 − 0.8 − 1.0 = −1.8
2 Express the transfer function in asymptotic approximation
Log Magnitude at the Upper Frequency Limit:
form.
4 − 0.8 − 3.7 − 5.4 ≈ −5.9
132 s
G3 ( s ) = The lower frequency sum estimate is too low an upper
( s + 2) ( s 2 + 30s + 400) limit for the log-magnitude axis. The log-magnitude line will
rise with a slope of positive one, due to the differentiator
132 s from the lower frequency limit of w = 0.1 to the corner fre-
G3 ( s ) =
s   s2 30  quency of the first-order factor denominator factor at w = 2.
2  + 1 400  + s + 1
2   400 400  The upper limit must be at least log-magnitude equal zero.
There are eight horizontal divisions, so we will use log-mag-
0.165s nitude equal zero as the upper limit.
G3 ( s ) =
s  s
2
30  7 Sketch Two, Fig. 10.60 Sum the factors to create the com-
 + 1  + s + 1 posite log-magnitude line for the transfer function. Work
2  400 400 
from low frequency to high frequency. Sum the slopes at
each inflection point.
620 10  Frequency Response

Fig.  10.60   Asymptotic ap- Corner Frequencies


proximation of the log-magni- ωc = 2 ωc = 20
tudes of the transfer function,
G3 ( s ) = 132s /(( s + 2)( s 2 + 30 s + 400)) 1 20

0 0
1 decade
-1 +1 -20

-2 1 decade -40
-2
log|G( jω)| -3 -60 dB
-4 -80

-5 -100
132s
___________
G3 (s) =
( )( )
-6 s2 30s -120
s + 1 _____
__ + _____ + 1
2 400 400
-7 -140
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Fig. 10.61   True log-magnitude Corner Frequencies


curve ( black trace) and the ωc = 2 ωc = 20
asymptotic approximation of
the log-magnitude ( lighter 20
1
trace) of the transfer function,
Asymptotic
G3 ( s ) = 132s / (( s + 2)( s 2 + 30s + 400)), Approximation
0 0
damping ratio ζ = 0.75
-1 -20

-2 True Log Magnitude -40


Curve ζ = 0.75
log|G( jω)| -3 -60 dB
-4 -80

-5 -100
132s
G3 (s) = ___________
-6

-7
( )(s2
s + 1 _____
__
2
30s
+ _____ + 1
400 400 ) -120

-140
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

When comparing the asymptotic approximation with the true 10.6.7 Bode Phase-Angle Plots of Transfer
log-magnitude curve, we find surprisingly good agreement, Functions
per Fig. 10.61. The surprising aspect is that the effect of
the damping ratio, ζ, is not included in the asymptotic ap- We will now create phase-angle plots for transfer func-
proximation. Since the system’s damping ratio of ζ = 0.75 is tions which have a combination of factors. The properties
greater than ζ = 0.7, we may conclude the system log-mag- of exponentials apply to complex exponentials, allowing the
nitude plot has no resonant peak. If we reduce the damping phase-angle plots of each factor to be calculated separately.
to ζ = 0.075, then we find substantial error in the frequency Then the phase angles of the factors are summed, to yield the
range of the resonant peak, Fig. 10.62. phase-angle plot for the transfer function.
10.6  Asymptotic Approximation of the Log-Magnitude Bode Plot 621

Fig. 10.62   True log-magnitude Corner Frequencies


curve (black trace) and the ωc = 2 ωc = 20
asymptotic approximation of
the log-magnitude (lighter 1 20
trace) of the transfer function, True Log
G4 ( s ) = 132 s /(( s + 2)( s 2 + 3s + 400)), Magnitude Curve 0
0
damping ratio ζ = 0.075 ζ = 0.075
-1 -20
Asymptotic
-2 Approximation -40

log|G( jω)| -3 -60 dB


-4 -80

-5 -100
132s
G4 (s) = ___________
( )( )
-6 3s
s 2 _____ -120
s + 1 _____
__ + +1
2 400 400
-7 -140
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

Fig. 10.63   Phase-angle Bode Corner Frequency ωc = 8


plot of the factors of and the
complete transfer function, 45 o 0.785
G1 ( s ) = 32 / ( s ( s + 8))
o
0 0

-45
o
( 1
____
s+8 ) -0.785
( )
1
__
s
G( jω) -90
o
-1.57 Rad
G1(s)
-135o -2.36

-180
o
-3.14
32
________
G1(s) =
s(s + 8)
-225
o -3.93
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

10.6.7.1  Example One: Phase-Angle Bode Plot 10.6.7.2  Example Two: Phase-Angle Bode Plot
It is not necessary to express a transfer function in the asymptot-
ic approximation form, in order to create its phase-angle Bode 32( s + 2)
Phase-angle Bode plot of G2 ( s ) =
plot, Fig. 10.63. It is important to factor the transfer function. s ( s 2 + 30 s + 200)
Each factor contributes to the angle of the transfer function. Check the eigenvalues of the second-order factor. If the
32 32 4 eigenvalues are real, express the second-order polynomial as
G1 ( s ) = → G1 ( s ) = = the product of two first-order factors, Fig. 10.64:
s ( s + 8) s  s 
8s  + 1 s  + 1
8  8  s 2 + 30 s + 200 = 0 → s1 = 10 and s2 = 20

  s    s  s 2 + 30 s + 200 = ( s + 10)( s + 20)


G1 ( s ) =  ( 4) −   s  + 1  = 0 −   ( s ) +   + 1 
 8      8 
32 ( s + 2) 32 ( s + 2)
G2 ( s ) = =
  s 
G1 ( s ) = 0 −  90o +   + 1  ( 2
s s + 30 s + 200 ) s ( s + 10)( s + 20)
  8 
622 10  Frequency Response

Fig. 10.64   Phase-angle Bode ωc = 10


plot of the factors of and the Corner Frequencies ωc = 2 ωc = 20
complete transfer function,
G2 ( s ) = 32( s + 2) / ( s ( s + 10)( s + 20))
o
90 1.57
(s + 2)

( s_____
+ 10 )
1

( s_____
+ 20 )
1
o
0 0

G( jω) ( )
1
_
s Rad
o
-90 -1.57
G2(s)

G2 (s) =
( )
0.32
s
___
2
___________
+1
-180
o
-3.14
( )( )
s
s ___
10
+1
20
s
___
+1

0.1 1 10 100 1,000


rad
Input Frequency ω, ___
sec

Fig. 10.65   Phase-angle Bode Corner Frequencies ωc = 2 ωc = 20


plot of the factors of and the
complete transfer function,
s
G2 ( s ) = 132 s /(( s + 2)( s 2 + 30 s + 400)) o
90 1.57
G3(s)

G( jω) 0o 0 Rad
( 1
____
s+2 )
o
( _______1
s 2 + 30s + 400 )
-90 -1.57

132 s
__________
G 3(s) =
o (s + 2)(s 2 + 30s + 400)
-180 -3.14
0.01 0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

 32 ( s + 2)  Check the eigenvalues of the second-order factor. The ei-


G2 ( s ) =    genvalues are complex conjugates. Keep the second-order
 s ( s + 10)( s + 20) 
polynomial in its current form.
=  (32 ( s + 2)) −  ( s ( s + 10)( s + 20))
s 2 + 30 s + 400 = 0 → s1 , s2 = σ ± j w d = −15 ± j13.2
= (  32 +  ( s + 2))
− (  s +  ( s + 10) +  ( s + 20)) Determine its ideal, undamped natural frequency, ωn, which
(
= ( 0 +  ( s + 2)) − 90 +  ( s + 10) +  ( s + 20)
o
) is its corner frequency, and damping ratio, ζ :

(
G2 ( s ) =  ( s + 2) − 90o +  ( s + 10) +  ( s + 20) ) s 2 + 30 s + 400 = s 2 + 2ζw n s + w n2
30
10.6.7.3  Example Three: Phase-Angle Bode Plot w n = 200 = 20 → ζ = = 0.75
2w n
132 s
Phase-angle Bode plot of G3 ( s ) = ,
( s + 2)( s 2 + 30 s + 400)
Fig. 10.65.
10.7  Nyquist Polar Plots 623

o
  180
132 s
G3 ( s ) =   
(
 ( s + 2) s + 30 s + 400
2
) 
90
o

( (
=  (132 s ) −  ( s + 2) s 2 + 30 s + 400 )) arg(G( jω)) 0o
(
= 132 0 +  s )
( ))
o
-90
(
−  ( s + 2) +  s 2 + 30s + 400

( ( ))
o
-180
G3 ( s ) = 90o −  ( s + 2) +  s 2 + 30s + 400
0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec
10.6.8  A Complex Number’s Argument
Fig. 10.66   Mathcad’s arg() function, and MATLAB’s angle() function
The “argument” of a complex number is the “primary,” or report the angle of a complex number as ± 180°
smallest angle needed to locate the direction of the com-
plex number, when it is represented in magnitude and
angle, or magnitude and complex exponential unit vector 10.7  Nyquist Polar Plots
form. Unfortunately, the primary angle can be defined as
either −180o ≤ G ( jw ) ≤ 180o , or as 0o ≤ G ( jw ) < 360o. Nyquist polar plots, named for Harry Nyquist (1889–1976),
Mathcad’s function arg() and MATLAB’s function angle() present the frequency response of a system, by plotting the
both use the definition −180o ≤ G ( jw ) ≤ 180o. This is fine result of evaluating transfer function, G( s), for s = jω in Car-
for plotting the phase angle of individual factors, since the tesian coordinates on the “G( s) plane.” Although we speak
maximum possible phase angle of a factor is ± 180°. How- of “the” complex plane, every complex function is entitled
ever, this definition creates errors, when the phase angle of a its own plane to plot the result of evaluating it, hence the
transfer function exceeds 180°. This situation occurs, when G(s) plane.
the “net order” of the transfer function, i.e., the order of the Bode solved the problem of presenting three variables,
denominator minus the order of the numerator, is third order the input frequency, ω, G ( jw ) , and G ( jw ) , on two-
or above. Erroneous phase-angle plots are easy to spot, be- dimensional paper, by repeating the input frequency as the
cause they have a step discontinuity of 180°. For example, abscissa on two plots. Nyquist used the alternative approach
the argument calculated by Mathcad of the fourth-order of parameterizing (marking) the G( jω) curve with the input
transfer function frequency. The axes of a Nyquist plot are the real and imagi-
nary components of the transfer function being evaluated.
1
G (s) = The transfer function is evaluated over a frequency range or
( )(
s 2 + s + 4 s 2 + 20 s + 900 ) spectrum, as with Bode plots. The resulting Nyquist polar
plot curve can be interpreted as either Cartesian coordinates,
is plotted in Fig. 10.66. To avoid errors, phase-angle plots or as the trace of the tip of a complex exponential, repre-
created in Mathcad or MATLAB must use superposition (or sented by the vectors drawn from the origin of the G( s) plane
summation) of the phase angles of the individual factors. to the curve in Figs. 10.67 and 10.68.

Fig. 10.67   Nyquist polar plot 0.2 1


________
of G ( s ) = 1/( s 2 + s + 3) G(s) =
s2 + s + 3
ω=5 ω=0
Im (G( jω) ) 0
ω = 0.5

-0.2 ω=1

-0.4
ω=2
ω = 1.5
-0.6
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
Re(G( jω))
624 10  Frequency Response

Fig. 10.68   Nyquist polar 0.6


plot of G ( s ) = s /( s 2 + s + 3)
ω=1 ω = 1.5
0.4

0.2
ω = 0.5
Im (G( jω)) 0
ω=0
s
ω = 10 G(s)= ________
-0.2 s2+ s + 3
ω=5
-0.4 ω=2

-0.6
-0.4 -0.2 0 0.2 0.4 0.6 0.8 1.0 1.2

Re (G( jω))

Fig. 10.69   Portion of


the Nyquist polar plot of o
G( j0) = 0
G ( s ) = 1/( s 2 + s + 3), showing
ω = 10 ω=0
the steady-state response to the 0
low and high input frequencies
ω=5
Im (G( jω))
ω = 0.5
o
G( j¥) = -180

ω=1
-0.2 s
G(s)= ________
2
s +s+3
0 0.2 0.4
Re (G( jω))

The Nyquist plots Figs. 10.67 and 10.68 are of the trans- we could not estimate the angle of the final segment visu-
fer functions for the spring force and the mass velocity of ally, without greatly enlarging the plot. In contrast, the log-
the spring-mass-damper system with the parameter values, magnitude Bode plot enhances small magnitude features,
M = 1, b = 1, and K = 3. which is very helpful, when using frequency response data
to determine a system’s transfer function, i.e., system “iden-
FK ( s ) 1 V1g ( s ) s tification”.
= and =
F (s) s2 + s + 3 F (s) s2 + s + 3 The small scale of Nyquist plots at high frequencies must
be weighed against the log scale of Bode plots producing a
These values yield an underdamped system with the eigenvalues, false sense of precision. Although a frequency of 10 rad/sec
s1 , s2 = −0.5 ± j1.66 , the natural frequency, w n = 3 = 1.73, is usually well within the capability of experimental appara-
and the damping ratio, ζ = 0.289 . tus, frequency response data are limited by the power of the
An advantage of Nyquist plots is the magnitude and angle input (or excitation), and the signal to noise ratio of the out-
of the transfer function evaluated for an input frequency are put variable. Depending on its mass, or mass moment of in-
shown to true linear scale. The true-scale magnitude is also ertia, an input frequency of 10 rad/sec may be at the limit of
a disadvantage of Nyquist plots. The plot loses resolution, experimental feasibility for a mechanical system described
when the magnitude becomes small. For example, upon by the transfer function, G ( s ) = 1 / ( s 2 + s + 3).
enlarging the Nyquist plot of G ( s ) = 1/ ( s 2 + s + 3), per A second drawback of Nyquist polar plot is that plots of
Fig.  10.69, all that can be determined above the input fre- numerator and denominator factors not only lack the sym-
quency of w = 10 , is that the curve approaches the origin at metry found in Bode plots, they do not even resemble one
an angle, −180o ≤ G ( jw ) ≤ −160o , as the input frequency another. A plot of a factor, say, ( τs + 1), is completely dif-
approaches infinity. Although we know that the angle of the ferent, if the factor is in the numerator rather that the de-
second-order transfer function approaches the origin at the nominator. It is only feasible to sketch Nyquist plots for very
angle of −180°, as the input frequency approaches infinity, simple cases. Recall that Bode plots contain mirror image
Summary 625

Fig. 10.70   Nyquist polar plots. 1


a Differentiator, G ( s ) = s. a G(s) = s
b G(s) = __
b Integrator, G ( s ) = 1/s s
1
10 ω = 10 0
ω = 1 ω = 0.75
8 -2 ω = 0.5
ω = 7.5 ω = 0.4
ω = 0.3
6 -4
Im (G( jω)) ω=5 Im (G( jω)) ω = 0.2
4 ω=4 -6
ω=3
2 ω=2 -8
ω=1
0 ω=0 -10 ω = 0.1
-1 -2 0 2 -2 0 2
Re (G( jω)) Re (G( jω))

Fig. 10.71   Nyquist polar plots


of first-order factors. a Numera- a 5
b
tor factor, G ( s ) = s + 1. b Denom- 0.4 o
Approaches the origin at -90
inator factor, G ( s ) = 1/ ( s + 1) as ω approaches infinity
4 ω=4 0.2

3
G(s) = s + 1 Im (G( jω)) 0 ω=0
ω =10 ω = 0.1
Im (G( jω)) -0.2 ω=4 ω = 0.25
2 ω=2 ω=2
-0.4 ω = 0.5
1 ω=1 -0.6 ω=1
ω = 0.5 1
____
ω = 0.25 ω = 0 -0.8 G(s) =
0 s+1
-1.0
0 1 2 -0.2 0 0.2 0.4 0.6 0.8 1.0 1.2
Re (G( jω)) Re (G( jω))

symmetry of numerator and denominator factors for both factor approaches zero at − 90°, and the second-order factor
the log-magnitude and phase-angle plots. The contribution, approaches zero at − 180°, respectively, as factors are domi-
a factor in the numerator, is the mirror image, relative to the nated by the frequency term.
log-magnitude = 0 line, and the phase angle = 0° line of the
same factor in the denominator. There is no mirror image
symmetry between numerator and denominator factors in Summary
Nyquist plots.
The only numerator and denominator factors with similar Sinusoidal inputs to energetic systems are common. Wave
shapes are differentiators and integrators, Fig. 10.70, since phenomena and rotation provide the ultimate source of sinu-
they both appear as straight lines on the imaginary axis. They soids. Any periodic input can be represented as a summation
do not have mirror image symmetry. They are reciprocals. of sinusoids at different frequencies with the Fourier series.
The magnitude of the integrator approaches infinity, as the Both the transient and steady-state response of an ener-
input frequency approaches zero. getic system to a sinusoidal input are important to consider
The Nyquist plots for first-order factors, Fig. 10.71, and in machine design.
second-order factors, Figs. 10.72 and 10.73, share similari-
ties, in that the magnitudes of the numerator factors head Transient Sinusoidal Response The maximum or mini-
to infinity with increasing frequency, as the magnitudes of mum value of the power variables in a system not subject
the denominators head to zero. The first-order denominator to resonance occurs during the transient period of sinusoidal
626 10  Frequency Response

Fig. 10.72   Nyquist polar plot of 80


the second-order numerator fac- G(s) = s2 + 4s + 64
tor, G ( s ) = s 2 + 4s + 64
60

ω = 10
Im (G( jω)) 40
ω = 7.5
ω=5
20 ω=4
ω=3 ω=2
ω=1 ω=0
0
-80 -60 -40 -20 0 20 40 60 80
Re (G( jω))

0.005 Bode plots and Nyquist polar plots are plots of the steady-
ω=0 state frequency response of a system. The corner frequencies
0 ω=1 of first-order factors are the inverse of their time constant,
ω=2 ω=3
Approaches the origin
w c = 1/τ , which equals the magnitude of the pole, or zero,
ω=4
ο
at -180 when ω → ¥ of the factor. The corner frequency of a second-order fac-
-0.01 ω=5 tor is its natural frequency. The phase angles of integrators
Im (G( jω)) ω = 10 1
__________
G(s) = 2 and differentiators are constants, − 90° and + 90°, respec-
s + 4s + 64
tively. The maximum change, or transition, of a first-order
-0.02
factor’s phase angle is ± 90° and of a second-order factor is
Crosses imaginary axis
at -90ο when ω = ω n ± 180°. Both the first- and second-order phase-angle curves
pass through half of their total transition at the factor’s cor-
-0.03 ner frequency. The phase angles of first-order factors have a
ω = 7.5
single shape, allowing them to be sketched accurately. The
-0.01 0 0.01 0.02 shape of the phase-angle Bode plot of a second-order factor
Re (G( jω)) is a function of the damping ratio. It requires an exemplar to
sketch accurately. Bode log-magnitude plots can be sketched
Fig. 10.73   Nyquist polar plot of the second-order numerator factor, to reasonable accuracy by using asymptotic approximation,
G ( s ) = 1/( s 2 + 4 s + 64) except that resonant peaks, if any, are omitted.
Nyquist plots are true, linear sketches of the system’s
A sin(ωt) A|G(jω) |sin(ωt+φ) transfer function, evaluated for s = jω. They must be param-
G(s) eterized by the input frequency. They cannot be sketched ac-
where φ = G(jω) curately.
Fig. 10.18  Steady-state frequency response relationship

Problems
excitation. Many failures occur during startup and shutdown
of systems. Calculation of the transient response of a sys- Problem 10.1 Fig P10.1 shows a translational mechanical
tem to a sinusoidal input is usually accomplished with either system. Mass M is supported by frictionless rollers. A dash-
Laplace transforms or state-space. pot with damping b and a spring with spring constant K are
attached between the mass and ground. The mass is acted
Steady-State Response The steady-state response of a upon by a sinusoidal force, F( t). The parameter values are
system to a sinusoidal input is easy to calculate from the M = 3, b = 2, and K = 100
system’s transfer function, using the frequency response 10.1.a Derive the system equation for:
relationship, Fig. 10.18. i The velocity of the mass.
The frequency response relationship only applies to the ii The force acting through the spring.
steady-state response. It represents only the particular re- iii The force acting to accelerate the mass.
sponse of a system to a sinusoidal input. The natural or ho- 10.1.b Derive the transfer function for the system equation
mogeneous response is not included. of part a and plot the poles and zeros (if any) of trans-
fer function on the s-plane.
Problems 627

x,v L
b

F(t) v(t) C R
M

K Fig. P10.3   RLC electric circuit

Frictionless sinusoidal torque, T( t). The parameter values are J = 5, b = 4,
rollers and K = 1,000.
10.2.a Derive the system equation for:
Fig. P10.1   Translational spring-mass-damper system acted on by the i The angular velocity of the flywheel.
force, F( t) ii The torque acting through the spring.
iii The torque acting to accelerate the flywheel.
Shaft rigidly attached 10.2.b Derive the transfer function for the system equation
part a and plot the poles and zeros (if any) of transfer
Torsion spring K
Flywheel
function on the s-plane.
rotational inertia J 10.2.c Determine the steady-state response of the transfer
functions of part b for:
i T (t ) = 10 N · m sin(7 t )
ii T (t ) = 10 N · m sin(14t )
iii T (t ) = 10 N · m sin(21t )
Torque 10.2.d Formulate the state-space representation of the sys-
input T(t) tem with the output variables:
i The angular velocity of the flywheel.
Hydrodynamic ii The torque acting through the spring.
bearing damping b
iii The torque acting to accelerate the flywheel.
10.2.e Solve the state and output equations of part d for the
Fig. P10.2   Rotational system consisting of a torque source, a flywheel
with damping, and a torsion spring sinusoidal inputs of part c for the duration, t ≈ 10τ ,
1
where τ = .
σ min
10.1.c Determine the steady-state response of the transfer
functions of part b for: Problem 10.3 Fig. P. 10.3 shows an RLC electric circuit
i F (t ) = 10 N sin(3t ) with a voltage source. The input is a sinusoidal voltage, v( t).
ii F (t ) = 10 N sin(6t ) The parameter values are L = 0.01 H, C = 4 × 10 −6 F, and
iii F (t ) = 10 N sin(12t ) R = 0.1 Ω.
10.1.d Formulate the state-space representation of the sys- 10.3.a Derive the system equation for:
tem with the output variables: i The current through the inductor.
i The velocity of the mass. ii The voltage across the capacitor.
ii The force acting through the spring. iii The current from the voltage source.
iii The force acting to accelerate the mass. 10.3.b Derive the transfer function for the system equation
10.1.e Solve the state and output equations of part d for the of part a and plot the poles and zeros (if any) of trans-
sinusoidal inputs of part c for the duration, t ≈ 10τ , fer function on the s-plane.
1 10.3.c Calculate the ideal, undamped, natural frequency, ωn,
where τ = .
σ min and the damping ratio, ζ, of the circuit.
10.3.d Determine the steady-state response of the transfer
Problem 10.2 Fig. P 10.2 shows a rotational mechanical functions of part b using the natural frequency deter-
system. The flywheel has rotational inertia J. The flywheel’s mined in part c for:
shaft is supported by a hydrodynamic bearing with damp-  wn 
i v(t ) = 24 VDC sin  t 
ing b and attached to a torsion spring with spring constant  2 
K. The other end of the torsion spring is rigidly attached to ii v(t ) = 24 VDC sin(w n t )
the machine frame, which is ground. The system is driven a iii v(t ) = 24 VDC sin(2w n t )
628 10  Frequency Response

10.4.d Determine the steady-state response of the transfer


C functions of part b using the natural frequency deter-
I mined in part c for
R1  wn 
P(t) i p (t ) = 3, 000 psi sin  t 
vent to Pump  2 
atmosphere R2
ii p (t ) = 3, 000 psi sin(w n t )
Fluid iii p (t ) = 3, 000 psi sin(2w n t )
Reservoir
10.4.e Formulate the state-space representation of the sys-
tem with the output variables:
Fig. P10.4   Schematic of a fluid system: i The volume flow rate through the fluid inertance.
ii The pressure in the fluid capacitance.
iii The volume flow rate into to the fluid capacitance.
10.3.e Formulate the state-space representation of the sys- 10.4.f Solve the state and output equations of part e for the
tem with the output variables: sinusoidal inputs of part c for the duration, t ≈ 10τ ,
i The current through the inductor. 1
where τ = .
ii The voltage across the capacitor. σ min
iii The current from the voltage source.
10.3.f Solve the state and output equations of part e for the Problem 10.5 An electromechanical schematic of a DC
sinusoidal inputs of part c for the duration, t ≈ 10τ , motor is shown in Fig. P10.5. The motor’s resistance is
1 R = 4 Ω. The relationship between the motor current and the
where τ = .
σ min motor torque is TM = KT iM , where KT = 8 N · m/A. The mo-
tor’s mass moment of inertia is J M = 0.3 kg · m 2 . The motor
Problem 10.4 Fig. P10.4 shows a fluid system with a pump turns a flywheel with mass moment of inertia, J L = 2 kg · m 2,
modeled as a pressure source, two fluid resistances, a fluid and damping b = 0.1 N · m · sec/ rad .
inertance, and a fluid capacitance (accumulator). The param- 10.5.a Derive the system equations for:
eter values are: i. The current from the source.
ii. The back EMF.
MPa MPa MPa · sec 2 m3
R1 = 1 R2 = 20 I = 5 C = 4 iii. The angular velocity of mass moment of
m3 m3 m3 MPa
inertia, JL.
and check their units.
10.4.a Derive the system equation for 10.5.b Derive the transfer function for the system equation
i The volume flow rate through the fluid inertance. of part a and plot the poles and zeros (if any) of trans-
ii The pressure in the fluid capacitance. fer function on the s-plane.
iii The volume flow rate into to the fluid capacitance. 10.5.c Determine the steady-state response of the transfer
10.4.b Derive the transfer function for the system equation functions of part b for:
of part a and plot the poles and zeros (if any) of trans- i v(t ) = 24 VDC sin(3.5t )
fer function on the s-plane. ii v(t ) = 24 VDC sin(7t )
10.4.c Calculate the ideal, undamped, natural frequency, ωn, iii v(t ) = 24 VDC sin(14t )
and the damping ratio, ζ, of the fluid system. 10.5.d Formulate the state-space representation of the sys-
tem with the output variables:
i. The current from the source.

Fig. P10.5   Schematic of a DC Flywheel Electric Motor


motor with an inertial load and Rotational Inertia JL Transduction KT
driven by a voltage source
Rotational Inertia J M

Electrical Resistance R
Bearings
Damping b Voltage Source v(t)
Problems 629

Rigid Attachement Angular Velocity Source, Ω(t)


Compliant Shaft K Compliant Shaft
Torsional Spring K
Gear Shaft rotates in
N2 teeth ideal bearing Pinion, N1 teeth

Pinion Rack, N2 teeth


N1 teeth per inch
Mass M

Ω(t)
Inertia J
Fluid coupling
Gear
Damping b N3 teeth
Lubricant Film
Fig. P10.6   Rotational mechanical system Damping b

ii. The back EMF. Fig. P10.7   Hybrid rotational-translational mechanical system
iii. The angular velocity of mass moment of
inertia, JL.
10.6.e Solve the state and output equations of part d for the
10.5.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t ≈ 10τ ,
sinusoidal inputs of part c for the duration, t ≈ 10τ ,
1
1 where τ =  .
where τ =  . σ min
σ min
Problem 10.7 The rotational-translational mechanical sys-
Problem 10.6 The rotational mechanical system shown in tem, shown schematically in Fig. P10.7, consists of an angu-
Fig. P10.6 consists of angular velocity source, Ω( t), driv- lar velocity source, Ω(t), acting on a compliant shaft modeled
ing a fluid coupling with damping, b = 80 N · m · sec /rad . as a torsion spring with spring constant, K = 120 N · m/rad .
The output shaft of the fluid coupling drives a pinion with The spring drives a pinion with N1 = 12 teeth, engaged in a
N1 teeth. The pinion with N1 = 10 teeth is meshed with rack with N 2 = 3 teeth per inch. The rack is attached to mass,
Gear Two with N 2 = 20 teeth and Gear Three with N 3 = 30 M = 5 kg . The rack and mass both slide on a lubricant film
teeth. Gear Two is mounted on a compliant shaft with spring with damping, b = 3 N ⋅ m ⋅ sec /rad .
constant, K = 600 N · m/rad . The other end of the compliant 10.7.a Derive the system equation for:
shaft is attached rigidly to the machine frame. Gear Three i The torque acting through the compliant shaft, K.
is mounted on a rigid shaft which carries rotational inertia, ii The velocity of mass, M.
J = 6 kg · m 2 , on an ideal, frictionless bearing. iii The force acting to accelerate mass M.
10.6.a Derive the system equation for: 10.7.b Derive the transfer function for the system equation
i The angular velocity of inertia, J. of part a and plot the poles and zeros (if any) of trans-
ii The torque acting through the compliant shaft. fer function on the s-plane.
iii The torque applied by the angular velocity source. 10.7.c Determine the steady-state response of the transfer
10.6.b Derive the transfer function for the system equation functions of part b for:
of part a and plot the poles and zeros (if any) of trans- rad
i Ω (t ) = 10 sin(24t )
fer function on the s-plane. sec
10.6.c Determine the steady-state response of the transfer rad
ii Ω (t ) = 10 sin(48t )
functions of part b for: sec
rad rad
i Ω (t ) = 10 sin(0.1t ) iii Ω (t ) = 10 sin(96t )
sec sec
rad 10.7.d Formulate the state-space representation of the sys-
ii Ω (t ) = 10 sin(0.5t )
sec tem with the output variables of part a.
rad 10.7.e Solve the state and output equations of part d for the
iii Ω (t ) = 10 sin(t )
sec sinusoidal inputs of part c for the duration, t ≈ 10τ ,
10.6.d Formulate the state-space representation of the sys- 1 .
where τ =
tem with the output variables of part a. σ min
630 10  Frequency Response

Problem 10.8 Plot the log-magnitude and phase-angle Bode 8(60 s 2 + 7800 s )
plots, using MATLAB or Mathcad, for transfer functions: iii G ( s ) =
s 3 + 320 s 2 + 19, 900 s + 168, 000
50 ( s + 3)
i G ( s ) = iv G ( s ) = 2
s ( s + 25) ( s + 5s + 4)( s 2 + 4 s + 8)
11
ii G ( s ) = 8(3s 2 + 0.6 s + 3)
s ( s + 3)( s + 16) v G ( s ) =
320 s (2 s + 50)( s 2 + 20 s + 1200)
iii G ( s ) =
( s + 4)( s 2 + 3s + 55)
Problem 10.12 Plot the Nyquist polar plot, using MATLAB
4 s ( s + 10) or Mathcad, for the transfer functions:
iv G ( s ) =
( s + 20)(4 s 2 + 8s + 64)
50
1 i G (s) =
v G ( s ) = s ( s + 25)
( s + 1) ( s 2 + s + 2)
2
11
ii G ( s ) =
Problem 10.9 Plot the log-magnitude and phase-angle Bode s ( s + 3)( s + 16)
plots, using MATLAB or Mathcad, for transfer functions: 320
iii G ( s ) =
28(8s 2 + 96 s + 1, 280) ( s + 4)( s 2 + 3s + 55)
i G (s) =
s ( s 2 + 112 s + 262) 4 s ( s + 10)
iv G ( s ) =
239 s 2 + 5497 s ( s + 20)(4 s 2 + 8s + 64)
ii G ( s ) = 1
0.5s + 11.8s 2 + 386 s + 1260
3
v G ( s ) =
2 ( s + 1) 2 ( s 2 + s + 2)
8(60 s + 7800 s )
iii G ( s ) = Problem 10.13 Plot the Nyquist polar plot, using MATLAB
s 3 + 320 s 2 + 19, 900 s + 168, 000
or Mathcad for the transfer functions:
( s + 3)
iv G ( s ) = 28(8s 2 + 96 s + 1, 280)
( s + 5s + 4)( s 2 + 4 s + 8)
2
i G (s) =
s ( s 2 + 112 s + 262)
8(3s 2 + 0.6 s + 3)
v G ( s ) = 239 s 2 + 5497 s
s (2 s + 50)( s 2 + 20 s + 1200) ii G ( s ) =
0.5s + 11.8s 2 + 386 s + 1260
3

Problem 10.10 For transfer functions below: 8(60 s 2 + 7800 s )


10.10.a Sketch the log-magnitude vs. frequency plot using iii G ( s ) =
s 3 + 320 s 2 + 19, 900 s + 168, 000
the asymptotic approximation.
( s + 3)
10.10.b Sketch the phase angle vs. frequency plot. iv G ( s ) = 2
( s + 5s + 4)( s 2 + 4 s + 8)
50
i G ( s ) = 8(3s 2 + 0.6 s + 3)
s ( s + 25) v G (s) =
11 s (2 s + 50)( s 2 + 20 s + 1200)
ii G ( s ) =
s ( s + 3)( s + 16)
320 Problem 10.14 Determine the transfer function of the
iii G ( s ) = system, which produced the Bode plots in Fig. P10.14.
( s + 4)( s 2 + 3s + 55)
4 s ( s + 10) Problem 10.15 Determine the transfer function of the
iv G ( s ) =
( s + 20)(4 s 2 + 8s + 64) system, which produced the Bode plots in Fig. P10.15.
1
v G ( s ) =
( s + 1) 2 ( s 2 + s + 2) Problem 10.16  Determine the transfer function of the
system, which produced the Bode plots in Fig. P10.16.
Problem 10.11 For transfer functions below:
10.11.a Sketch the log-magnitude vs. frequency plot using Problem 10.17 Estimate the transfer function that produced
the asymptotic approximation. the Nyquist plot in Fig. P10.17.
10.11.b Sketch the phase angle vs. frequency plots.
28(8s 2 + 96 s + 1, 280)
i G (s) = Chapter 10 Appendix
s ( s 2 + 112 s + 262)
239 s 2 + 5497 s
ii G ( s ) = Drawing Bode Plots in Mathcad
0.5s + 11.8s 2 + 386 s + 1260
3

Determine the desired input frequency range (spectrum). If in


doubt, start with a range which extends from approximately
Chapter 10 Appendix 631

Fig. P10.14   Bode plots of an 6 120


unknown system

4 80

2 40
log|G( jω)| dB
0 0

-2 -40

-4 -80
0.01 0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec
o
-45

-90o

o
G( jω) -135

o
-180

-225 o
0.01 0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

two orders of magnitude below the lowest corner frequency Where low is the logarithm of the lower limit of the frequen-
to two orders of magnitude above the highest corner fre- cy range and high is the logarithm of the upper limit of the
quency. For example, if the system’s corner frequencies were frequency range. Although Mathcad has the imaginary num-
0.8 rad/sec and 70 rad/sec, two orders of magnitude below ber i, the syntax to multiply by the imaginary number it must
is approximately 0.01 rad/sec and two orders of magnitude be written as 1i. It is easiest to define j.
above is approximately 10,000 rad/sec. ( )
Log-magnitude plots are log G ( j · w p ) using the mag-
The values of frequency within the range for which the nitude bars created by typing the single vertical bar above
log-magnitude and phase-angle plots are calculated should the forward slash. Be careful not to select the determinant
be spaced logarithmically to produce smooth curves. One operator from the matrix pallet which has the same symbol
thousand points is more than enough for a smooth curve. of two vertical bars. The code to calculate the log-magni-
Calculate the frequency values as shown in the Mathcad tude of the transfer function can be in the place holder of
code below: the ordinate (vertical axis) of the plot or the log-magnitude
can be calculated and assigned to a vector variable, say
j := −1 LogMagp, before plotting. Remember to multiply the fre-
low := −2 quency by the imaginary number, j, to evaluate the transfer
high := 3 function, G( j · w p).
Calculate the phase angle, which is the angle or the ar-
p := 0..1000
gument of G( j · w p ) with the arg() function. The arg() func-
 p 
low + ( high − low)·
w p := 10  1000  tion returns the “principle” angle of a complex number
632 10  Frequency Response

Fig. P10.15   Bode plots of an 1 20


unknown system

0 0

-1 -20
log|G( jω)| dB
-2 -40

-3 -60

-4 -80
0.01 0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec
o
45

o
0

o
-45
G( jω)
-90o

o
-135

o
-180
0.01 0.1 1 10 100 1,000
rad
Input Frequency ω, ___
sec

in radians. The principle angle of a complex number is Drawing Bode Plots in MATLAB
between −π ≤ G ( s ) ≤ π It is convention to plot the phase
angle in degrees versus the excitation (input) frequency MATLAB’s Control Tool Box, an add-on package, contains
in rad/sec, so multiply the argument by 180/π to yield a Bode-plot function. What follows is the procedure for cre-
degrees. Right click within the plot to access the Format ating Bode plots without the Control Tool Box.
menu. Format the frequency (horizontal) axis as a loga- The MATLAB function for creating a semilogarithmic plot
rithmic scale. with the x-axis logarithmic and the y-axis linear is semilogx().
Its argument list is the same as the plot() function. It expects
pairs of data vectors ordered as semilogx (xdata, ydata). The
Drawing Nyquist Plots in Mathcad xdata vectors provided by the user are not logarithms. The
function calculates the log(xdata) to create the plot.
Nyquist plots are direct plots of the complex value of a trans- The first step in creating a Bode plot is to determine the
fer function evaluated substituting for the Laplace variable desired input frequency range (spectrum). If in doubt, start
s = jω. Use Mathcad’s Re() and Im() functions to extract the with a range which extends from approximately two orders
real and imaginary components from the complex number. of magnitude below the lowest corner frequency to two or-
The Im() returns a purely real number with the correct sign. ders of magnitude above the highest corner frequency. For
The Nyquist plot is then produced as an x–y plot with the x- example, if the system’s corner frequencies were 0.8 rad/sec
axis Re(G ( j · w p )) and the y-axis Im(G ( j · w p )) . and 70 rad/sec, two orders of magnitude below is approxi-
mately 0.01 rad/sec and two orders of magnitude above is
approximately 10,000 rad/sec.
Chapter 10 Appendix 633

Fig. P10.16   Bode plots of an 2 40


unknown system

1 20

log|G( jω)| 0 0 dB

-1 -20

-2 -40
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec
o
135

o
90

o
45

G( jω) 0

o
-45

o
-90
o
-135
0.1 1 10 100 1,000 10,000
rad
Input Frequency ω, ___
sec

The values of frequency within the range for which the % MATLAB's common log function is log10( );
log-magnitude and phase-angle plots are calculated should logmag(p) = log10(abs(100/((j*w(p))^2+5*j*w(p)+100)));
be spaced logarithmically to produce smooth curves. One end
thousand points is more than enough for a smooth curve. The semilogx(w,logmag) % Create semilog plot
MATLAB code below spaces the frequencies at which val- grid on % Draw grid on the plot
ues are calculated logarithmically, calculates the log-magni-
tude values, and creates the plot: Although MATLAB has the imaginary number defined as
i and j, it advises the user to use the syntax 1i “for speed
% Semilog plot. The Bode log magnitude and robustness.” It is easiest to define j using the square root
%of the transfer function: G(s)=100/(s^2+5*s+100) function sqrt(). The magnitude of a complex number is cal-
j=sqrt(-1) % Define the imaginary number j culated using MATLAB’s abs() function, where abs is short
low = -2 % Exponent of the lower limit of freq range for absolute value.
high = 3 % Exponent of the upper limit of freq range Calculate the phase angle, using MATLAB’s angle() func-
range = high - low % Number of decades in freq range tion. The angle() function returns the “principle” angle of a
N = 1000 % Number of calculations in plot complex number in radians. The principle angle of a com-
w = zeros(N); % Preallocate and zero freq vector w plex number is between −π ≤ G ( s ) ≤ π . It is convention to
for p = 1:N;
plot the phase angle in degrees versus the excitation (input)
% Calculate the freq value w(p);
frequency in rad/sec, so multiply the argument by 180/π to
w(p) = 10^(low+range*(p/N));
yield degrees. The following MATLAB code calculates and
plots a phase-angle plot:
634 10  Frequency Response

Fig P10.17   Nyquist polar plot of


an unknown system Im
-2 -1 0 1 2 3 4
| o Re
95 -21.8
rad
___ |1.
ω 3 = 15 sec |3.7
1| rad
___
-144 -j
o
ω1 = 5 sec

-j2

-j3

-j4

rad -j5
___
ω 2 = 10 sec G(s) plane

-j6

% Semilog plot of the Phase Angle of % Nyquist G(s)=100/(s^2+5*s+100)


% G(s)=100/(s^2+5*s+100) j=sqrt(-1) % Define the imaginary number j
j=sqrt(-1) % Define the imaginary number j low = -2 % Exponent of the lower limit of freq range
low = -2 % Exponent of the lower limit of freq range high = 3 % Exponent of the upper limit of freq range
high = 3 % Exponent of the upper limit of freq range range = high - low % Number of decades in freq range
range = high - low % Number of decades in freq range N = 1000 % Number of calculations in plot
N = 1000 % Number of calculations in plot RealG = zeros(N); % Preallocate and zero RealG vector
w = zeros(N); % Preallocate and zero freq vector ImagG = zeros(N); % Preallocate and zero ImagG vector
phi = zeros(N); % Preallocate and zero logmag vector %Calculation loop
%Calculation loop for p = 1:N;
for p = 1:N; % Calculate value of freq w at a logarithmic spacing;
w(p) = 10^(low+range*(p/N)); w = 10^(low+range*(p/N));
% MATLAB's angle() returns principle angle in rad; % Calculate value of G(j*w);
% MATLAB has pi as a predefined constant; G = 100/((j*w)^2+j*w*5+100);
phi(p) = (180/pi)*angle(100/((j*w(p))^2+5*j*w(p)+100)); RealG(p) = real(G);
end ImagG(p) = imag(G);
semilogx(w,phi) % Create semilog plot end
grid on % Draw grid on the plot plot(RealG,ImagG) % Create semilog plot
grid on % Draw grid on the plot

Drawing Nyquist Plots in MATLAB References and Suggested Reading

Nyquist plots are direct plots of the complex value of a trans- Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs
fer function evaluated substituting for the Laplace variable Ogata K (2009) Modern control engineering 5th edn. Prentice-Hall,
s = jω. MATLAB’s real() and imag() functions extract the Englewood Cliffs
real and imaginary components from a complex number. The Rowell D, Wormley DN (1997) System dynamics: an introduction.
imag() function returns a purely real number with the correct Prentice- Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
sign. The Nyquist plot is then produced as an x–y plot with dynamics. Addison-Wesley, Reading
the x-axis real (G ( j · w p )) and the y-axis imag (G ( j · w p )) .
AC Circuits and Motors
11

Abstract
Alternating current (AC), varying sinusoidally at 50 Hz in Europe and Asia and 60 Hz in
North America and Japan, is the dominant form of electric power. AC provides the continu-
ously changing magnetic flux linkage needed for electrical transformation. Transformation
allows transmission at high voltage and low current, reducing resistive loss, with step-down
transformers near the point of use. Complex impedances are an efficient method for deriv-
ing the transfer function of an electric circuit. Analysis of AC circuits is an application of
frequency response. Phasors are vectors in the complex resistance–reactance plane, which
can represent either variables or operators. As variables, phasors are analogous to complex
exponentials except that their amplitude is the root-mean-squared value rather than the
peak amplitude of the sinusoid. As operators, they are analogous to transfer functions.

11.1 Introduction 11.1.1 Alternating Current

Development of electrical science and technology began with The input to an AC circuit is always a sinusoidally varying
systems powered by voltaic “piles” or batteries which pro- voltage v(t ) = V0 sin(w t ). AC circuit analysis is frequency re-
duced a constant voltage. It was natural that the first electrical sponse analysis. Consequently, AC analyses apply only to
generators (or dynamos) produced constant voltage, yielding the steady-state condition. If we need to know the transient
what is termed “direct” or non-alternating current. Alternat- response of a system driven by AC power, we must use a
ing current (AC) electric motors have powered stationary conventional system dynamics analysis.
machines for over 100 years, progressively replacing expen- By convention, the input voltage is assumed to have zero
sive steam engine driven shaft and belt systems. AC electric phase shift and, when plotted as a vector, lies on positive real
motors are now used in combination with internal combus- axis. Usually, the output variable of interest is the current
tion engines in hybrid automobiles and have replaced internal drawn by the AC circuit. The phase angle ϕ of the current
combustion engines in automobiles designed for commuters. drawn by a system, i (t ) = I 0 sin(w t + φ ), is measured relative
Mechanical engineers must understand AC motors and AC to the input voltage. Current phase leads or lags are created
circuits in order to use them effectively in their designs. by capacitive or inductive energy storage within a circuit.
Alternating current circuits are driven by a sinusoidal For many of our analyses, we will find it convenient to rep-
voltage source and are analyzed in steady-state. Consequent- resent the circuit being driven by the source as a single com-
ly, the analysis of AC circuits, including AC motors, is an plex impedance load. Complex impedances will allow us to
application of frequency response analysis. We will begin work with “phasors.” Phasors are complex exponentials with
by introducing complex impedance, which is a method to magnitudes which are RMS values defined below. The angle
generalize the reduction of an electric circuit or linear graph, of the phasor is the phase shift by itself, equivalent to set-
which allows a graphical reduction of the energetic network ting time t = 0, at an arbitrary moment in steady-state. Conse-
to a transfer function. We will then introduce some of the quently, AC circuit analyses are quasi-steady-state analyses.
terminology and technology of AC motors, including “three-
phase” AC power. Then we conclude with examples of
power calculations for three-phase electric motor.
K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_11, © Springer Science+Business Media New York 2014 635
636 11  AC Circuits and Motors

11.1.2 AC Power we will consider purely resistive circuits. If a circuit has no


energy storage, then the electric power drawn by the circuit
Alternating current is the worldwide standard for the distri- is calculated by the product of current and voltage, P = vi .
bution of electric power. When Thomas Edison (1847–1931) Accordingly, an increased voltage allows a decreased current
developed the first direct current (DC) power generation and to deliver the same power.
distribution system in 1882, it was known that the AC could Recall the relationships between voltage, current, induc-
be transformed by magnetic induction to increase voltages tance L, and flux linkage λ from Sect. 5.7.3,
and lower currents for transmission, reducing resistive power
di
loss. AC power could be used to illuminate the carbon fila- v12 = L (5.45)
ment incandescent lamps, also developed by Edison. How- dt

ever, electric motors at the time were all DC motors. George

Westinghouse (1846–1914) created the first AC power = v12 (5.46)
  dt
system in 1886 for lighting. It was not until Nikola Tesla
(1856–1943), who received the patent for the AC induction
motor in 1888, began working with Westinghouse that AC λ = N φ = Li (5.47)

power gained popularity. The long distance transmission of
hydroelectric power, three phase at 40 Hz was demonstrated AC power’s sinusoidally varying current produces time-
in 1891, Frankfurt, Germany. Westinghouse and Tesla are varying magnetic fluxes φ in coils and time-varying flux
credited with the first large scale hydroelectric power plant, linkages λ. Consequently, by magnetic transformation the
which started operation at Niagara Falls in 1895. It generated voltage of AC power can be “stepped up” for transmission
two-phase power at 25 Hz in the USA. and “stepped down” for use, decreasing voltage and increas-
The United States and Canada now use the frequency of ing current. Although it is possible to step up and step down
60 Hz (cycles/sec) established by Tesla. Elsewhere, AC fre- DC voltages of large power flows, it is more expensive. DC
quency is 50 Hz with a few exceptions. Why we now use electrical energy must be converted to mechanical energy by
60 Hz and the rest of the world uses 50 Hz is a good ques- a DC motor, which then drives a DC generator to produce
tion. The higher frequency is more efficient. (One suggestion DC electrical power at a different voltage. Paired DC motors
why 50 Hz is used outside of North America is that it is more and DC generators are expensive to purchase and to operate,
“metric” than 60 Hz. The rationale for 60 Hz was the need since they need maintenance. AC voltages can be stepped up
to synchronize generating stations within a grid, which was or down using transformers which are far less expensive to
done twice a day using a master clock.) Some AC calcula- build and require little maintenance.
tions are performed with frequency in Hertz, but others will A variety of voltages are used for industrial motors. Large
require angular frequency with units of rad/sec. One conver- motors use voltages greater than 110–120 VAC for a number
sion that is helpful to know is 60Hz = 377rad/sec. of reasons, including economy. The hazard posed by higher
The residential service voltage in North America is voltages is not an issue because large industrial motors are
110–120 VAC. It is a range and does vary with the load on a hardwired by electricians. The following are the typical volt-
system and the distance from the electrical distribution line. ages: 208 VAC, 230 VAC, 460 VAC, 575 VAC, and 2300
Typically, the voltage of 115 VAC is assumed in an analy- VAC. “High voltage” is above 600 VAC. The advantage
sis. This voltage dates back to Edison’s carbon filament of high voltage is reduced current for the same power. An
incandescent lamp. Edison generated 110 VDC to deliver induction motor using 2300 VAC draws only 49 Amps to
100 VDC, allowing for approximately 10% resistive loss in produce 200 hp. In comparison, the same type of motor
transmission. Countries that use 50 Hz also distribute power designed to run at 230 VAC would need 490 Amps, and a
at higher voltage, which compensates for the lower efficien- lot more copper, to produce 200 hp. Industrial motors also
cy of that frequency. The European Union is currently imple- use variable frequency AC current for speed control. AC is
menting a dual voltage system of 240 and 400 VAC with the supplied to the speed control circuitry at 60 Hz. The AC is
lower voltage used for household appliances and lighting, converted to a direct current (DC), and then to the needed
and the higher voltage for large motors. frequency using circuitry known as an “inverter.”
Resistive energy loss during transmission of electric
power is a function of current flow, not voltage.
11.1.3 Root-Mean-Square (RMS) or Effective
PLost = Ri 2 (11.1) Values

We shall see that AC power calculations must consider the AC voltage measurements do not report the true amplitude
phase shift between current and voltage, which occurs when of the sinusoidal voltage or current. The value reported as
there is energy storage in the system. For the time being, volts AC, abbreviated as VAC, represents the “average” or
11.1 Introduction 637

V0 200 a b c
V RMS
100

v(t) 0 VDC 0
VDC
-100

-V0 -200

(n-1)T nT (n+1)T 0 0.01 0.02 0.03 0.04


t, periods t, sec

Fig. 11.1   Root-mean-square voltage VRMS level plotted against Fig. 11.2   Three electrical phases, a, b, and c, of 60 Hz (377 rad/sec)
the sinusoidal voltage. The units VDC mean the true amplitude of the 120 VAC power. Time t = 0 is an arbitrary time after the system has
voltage reached steady-state

“effective” voltage. The “effective” value corresponds to 11.1.4 Sinusoidal and RMS Voltages
the constant, or DC, voltage which would deliver the same
power to a resistance. Calculation of the effective value of Voltages reported as VAC (volts AC) are root-mean-square
an AC voltage cannot be based on the average value of the (RMS) values, and not the peak or true amplitude of the si-
sine wave. The average value of a sinusoidal wave, such nusoidal voltage. The terminology can be confusing because
as v(t ) = V0 sin(w t ), which has no vertical offset from zero, the true amplitude of a sinusoidal voltage is reported as VDC
Fig. 11.1, is zero due to the symmetry of the sine wave. (volts DC) even though it is a sinusoid. VAC units of voltage
Consequently, a straight average of v(t ) = V0 sin(w t ) over with RMS amplitude are always used in AC calculations.
a period is not a useful calculation. Squaring v( t) before
calculating the average eliminates the problem of negative
values over half the period, but also inordinately increases 11.1.5 Three-Phase Alternating Current
the contribution of large numbers to the calculation. A way
of compensating is to take the square root of the average Electric power is generated as “three-phase” alternating cur-
of squared function. This yields the “root-mean-square” or rent, which consists of three sinusoidally varying voltages
RMS value. spaced in time by phase angles that differ by 120 degrees, or
Expressed mathematically, the root-mean-square (RMS) 2π /3 rad. Three-phase power is generated simultaneously in a
value of the v( t) is single generator. Each phase is generated and transmitted in a
T
separate circuit, the windings or coils of which are arranged in
1 the repeating pattern of a, b, c around the circumference of the
VRMS = ∫ (v(t ))2 dt (11.2)
 T 0 generator’s stator. Each of the three phases is transmitted on a
different electrical conductor or “line.” The three sinusoidally
or varying voltages are known as “phases” due to their different
T phase angles. The word phase in this context is similar to its use
1
(V0 sin (w t )) dt
2


VRMS = ∫
T 0
(11.3) in the phrase, “the phase of the moon.” Phase denotes a point
in a cycle. In the case of three-phase current, the sinusoids of
three circuits are shifted one third of the cycle apart. When one
This integral is evaluated in the appendix of this chapter. The of the circuits is arbitrarily chosen as “phase a,” Figs. 11.2 and
result is the effective or root-mean-square (RMS) value of 11.3, it defines the beginning of the cycle. Then the other two
v (t ) = V0 sin (w t ) circuits’ sinusoids are seen as shifted “in phase” from phase a.
An individual line can be used as single-phase power
V0 when its voltage is dropped across the electrical load to a
VRMS = = 0.707 V0 (11.4)
2 “neutral” conductor at ground potential. Alternatively two

or three of the lines can be used together as “polyphase”
The amplitude, or peak value, V0 of the sine can be calculated power. Two phases are delivered to residential electrical ser-
from its root-mean-square value vices. Most small motors in home appliances or hand tools
run on single phase 120 VAC. Large motors, such as central
V0 = 2VRMS = 1.41 VRMS (11.5) air conditioning compressors, and large resistive loads, such
 as electric stoves and electric dryers, use the voltage drop
638 11  AC Circuits and Motors

phase c A sin(ωt) A|G(jω) |sin(ωt+φ)


G(s)
where φ = G(jω)
o
φc = -240
-4π Fig. 11.4   Frequency response relationship
___
= phase a
3

1 R 2 L 3
φ b = -120 o
-4π
___ iL
=
3 iR
phase b i iC C
Fig. 11.3   “Phasor” diagram of three-phase AC voltage. A phasor’s
magnitude equals the RMS or effective magnitude of the AC voltage or
current. The phasors are drawn at their phase angle, corresponding to g
the beginning of an AC cycle
Fig. 11.5   Sinusoidally excited RLC circuit. The sinusoid in the source
symbol denotes AC voltage and that the system is in steady-state
b­ etween the two phases. Similarly, industrial equipment can
be one, two, or three phase. Three-phase AC power is widely
used for industrial motors, as discussed in Sect. 11.6. (or modeled as being linear), (2) the input is a sinusoid, and
Figure  11.2 shows three sinusoidal voltages, a( t), b( t), (3) the system has reached steady-state. AC circuit analysis
and c( t). The peak voltages are V0 = 170 VDC. Since: meets these criteria.
We will introduce the specialized techniques developed
170 VDC for analyzing AC circuits by applying the general purpose
= 120 VAC
2 method we have applied to other energetic systems. Our ex-
ample is the RLC circuit of Chap. 1, Fig. 11.5, now excited
We are free to choose which of the three conductors carries by AC voltage, as indicated by the sinusoid in the source
phase a and to define time t = 0 to be the beginning of phase symbol. If it is powered directly from the “grid” or through
aʼs cycle anytime once the system has reached steady-state. a transformer to step down the applied voltage, the input
In calculations, the voltage of phase a is always assumed to frequency is 60 Hz = 377 rad/sec. The output variable of in-
have the phase angle φa of zero. In Fig. 11.2, the solid trace terest is the current i flowing from the voltage source.
has been named phase a. Following our standard analysis, we write the energetic
equations which describe the system.
a (t ) = V0 sin (w t ) (11.6)
 Compatibility:  v (t ) = v12 + v23 + v3 g
Phase b peaks after or lags behind, phase a by one-third of a
cycle of 2π rads, φb = − 2π /3, Fig. 11.3. Continuity:  i = iR iR = iL iL = iC
diL dv3 g
  2π  Elements:  v12 = RiR v23 = L iC = C
b (t ) = V0 sin  w t −  (11.7) dt dt
 3
  1 2 1
Energy: E sys = E L + EC EL = LiL EC = Cv32g
Phase c lags behind phase a by two-thirds of a cycle, 2 2
φc = −4π /3 .
Recall that our only use for the energy equations has been
  4π  to establish the initial condition of the output variable when
c (t ) = V0 sin  w t − 
 3 (11.8) an input is applied to an energized system. The energy equa-
tions have no use in a steady-state analysis since the effect of
the initial condition of the system has decayed to zero when
the system reaches steady-state.
11.2 Frequency Response of Electric Circuits To use the frequency response relationship, we must
derive the transfer function for the input voltage and the
The frequency response relationship, introduced in output variable, and the current from the voltage source.
Sect. 10.2, Fig. 11.4, applies to all types of energetic sys- Thus far, our procedure for deriving a transfer function has
tems, including electric circuits, if (1) the system is linear been to (1) reduce the equation list to a differential system
11.3  Complex Impedance 639

I s
equation, (2) perform the Laplace transformation on the () =
Cs

I (s)
=
Cs
differential system equation, and (3) create the ratio needed V (s) Cs V (s) CRs + CLs 2 + 1
for the transfer function. We will now vary the sequence CsR + CsLs +
Cs
of operations and perform the Laplace transformation on
energetic equations before we reduce the set of energetic C
I (s) I (s) s
equations. Cs CL
= → =
An unknown, time-domain variable transforms to an un- V ( s ) CLs + CRs + 1
2
V (s) CR 1
s2 + s+
known Laplace-domain variable. We often omit indicating a CL CL
variable is in the time-domain, but we usually indicate that a 1
variable is in the Laplace-domain. I (s) s
= L
(11.11)
V (s) R 1
Compatibility: s2 + s +
L CL

L {v(t )} = L {v 12 + v23 + v3 g }
By performing the Laplace transformation before reducing
V ( s ) = V12 ( s ) + V23 ( s ) + V3 g ( s ) the set of energetic equations, we are able to obtain the trans-
fer function without deriving the differential system equa-
Continuity: tion. This is the essence of complex impedance.

L {i} = L {i } R → I (s) = I R (s)


11.3 Complex Impedance
L {i } = L {i }
R L → I R (s) = I L (s)
L {i } = L {i }
L C → I L ( s ) = IC ( s ) Complex impedance is a generalized, frequency-dependent
resistance. The complex impedance of an electric circuit ele-
ment is created by:
Elements:  L {v } = L { Ri }
12 R → V12 ( s ) = RI R ( s ) 1. Transforming the elemental equation using the Laplace
transform.
 diL  2. Rearranging the equation into the ratio of voltage over
L {v } = L
23 L  → V23 ( s ) = LsI L ( s ) current. The quantity on the other side is the complex im-
 dt 
pedance.
 dv3 g 
L {i } = L
C C  → I C ( s ) = CsV3 g ( s ) V (s)
 dt  ≡ Z ≡ Complex Impedance (11.12)
I (s)

Reduction: Input V( s), Output I( s)
Capital Z is the symbol for complex impedance.
V ( s ) = V12 ( s ) + V23 ( s ) + V3g ( s ) The form of the complex impedance of an electrical ele-
 IC ( s ) ment is easy to remember once it is recognized as the same
V ( s ) = RI R ( s ) + LsI L ( s ) + (11.9) form as electrical resistance.
Cs
The term “complex” refers to the Laplace transforma-
tion of the time-domain differential equations. The Laplace
I (s)
 V ( s ) = RI ( s ) + LsI ( s ) + variable, s = σ + jw , is a complex variable. The term “im-
Cs pedance” connotes a generalized resistance to power flow
 1 through an element, since resistance impedes the flow of
V ( s ) =  R + Ls +  I ( s ) current and power in a circuit.
 Cs 

Output ( s ) I (s) 1
= = (11.10) 11.3.1 Complex Impedance of a Resistor
Input ( s ) V ( s) 1
R + Ls +
Cs
Illustrating with the elemental equation of a resistance

Clear fractions and write the denominator polynomial in v (t ) = Ri (t ) → L {v (t )} = L { Ri (t )} → V (s) = I (s) R


standard form by ordering the terms and clearing the coef-
V (s)
ficient of the highest-order term. = R ≡ ZR (11.13)
I (s)
640 11  AC Circuits and Motors

Fig. 11.6  a Linear graph of


RLC circuit shown schemati-
a R L b ZR ZL
cally in Fig. 11.5, b Linear graph
1 2 3 1 2 3
of the RLC circuit with complex
impedances in place of time
domain elements C ZC

g g g g


The complex impedance of a resistor, ZR, equals its resis- V (s) I (s) 1 1
tance R because the elemental equation is algebraic, not dif- = R ≡ ZR → = = ≡ YR (11.16)
I (s) V (s) R ZR
ferential with respect to time. The Laplace transformation
transforms the variables v(t) and i(t) from the time-domain
to V(s) and I(s) in the Laplace-domain. The complex admittances of a capacitance and an inductance
have the same form as the complex impedance of a resis-
tance.
11.3.2 Complex Impedance of a Capacitor
 V (s) 1 I (s) 1
= ≡ ZC → = Cs = ≡ YC (11.17)
The elemental equations of energy storage elements are trans- I ( s ) Cs V (s) ZC
formed from differential equations to algebraic equations.
The transformation of differentiation with respect to time  V (s) I (s) 1 1
leads to the Laplace variable s being included in the complex = Ls ≡ Z L → = = ≡ YL (11.18)
I (s) V ( s ) Ls Z L
impedance. The complex impedance of a capacitor is:
Complex admittances are useful in reducing a system with a
 number of parallel elements, as we shall see in Sect. 11.3.6.
 dv (t ) 
L {i (t )} = L C  → I ( s ) = CsV ( s )
 dt 
V (s) 1 11.3.5 Reduction of the RLC Circuit Using
= ≡ ZC (11.14) Complex Impedances
I ( s ) Cs
We will illustrate the use of complex impedances as a simpli-
11.3.3 Complex Impedance of an Inductor fication of the reduction of the RLC circuit of Sect. 11.2. In
Sect. 11.3.5, we will introduce an even more efficient graphi-
Inductors store energy in the magnetic field built by their cal reduction method.
current. The time dependency of the flow of energy into and The first step is to replace the elements of the linear graph
out of the magnetic field in the time-domain requires a dif- of the RLC circuit with the corresponding complex imped-
ferential elemental equation. Hence, the complex impedance ance, Fig. 11.6.
of an inductor also includes the Laplace variable s. The key aspect of complex impedance is that all com-
plex impedances can be manipulated as if there were
  di (t )  resistances in a time-domain resistive circuit. Figure 11.6b
L {v (t )} = L L  → V ( s ) = Ls I ( s ) shows the complex impedances of resistance, inductance,
 dt 
and capacitance in series. These three different types of
V (s) complex impedances can be manipulated as if they were all
= Ls ≡ Z L (11.15)
I (s) the complex impedance of resistance. In Chap. 5, we proved
that, in the time-domain, resistances in series sum to form a
single equivalent resistance. We will now demonstrate that
11.3.4 Complex Admittance complex impedances in series and in parallel also sum to the
equivalent complex impedance.
Complex admittance is the inverse of complex impedance,
as conductance is the inverse of resistance. Complex imped- 11.3.5.1 Reduction of Impedances in Series
ances have the form of resistance. Complex admittance has We derived the relationships to replace resistors in parallel
the inverse form. The symbol for complex admittance is Y. or in series with single equivalent resistances in Chap. 5.
Recall that the equivalent element carries the same current
11.3  Complex Impedance 641

Fig. 11.7  a Impedances in


series, b Single equivalent a b
impedance Z1 Z2 Z equiv
1 2 3 1 3

Fig. 11.8  a Impedances in


parallel, b Single equivalent
a Z1 b Z equiv
impedance
1 2 1 2

Z2

(through variable) flow and has the same voltage (across time consuming, however, because the result must be placed
variable) drop across it as the portion of the network it re- over a common denominator and inverted to be useful.
places. The energetic equations for the complex impedances Z1
The energetic equations for the complex impedances Z1 and Z2 in parallel, Fig. 11.8a, are:
and Z2 in series, Fig. 11.7a, are:
Continuity:  I1 ( s ) + I 2 ( s ) ≡ I ( s )
Continuity:  I1 ( s ) = I 2 ( s ) ≡ I ( s )
Compatibility:  V12 ( s ) = V12 ( s )
Compatibility:  V13 ( s ) = V12 ( s ) + V23 ( s )
Element Impedances:
Element  Impedances:  
V12 ( s ) V12 ( s ) V12 ( s )
= Z1 = Z2 = Z Equiv
V12 ( s ) V23 ( s ) V (s) I1 ( s ) I2 (s) I (s)
= Z1   = Z 2   13 = Z Equiv Parallel
I1 ( s ) I2 (s) I (s) Series

Reduction of impedances in parallel to a single equivalent


Reduction of two impedances in series to a single equivalent impedance:
impedance:
V12 ( s ) V12 ( s ) V12 ( s )
V13 ( s ) = V12 ( s ) + V23 ( s ) I1 ( s ) + I 2 ( s ) = I ( s ) → + =
Z1 Z2 Z Equiv
Z REquiv I ( s ) = Z R1 I1 ( s ) + Z R2 I 2 ( s )   1 1  V (s) 1 1 1
V12 ( s )  +  = 12 → = +
  Z1 Z 2  Z Equiv Z Equiv Z1 Z 2
Z Equiv I ( s ) = Z1 I ( s ) + Z 2 I ( s ) parallel

1 Z2 Z Z + Z2
Z Equiv I ( s ) = ( Z1 + Z 2 ) I ( s ) = + 1 = 1
Z parallel Z1Z 2 Z1Z 2 Z1Z 2

Z Equiv = Z1 + Z 2 Z1Z 2
(11.19) Z parallel = (11.21)
series
Z1 + Z 2

This reduction is applicable to any complex impedance. It The reduction of two complex impedances in parallel can be
can also be extended to any number of complex impedances extended to any number of complex impedances in parallel.
in series. Impedances in parallel sum as inverses to yield the inverse of
Complex impedances in series: the equivalent impedance:

Z Series = Z1 + Z 2 +  + Z n (11.20) 1 1 1 1
  = + + + (11.22)
Z parallel Z1 Z 2 Zn

11.3.5.2 Reduction of Impedances in Parallel 1


The effort is in inverting to Z parallel . The ­result for
The reduction of complex impedances in parallel to a single Z parallel
equivalent impedance is straightforward. It is generally more two elements in parallel is easy enough to invert. Inverting
642 11  AC Circuits and Motors

a ZR ZL b The input variable of an AC circuit is always voltage,


1 2 3 1 Vsource( s). Notice that the driving point impedance is the in-
verse of the transfer function for the current from the source,
Isource( s).
ZC Z D.P
1 I
= source
(s) = Cs
Z D.P. Vsource ( s ) LCs 2 + RCs + 1
g g g

Fig. 11.9   Sinusoidally excited RLC circuit. The sinusoid in the source
When the denominator is expressed in standard form by
symbol denotes AC voltage and that the system is in steady-state clearing the highest-order term of its coefficient,

 1 
the sum of three inverted impedances for three elements in 1 I source ( s )  LC  Cs
= =
parallel is more involved. Z D.P. Vsource ( s )  1  LCs 2 + RCs + 1
 
 LC 
1 1 1 1 
= + + 1
s
Z parallel Z1 Z 2 Z3 1 I
= source
(s) = L (11.25)
1 Z 2 Z3 ZZ ZZ Z D.P. Vsource ( s ) R 1
s2 + s +
= + 1 3 + 1 2 L LC
Z parallel Z1Z 2 Z3 Z1Z 2 Z3 Z1Z 2 Z3

1 Z 2 Z3 + Z1Z3 + Z1Z 2 the result, Eq. 11.25, equals the transfer function we derived
= previously, Eq. 11.11.
Z parallel Z1Z 2 Z3

Z1Z 2 Z3
Z parallel = 11.3.7 Graphical Reduction of Networks
Z 2 Z3 + Z1Z3 + Z1Z 2
of Complex Impedances

11.3.6 Driving Point Impedance The reduction linear graph or a circuit in which the elemen-
tal equations are expressed as complex impedances can be
When a system is reduced to a source and a single equiva- performed as a graphical technique. The method is to suc-
lent impedance, the equivalent impedance is called the “driv- cessively combine elements first those in parallel and then
ing point” impedance, since it is the impedance the source those in series, replacing them with equivalent elements.
driving the system “feels.” A driving point impedance is The process is repeated until the network is reduced to the
used to determine the power the source must provide to drive source element and a single, driving point, complex imped-
the system. The definition of driving point impedance is ance. The graphical reduction to the driving point impedance
is quick and easy. The work is in the back-substitution and
Vsource ( s ) clearing fractions. The graphical algorithm is illustrated in
Z D.P. = (11.23)
 I source ( s ) Figs. 11.10 and 11.11.

We will replace the three impedance in series, Fig. 11.9a,


with a single equivalent impedance, Fig. 11.9b, by use of 11.4 Phasors and Phasor Operators
Eq. 11.20. The equivalent impedance represents all of the
system, except the source. The equivalent impedance is the Phasors are vectors used in AC circuit calculations to
driving point impedance, ZD.P. represent the variables, voltage and current. The phasor rep-
resentation of voltage and current differs from the complex
1
Z R + Z L + ZC = Z D.P. → R + Ls + = Z D.P. exponential representation of those variables in just three as-
Cs pects. First, AC analysis is steady-state. The magnitude of
The effort is expressing the driving point impedance as a the phasor is constant. Second, the magnitude used is the
proper ratio. RMS magnitude, not the true magnitude, again, because it
is an AC analysis. All AC calculations are in steady-state.
Cs Cs 1 RCs + LCs 2 + 1
R + Ls + = Z D.P. → = Z D.P. Second, only the phase angle is used as the angle. The time
Cs Cs Cs Cs varying angle, ωt, is not included. Consequently, a phasor
 Vsource ( s ) LCs 2 + RCs + 1 diagram shows the relative position of the variables at the
Z D.P. = = (11.24)
I source ( s ) Cs
11.4  Phasors and Phasor Operators 643

Fig. 11.10  a Initial circuit, b


Parallel impedances Zα and Zβ
a Zα b Z equiv c
1 2 1 2 1
replaced by their equivalent im-
pedance Zequiv, c Series imped-
ances Zequiv and Zγ replaced by Zβ
their equivalent impedance, the Zγ Zγ Z D.P
driving point impedance, ZD.P.

g g g g g

Fig. 11.11  a Initial circuit, b


Parallel impedances Zβ and Zγ
a Zα b Zα c
1 2 1 2 1
replaced by their equivalent im-
pedance Zequiv, c Series imped-
ances Zα and Zequiv replaced by
their equivalent impedance, the Zβ Zγ Z equiv Z D.P
driving point impedance, ZD.P.

g g g g g

instant that the input sinusoid begins a new cycle, the angle
equals an integer multiple n2π. The phasors are vectors of
jΩ
reactance
voltage or current with RMS magnitude vector-like repre-
Z L= jωL
sentations of complex exponential.
ZR = R
11.4.1 Reactance and Resistance Ω, resistance
1
____ -j
___
Phasor operators are vectors used to represent the complex ZC = =
impedance or transfer function of the AC circuit. Repre- jωC ωC
senting complex impedances and complex admittances as
vectors allows graphical computations. To eliminate pos- Fig. 11.12   Phasor operators for the electrical properties of resistance,
capacitance, and inductance
sible confusion with admittance, jY is not used to identify
the imaginary (vertical) axis of a phasor operator’s complex
plane. Unfortunately, the vertical imaginary axis is identified
as jX, which takes some getting used to, or alternatively, by jΩ
its units. The units of the imaginary axis are jΩ and are called reactance
“reactance” to sound similar but distinctive from the units of
the real axis, which are units of resistance, Fig. 11.12. Al- Ω, resistance
though reactance appears as an “imaginary resistance” it is
a physical quantity. Physically, reactance represents energy
temporarily stored in inductance or capacitance and then re-
Z DP = 3 Ω - 1,330 jΩ
turned to the power grid. Complex impedances can be used
to derive transfer functions for either transient analyses or
frequency response analyses. Phasor operators are only used Fig. 11.13   Phasor operator for the driving point impedance of the RLC
in AC circuit calculations. circuit, Eq. 11.24
A complex impedance or admittance is expressed as
a phasor operator by evaluating the impedance or admit-
tance using s = jw , where ω is the frequency of the AC driving point impedance of the entire system, Tables 11.1
circuit, 377 rad/sec, and then plotting the result as a vec- and 11.2.
tor. Note that this is the same substitution used with the The driving point impedance ZD.P., Eq. 11.24, can be ex-
frequency response relationship. Again, AC analysis is a pressed as a phasor operator when its parameter values are
specialized frequency response analysis. Phasor operators known, Fig. 11.13. We will use R = 3 Ω, L = 10 µ H, and
can represent the impedances of individual elements or the C = 2 µ F and the input frequency of w = 377 rad/sec.
644 11  AC Circuits and Motors

Table 11.1   Impedances of electrical elements the flux linkages are related as:
Resistance Capacitance Inductance Ns
V (s) V (s) 1 V (s) λs = λp
= R ≡ ZC = ≡ ZC = Ls ≡ Z L Np
I (s) I ( s ) Cs I (s)

Exciting a transformer with a time-varying alternating cur-


Table 11.2   Phasor operators of electrical elements rent produces a time-varying flux φ. Consequently, the flux
Resistance Capacitance Inductance linkage λ is time varying and produces voltage across the
coils.
V ( jw ) V ( jw ) 1 −j V ( jw ) The relationship between the input and output voltages in
= R ≡ ZC = = ≡ ZC = jLw ≡ Z L
I ( jw ) I ( jw ) jCw Cw I ( jw )
the transformer is

N 
R 1 V34 =  s  V12
Vsource ( s ) LCs + RCs + 1 s2 + Cs +
 Np 
2
Z D.P. = = → Z D.P. = L LC
I source ( s ) Cs 1
s
L The relationship between the currents through the primary

2
rad  3Ω rad 1 and secondary coils is calculated using energy conserva-
 j 377  + j 377 + tion, neglecting the energy lost as heat in the transformer.
sec 10 µH sec 2 µF (10 µH )
Z D.P. = In steady-state, using the sign convention that power flow is
1 rad
j 377 positive into the transformer,
10 µH sec
1
( 10 8
Vsource ( s ) 5 × 10 + j1.13 × 10 sec 2 ) Pp + Ps = 0 → V12 i p = −V34 is
Z D.P. = = N  N 
I source ( s ) 3.77 × 107
1
V12 i p = −  s  V12 is → i primary = −  s  is
Ω ·sec 2  Np   Np 
Vsource ( s )
Z D.P. = = 3 Ω − 1,330 jΩ.
I source ( s ) Most engineering students are familiar with the use of trans-
formers to step up and step down voltages in electric power
distribution systems. Transformers have other important ap-
11.5 Electrical Transformers plications. One is to provide a magnetic interface between
electrical subcircuits. A second is to allow “impedance
Electrical transformers consist of two or more coils of wire matching” in alternating current (AC) circuits.
wrapped around the same ferromagnetic core, Fig. 11.14.
The same magnetic flux φ flows through both coils. By con-
vention, the input coil is called the primary and the output 11.5.1 Model of a Transformer with a Resistive
coil is called the secondary. Although schematics show the and Capacitive Load
primary and secondary coils at different locations on the fer-
romagnetic core, in practice both coils are wound together. We will consider an RC circuit connected to the secondary
The black dots are the “dot convention” which mark trans- coil of a doorbell transformer, Fig. 11.15.
formers and schematics with the lead that are positive simul- How does one visualize the response of this circuit? First
taneously. Note that the schematic also shows the AC source analyze, or break the system into pieces, to create simpler
in the positive half cycle. models. The effect of the primary side of the transformer is
The number of turns in the primary and secondary wind- to produce a sinusoidal voltage across the coil of the second-
ings are Np and Ns, respectively. Because the same flux φ ary side. If we are not interested in a power variable on the
passes through both coils, the flux linkage of one coil is re- primary side, we can replace the transformer with an equiva-
lated to the flux linage of the second coil. Using the defini- lent AC voltage source.
tion of flux linkage: We have simplified our model but we still have the ques-
tion, how does one visualize the response of a circuit excited
λ = Nφ by a sinusoidal input voltage? We begin with two “quasi-
static” cases (“quasi” means resembling or sharing some at-
and writing the flux through the two coils as: tributes of the actual) by considering the input voltage at its
maximum and minimum values. We can interpret these two
λp λp cases as if they were driven by a DC source. The jargon for
φ= =
Np Ns the two extremes is to say the input has “swung” positive
11.6  Three-Phase Power 645

Fig. 11.14   AC circuit with a


transformer with a load imped- φ is
ip
ance on the secondary side 1 3
+ +
+ Z load
AC Source
-
- -
2 4

Primary Coil Secondary Coil


Np turns N s turns

Fig. 11.15   AC circuit with a


transformer with a resistive and
φ is
ip
capacitive load on the second- 1 3
ary side
+ +
+ R
AC Source
5
- C
- -
2 4

Primary Coil Secondary Coil


Np turns N s turns

3 R 5 grees to draw phasor diagrams, and also when we report


our results, since degrees are easier to visualize than are
iR radians.
i iC C Algebraic expressions of trigonometric functions can be
difficult to visualize. For example, consider the voltages dif-
ferences a( t) − b( t), b( t) − c( t), and c( t) − a( t).

4=g
( )
a (t ) − b (t ) = V0 sin (w t ) − sin w t − 120o 
Fig. 11.16   Transformer replaced with the equivalent AC voltage
source driving the resistive and capacitive load on the secondary side ( ) (
b (t ) − c (t ) = V0 sin w t − 120o − sin w t − 240o  )
( )
c (t ) − a (t ) = V0 sin (w t ) − sin w t − 240o 

or has “swung” negative. They are represented by signs ad-


jacent to the AC voltage source. The positive case follows Note that the phase shifts would need to be converted to ra-
the “dot” convention, if the electrical schematic or transform dians to evaluate these equations. Figure 11.18 shows a( t),
provides it. If we are only interested in the variables on the b( t), and a( t) − b( t). Note that the voltage a( t) − b( t) has a
secondary side, Figs. 11.16 and 11.17, it will not matter if phase shift relative to both a( t) and b( t). The phase shift is
we are in phase with 180o degrees or out of phase with the easy to see in a phasor diagram.
primary side. The algebraic sum or difference of two sinusoids with the
same frequency will be a sinusoid at that frequency. What
are the magnitudes and phase shifts relative to a( t), b( t), and
11.6 Three-Phase Power c( t)? Fig. 11.18 shows the voltages of phases a and b and the
voltage difference between phases a and b.
We conventionally mix angular units when working with The magnitude of a( t) − b( t), b( t) − c( t), and c( t) − a( t)
AC. We must use [w ] ≡ rad/sec and [φ ] ≡ rad when we may be a surprise. It is a common error to presume that the
evaluate transfer functions and complex impedances. amplitude is V0, but V0 is the magnitude of the voltage of a
However, the phase angle φ is usually expressed in de- phase relative to ground. Comparing the voltages a( t) and
646 11  AC Circuits and Motors

Fig. 11.17   AC circuit on the


a b R
secondary side with a trans-
3 R 5 3 5
former, a resistive and capaci-
tive load. a AC source “swung”
positive, b AC “source” swung + iR - iR
negative
i iC C i iC C
- +

4=g 4=g

400
a-b c-a b-c Im o
j0
|V0 | a(0) = V0e
200
Re
o
-120

0|
VDC 0

|V
-200 o
-j120
b(0) = V0 e
-400
0 0.01 0.02 0.03 0.04
t, sec Fig. 11.20   Complex exponential vectors which represent a( t) and b( t)
Fig. 11.18   Voltages of phases a and b and the difference in voltage at t = 0
between phases a and b

The first step is to create complex exponentials which repre-


350 sent a( t) and b( t), Fig. 11.20. The vectors must have the correct
a-b
a amplitude and phase angle ϕ. Since we are evaluating steady-
200
state relationships that must hold for any arbitrary time t, we
are free to choose the time t at which we will draw the vector
VDC 0 diagrams. The most convenient time is t = 0. Phase a( t) has a
phase angle of zero, so it plots on the positive real axis. Posi-
-200 tive angles are counter-clockwise. Phase b( t) has a negative
b phase angle since it lags, or peaks later in time than does a( t).
-350 We next subtract the vectors to determine the magni-
0 0.01 0.02 0.03 0.04
tude of their difference. First, we multiply vector b(0) by
t, sec
minus one which is equivalent to rotating the vector 180o,
Fig. 11.19   Voltages of phases a and b and the difference in voltage Fig. 11.21, and then create a vector sum a (0) + −b(0) to per-
between phases a and b form the subtraction, Fig. 11.22.
Now the hard part. We need to find the magnitude and phase
angle of the vector a (0) − b(0) . We know the lengths of two of
b( t) with their difference a( t) − b( t), we see that the ampli- the sides of the triangle formed by a(0), -b(0), and a (0) − b(0)
tude of the difference a( t) − b( t) is greater for the simple rea- and we know the angle between a(0) and -b(0). There is surely
son that they are on opposite sides of the v = 0 line for most a sophisticated formula which we could hunt for in a handbook
of a cycle. What is likely a surprise is that the difference of mathematics that would yield magnitude and phase angle of
a( t) − b( t) peaks before either a( t) or b( t), Fig. 11.19. a (0) − b(0) , but it is usually faster to use what we know then
search for what we may not be able to find. We can create a
right triangle with a (0) − b(0) as its hypotenuse by dropping a
11.6.1 Line-to-Line Voltage line perpendicular to the real axis, Fig. 11.23.
Calculate the magnitude of the vector a (0) − b(0) with
We wish to determine the magnitude of a( t) − b( t), that is, the the Pythagorean t­heorem.
( )
magnitude sin (w t ) − sin w t − 120o . It is possible to use trig-
onometric identities and formulae but a vector construction 
( V (1 + cos (60 ))) + ( V sin (60 ))
2 2

using complex exponentials and Euler’s sine formula is easier. a ( 0) − b ( 0) = 0


o
0
o
11.7 Physical Principles of Three-Phase AC Motors 647

o o

Im -b(0) = -V0 e
-j120
= V0 e
j60 Im
)
- b(0

0|
(0)

|V
|V0| sin(60o)

0|
a

|V
o
60
|V0 | a(0) = V0e j0 |V0 |

|V0 | Re |V0 | cos(60o) Re


o
-120
0|

Fig. 11.23   Complex exponential vectors which represent a( t), b( t),


|V

and a( t) − b( t) at t = 0

o
-j120
b(0) = V0 e
This is a useful relationship to remember. The “line-to-line”
voltage (in this case a (t ) − b(t ) in a three-phase machine is
Fig. 11.21   Complex exponential vectors which represent a( t), b( t), 3 times the voltage of a single phase relative to ground.
and − b( t) at t = 0
Returning to the time-domain plots, note that it is difficult
to visualize the phase angle of a (t ) − b(t ) relative to a(t) from
these plots. Also note that the phase angles of the voltages
Im
a (t ) − b(t ), b(t ) − c(t ), and c(t ) − a (t ) differ from each other
) and from a(t), b(t), and c(t). The phase angle of a (0) − b(0)
- b(0 relative to a(0) is easily derived from the vector construction
(0)
0|

a using trigonometry.
|V

o
-b(0)= V0e j60
|V0 |
j0 o Re φab = tan −1 
 ( )  = tan  sin (60 ) 
V0 sin 60o −1
o

a(0) = V0e
 V0 (1 + cos (60 ))
o
 (1 + cos ( 60 ))  o

Fig. 11.22   Complex exponential vectors which represent a( t), b( t),


and a( t) − b( t) at t = 0  0.866 
φab = tan −1  = 30o
 1.5 
Factoring V0 out of the radical.

 11.7 Physical Principles of Three-Phase


( V ) (1 + cos (60 )) + (sin (60 )) 
2 2
a ( 0) − b ( 0) =
2 o o
0
AC Motors

(1 + cos (60 )) + (sin (60 ))
2 2
a ( 0) − b ( 0) = V0 o o
11.7.1 Rotating Magnetic Vector

Expanding the first term in the radical A “magnetic vector” is a vector in the direction of a mag-
netic field. If the magnetic field is symmetric, like a field of
(1 + cos (60 )) ( ) ( )
2
o
= 1 + 2 cos 60o + cos 2 60o a solenoid, the magnetic vector lies on the axis of symmetry.
The magnitude or length of the vector conveys the strength
Rearranging of the field. In three-phase AC motors, a magnetic vector is
 created by the windings (coil) of each phase. The windings
( )
a ( 0) − b ( 0) = V0 1 + 2 cos 60o + cos 2 60o + sin 2 60o  ( ) ( ) of three phases are on the stator, which is the stationary por-
tion of a motor, and spaced one-third of the circumference
Using the trigonometric relationship from a unit circle or 120° degrees apart around the axis of rotation of the rotor.
cos 2 (α ) + sin 2 (α ) = 1 The intensity of the magnetic field of each phase varies and
reverses direction with the sinusoidal variation in current
 through the windings. The symmetry of the sinusoidal varia-
a ( 0) − b ( 0) = V0 1 + 2 cos 60o + 1 ( ) tion of current and the spacing of the three sets of windings
120° apart on the stator lead to the resultant of the vector sum
yields the resulting line-to-line voltage. of the magnetic vectors of the three phases to have a con-
 stant magnitude and to rotate about the axis of the rotor once
a ( 0) − b ( 0) = V0 3 (11.26) every electrical cycle, Figs. 11.24 and 11.25. The rotating

648 11  AC Circuits and Motors

Im t3
)
φstator
(0
) -b
a(0 |V0| sin(60o)

0|
b

|V
φab |V0|
o
60 φstator
t4 a c
|V0 | cos(60o) Re φstator
t2
Fig. 11.24   Complex exponential vectors which represent a( t), b( t), and
a( t) − b( t) at t = 0 showing the phase angle ϕ of a( t) − b( t) c a

b
resultant magnetic vector drives the rotation of the motor’s
rotor, Fig. 11.26.
φstator
t1
11.7.2 Three-Phase Synchronous Motors
Fig. 11.26   Resultants of the vector sums of the phases’ magnetic vec-
The vector sum of the three stator phases is a vector of con- tors and times t1, t2, t3, and t4. The resultant vector represents the stator’s
rotating magnetic field
stant magnitude which rotates about the axis of the stator at
the AC frequency. The magnetic field of the rotor is created
by a DC current either conducted to the rotor through “slip 11.7.3 Three-Phase Induction Motors
rings” or generated on the rotor by an “exciter.” Since the
rotors magnetic vector is created by a DC current, it is fixed Induction motors are the most common AC motor. The in-
in its position relative to the rotor, Fig. 11.27. The rotor’s duction motor was invented concurrently by Nikola Tesla
magnetic vector chases the stator’s rotating magnetic vector. and Galileo Ferraris (1847–1897). An induction motor’s
Synchronous motors must be in synchrony or they lose magnetic field is created by “eddy” currents induced in
torque. Consequently, synchronous motors are not self-start- the rotor’s conductive bars by the stator’s magnetic field,
ing. They must be started and brought up to synchronous Fig. 11.28. The requisite for induction of the eddy currents
speed by another motor, or by an auxiliary set of windings by the stator’s magnetic field is for that field to rotate faster
on the same rotor. than the rotor. The difference is rotational speeds of the sta-
tor’s magnetic field and the rotor is termed “slip.”

Fig. 11.25  a One cycle of


three-phase voltage, b Positive
a t1 t2 t3 t4 c
200
direction of magnetic vectors, b
c Vector sums of the phases’ a(t)
Resultant
magnetic vectors and time t1, t2, b(t) 0 a t2
t3, and t4 Resultant
c(t)
t1 b
- 200 c c
0 0.004 0.008 0.012 0.016
t, sec
b
Phase b
Phase a c

Resultant
c
b
t3
Resultant
Phase c b a
t4
Phase Windings Positive
Magnetic Vector Directions a
11.8  Three-Phase AC Circuits 649

Fig. 11.27   A schematic cross-


section of a three-phase synchro-
nous motor Three phase AC energizes the
three separate field windings, b
phases a, b, and c. The magnetic
vectors of the three phases sum to a c
create a magnetic vector which
rotates around the axis of
the stator at the AC frequency.

c a
DC current creates a magnetic
vector fixed in position relative
to the rotor. The rotor rotates to b φrotor
align its magnetic vector with the
stator’s magnetic vector.

Copper or 11.7.4 Variable Frequency Motors


aluminum bars Induced “eddy” current
Variable frequency motors are induction motors driven vari-
ous speeds by varying the AC frequency. The variable fre-
quency sinusoidal voltage is created by the motor’s “driver,”
which are the motor’s electronics and power supply. A vari-
able frequency motor driver approximates a sinusoidal volt-
age from rectangular pulses, a process known as “inverting,”
The bars of the squirrel cage are “shorted” since half of the sine wave is negative.
or electrically connected by the end caps The idea of a variable frequency motor dates to the nine-
which also provide mechanical support. teenth century but was not practical until electronics were
developed and not economical until the electronics became
Fig. 11.28   A schematic of the rotor of an induction motor, also known
as a squirrel cage motor because the rotor resembles the exercise wheel inexpensive, as it has in the recent decades.
in a pet rodent’s cage

400
11.8 Three-Phase AC Circuits
Design A
Design D 11.8.1 Wye (Y) and Delta (Δ) Three-Phase
Percent of Rated Torque

300 Connections

We will now more formally consider three-phase electri-


200 cal current using phasors. Three phase lines can be con-
nected in a “Y” (also spelled as “wye”) connection or a
Design C “delta” connection, Fig. 11.30. The names refer to the con-
100
figuration of the connections, shown below, where the lines
Design B are phasors which represent the voltage of each of the three
0 phases. The voltage drop in the wye connection is between
0 20 40 60 80 100 one phase and neutral, which is a “common” return line used
Percent of Rated Speed to complete each of the three circuits. Consequently, a wye
connection requires four lines. The voltage drop in the delta
Fig. 11.29   NEMA standards specify four different induction motors connection is between two of the phases. The vector calcu-
designs by their torque-speed curves
lation we performed above found the “line-to-line” voltage
between phases a and b. This is a delta connection. We de-
Induction motors are self-starting. They have no brushes, termined that the voltage drop between the two phases was
since there is no need to conduct current to the rotor, and larger than the voltage of either phase to neutral by a factor
have lower maintenance cost than motors with brushes. of 3. The delta connection needs only three lines and has
NEMA standards specify “designs” of induction motors, a larger voltage drop across a load than does the wye con-
categorized by the shape of the torque-speed curve, Fig. 11.29. nection for the same phase voltages. If the delta connection
650 11  AC Circuits and Motors

a b b There is an odd aspect of the wye connection generator


schematic. Notice that the AC sources are marked with a
positive and negative polarity. What do polarity markings
b mean if AC voltage is a sinusoid symmetric about zero? The
polarity of an AC source refers to the phase shift. Reversing
neutral the source polarity it is equivalent to inverting the sine wave,
a
which is equivalent to a phase shift of 180°.
The neutral line is needed to complete each of the three cir-
c a cuits. If you have done any home wiring, you know that neu-
tral and ground are the same terminals in a residential electric
c service. Grounding the neutral line assigns it a voltage of zero.
If we consider just one phase, we can visualize the “positive”
Fig. 11.30  a “Y” or wye connection. The center terminal is neutral terminal of the phase swinging through the sinusoid voltage,
which is at ground voltage. b Delta connection alternately pushing current into the neutral line and pulling
out of the neutral line. As it happens, when the three phases
Fig. 11.31   A three-phase wye
are connected to the same neutral line, if they have a “bal-
b anced” load such as a three-phase motor which pulls the same
connection generator
+ amount of power from each line, the neutral line will carry no
current. At any instant, the net current being “pushed” into
the neutral line equals the net current being “pulled” out of it.
- You will see schematics in which the neutral line is omitted.
n - +
a However, it must be present for safety. It will carry current in
-
the event the load becomes “unbalanced.”
AC Source It is conventional to analyze three-phase circuits by con-
+ sidering just one of the phases and then tripling the resulting
current and power calculated. For example, a three-phase
c generator driving a three-phase motor is shown in the sche-
matic in Fig. 11.32.
Any of the three phases can be considered in isolation,
needs fewer wires and provides a larger voltage drop, why but it is convenient to work with phase a since the voltage
use a wye connection to drive a three phase motor? Many supplied by phase a has a phase shift of zero. The schematic
three-phase motors in fact use both types of connections to becomes a familiar first-order RL circuit when it is redrawn
maximize performance by varying torque or power, or to for any one of the three phases, Fig. 11.33.
provide two speeds, where a wye connection is used for low An alternative way to draw the wye generator–wye motor
speed and a delta connection is used for high speeds. Mo- schematic is to rearrange it so that it resembles a radiation
tors operated this way are called “single winding-two speed” warning symbol, Fig. 11.34.
motors.

11.8.3 Delta Connected Machines


11.8.2 Wye Connected AC Machines
Delta connected machines do not have a neutral line,
A wye connection is easier to understand and handle analyti- Fig. 11.35. The voltage drop through the windings is from
cally. A schematic of a three-phase wye connected genera- one phase to another. The voltage drop is greater than the dif-
tor is shown Fig. 11.31. The three AC sources, a, b, and c, ference between the voltage of one phase and ground when
shown in the schematic are in fact three sets of “armature” one of the two phases is negative.
windings within one generator. The a, b, and c windings are It is difficult to visualize the equivalent circuit of a single
spaced around the stationary frame of the generator, called phase in delta connections since the three phases are refer-
the “stator,” such that when the generator’s rotor carrying enced to each other rather than to neutral as they are in the
the “field” winding (a DC electromagnet) passes over the wye connection. Fortunately, it is easy to convert a schematic
armature windings in succession, the voltages induced in the from a delta connection to a wye connection for analysis. In
three windings peak at successively later time. This yields fact, we developed the method when we calculated the “line-
three sinusoids which have phase shifts of 120°.
11.8  Three-Phase AC Circuits 651

Fig. 11.32   Three-phase genera- b b


tor and motor in wye connections
+

-
n - +a n a
-
AC Source

+
c c

Fig. 11.33   Phase b windings of b Phase b b


a generator and a motor in wye
connections +
Phase b windings Phase b windings
of AC generator of AC motor
-
n n
neutral

b a 11.8.4 Example 1 Three-Phase Delta Connected


+ Motor Phase Voltage
n
The phase voltage of a delta connected motor is 460 VAC,
- Fig. 11.36. Determine the equivalent wye connected motor
b - for analysis.
n - + a We are given the phase voltage, or line-to-neutral, is 460
n VAC. The voltage dropping across the windings of the motor
is the line-to-line voltage. Consequently, we must increase
+ the voltage applied to the equivalent wye connected motor,
as shown in Fig. 11.37.
c c
Fig. 11.34   A three-phase generator and motor in wye connection vline = v phase 3 = ( 3 ) 460 VAC = 798 VAC
The equivalent wye connected motor is as shown in Fig. 11.37.
to-line” voltage between a and b. Our result, rewritten with
subscripts ab for line a to line b and an for line a to neutral.
11.8.5 Example 2 Three-Phase Delta Connected
vab = van 3 = vbn 3 Motor Line Voltage

We must know if a reported “terminal” voltage is line-to- The line voltage of a delta connected motor is 460 VAC,
line,  vab, or line-to-neutral,  van, in order to create the equiva- Fig. 11.38. Determine the equivalent wye connected motor
lent single-phase circuit. In industry, the term “line voltage” for analysis.
means line-to-line. The term “phase voltage” means line-to- In this case, the applied voltage of 460 VAC is reported as
neutral. line-to-line. The voltage which drops across the windings of
a motor in a delta connection is the line-to-line voltage. We
vline = v phase 3 do not need to adjust the magnitude of the applied voltage
if we show the voltage applied to equivalent wye connected
motor as line-to-neutral, Fig. 11.39.
652 11  AC Circuits and Motors

Fig. 11.35   Three-phase genera-


tor and motor in delta connec-
b b
tions - +

+ -
c - a c a
+

Delta Connected Delta Connected


Three Phase AC Generator Three Phase AC Motor

460 VAC
460 VAC 460 VAC 460 VAC

460 VAC 460 VAC

c b a c b a

Fig. 11.38   A three-phase motor in a delta connection. “Line” volt-


Fig. 11.36   A three-phase motor in delta connection with 460 VAC age is voltage any between two of the three phases. If the reference
phase voltage voltage(s) are not indicated, the default assumption is that voltages are
“Line” voltages

798 VAC 798 VAC 798 VAC 460 VAC 460 VAC 460 VAC
c b a c b a

neutral neutral

Fig. 11.37   Equivalent wye connected motor with line-to-neutral Fig. 11.39   Equivalent wye connected motor with line-to-neutral
voltages equivalent to the delta connected motor of Fig. 11.36 voltages equivalent to the delta connected motor of Fig. 11.38

We do need to adjust the voltage applied to the equivalent 11.9 AC Power Calculation
wye connected motor if we show the voltage as line-to-line.
A line-to-line voltage drop is greater than a line-to-neutral The product of current and voltage phasors is called “appar-
voltage drop, so we must increase the applied line-to-line ent” or “total” and has the units of volts · amperes.
voltage to have the correct line-to-neutral voltage Fig. 11.40.
Apparent Power ≡ Total Power ≡ V ⋅ I
vline = v phase 3 = ( 3 ) 460 VAC = 798 VAC V ⋅ I = V e j φV I e j φI = V I e j φV e j φI

Apparent Power = V I e ( V I )
j φ +φ
11.9  AC Power Calculation 653

798 VAC 798 VAC 798 VAC sents the flow of energy into either the magnetic field of an
c b a inductance or the electric field of a capacitance. Reactive
power is real power, but it is not converted into mechanical
power in an electric motor. The energy stored in the mag-
netic and electric fields of an AC circuit or machine flows
in and then back out each cycle of the sinusoidal current.
Reactive power is genuine power. Reactive power is neces-
sary for a motor to run, since an electric motor’s torque
is the result of the magnetic fields of the stationary and
rotating components attempting to align. Reactive power
is “imaginary” in the sense that it cannot be taken out of
neutral the motor as mechanical power on the output shaft. Reac-
tive power has units of var (lower case), which stands for
Fig. 11.40   Equivalent wye connected motor with line-to-line voltages
“volts-amps-reactive.”
Electric motors list a “power factor” on the “patent
plate” (name plate) and in the motor’s specifications. The
jQ Reactive power factor is cos(θ ), where θ is the angle between the
Power, var total or apparent power and the positive real axis. Power
P P factors are often presented as 100 cos(θ ). The product of the
θ Active Power power factor and the apparent power vector is the projec-
watts tion of the apparent power vector onto the real axis, which
P is the useful active power of the motor plus the power dis-
+j
Q sipated as heat.
jQ
Apparent
Power, VA 11.9.1 Example AC Power Calculations Using
Motor Specifications
Fig. 11.41   Vector decomposition of apparent power. The two com-
ponents of Total or Apparent Power vector projected on the real and
imaginary axes are “active” and “reactive” power
11.9.1.1 Example 1
A “General Purpose” single-phase AC 10 hp motor specifi-
cations are shown ­below. Calculate the motor’s power.
Since the AC source voltage has no phase shift, φ V = 0
Single-phase AC 10 hp motor specifications
j φI Horsepower/Kilowatt 10/7.46
Apparent Power ≡ V ⋅ I = V I e
Voltage 230
[ Apparent Power ] = [ volts ⋅ amps] = [ VA ] ≠ [ watts ] Hertz 60
Phase 1
Full load amps 41
Isn’t a watt = i · v ? It is for DC power but not for AC power.
RPM 1750
DC power is the product of two real variables. Total power is
Service factor 1.15
the product of a real and a complex variable or two complex Rating 40C AMB-CONT
variables. When total power is expressed as a vector sum Locked rotor code H
the real component is the “active” power and the imaginary NEMA design code L
component is the “reactive” power. Insulation class H
The real component P is “active” power. Active power is Full load efficiency 84
power that is either dissipated as heat in the system or trans- Power factor 93
duced from electrical to mechanical power in a motor, or Ship weight 174 lbs.
transformed from one AC current and voltage to another in
an electrical transformer. Excluding the amount dissipated as We are given the voltage, the “Full Load Amps” (current), and
heat, active power is what we can use as mechanical power. the motor’s power factor.
Active power has units of watts (Fig. 11.41). Calculate the apparent power.
The imaginary component of the apparent power vec-
tor, jQ, is called “reactive” power. Reactive power repre- Apparent Power = (230 VAC)(41A) = 9, 430 VA
654 11  AC Circuits and Motors

jQ Reactive jQ Reactive
Power, var Power, var
P = 26.2 kW P
P = 8.7 kW P
21.5 o
Active Power 34 o
Active Power
P+j
Q=9 watts P+ watts
.4 k jQ
VA =3
jQ 1.6
Apparent kV
A
Power, VA jQ
Apparent
Fig. 11.42   Vector decomposition of apparent power Power, VA
Fig. 11.43   Apparent power phasor and its projection on to the real,
Calculate the active power. Multiplying the apparent power active power, and imaginary, reactive power, axes
by the power factor over 100 yields the projection of the ap-
parent power vector on to the real axis.
Power Factor 1. The “Full Load Amps” (current) given in a specification
Apparent Power × = Active Power of a three-phase motor is for one “leg” (or electrical sup-
100
ply line) of the three phases.
(9, 430 VA )(0.93) = 8, 770 W = Active Power 2. The “rated” voltage given in the specification (or on the
nameplate) of a three-phase motor is the “line-to-line”
Reduce by the “full load efficiency.” Note that the “Full voltage. Line-to-line voltage is measured between two of
Load Efficiency” is calculated after projecting the apparent the three phases, i.e., vab, vbc, vca. The voltage needed for
power vector onto the real axis. the power calculation is the “line-to-neutral” voltage. The
line-to-line voltage is a factor 3 greater than the line-to-
Active Power × Efficiency = Available Power
neutral voltage.
(8, 770 W )(0.84) = 7,370 W = Available Power 1
vrated = vline − to − neutral
The motor is rated as 7.46 kW, so our calculations are 3
accurate to the precision of the data. Repeat the above calculations using the data for the three-
To represent these results as phasors, we need the angle phase motor.
between the apparent power vector and the real axis, Calculate the apparent power for phase a, Fig. 11.43.
Fig. 11.42.
 1  
 power factor  Apparent Power =    230 VAC ( 238 A ) = 31.6 kVA
−1  93 
θ = cos −1
 = cos   = 21.5
o
 3 
 100 100 
Calculate the power for phase a. Multiply the apparent power
by the power factor over 100 to yield the projection of the
11.9.1.2 Example 2  active power vector on to the real axis.
“General Purpose” three-phase, Delta connected, 100  hp
Power Factor
motor specifications are shown below Apparent Power = Active Power
100
Three-phase, Delta connected, 100 hp motor specifications
Horsepower/kilowatt 100/74.6
(31.6 kVA )(0.83) = 26.2 kW = Active Power
Voltage 230/460
Reduce the active power of phase a by the “full load effi-
Hertz 60
ciency” to the shaft power.
Phase 3
Full load amps 238/119
RPM 1185 ( Active Power )( Efficiency ) = Shaft power of one phase
Service factor 1.15
Rating 40C AMB-CONT
( Active Power )( Efficiency ) = ( 26.2 kW )(0.941)
Locked rotor code G Shaft power of one phase = 24.6 kW
NEMA design code B
Insulation class B The average shaft power from the other two phases is the
Full load efficiency 94.1 same. The average shaft powers over a cycle add as scalars,
Power factor 83
Ship weight 1267 lbs
11.10  Single-Phase AC Motors 655

so the shaft power is three times the average shaft power Main Winding
from one phase. L1

(3)( 24.6 kW ) = 73.8 kW


+
Starting
The rated power is 74.6 kW. Our calculation is within - Winding
L2
approximately 1%.
You will see the power calculations for three-phase mo-
Squirrel Cage
tors performed using a formula which simplifies the calcula- Centrifugal Rotor
tion but removes some of the meaning. Switch
C

Rated Power kW = 3 × Rated VAC × Full Load Amp Fig. 11.44   Schematic of a capacitor start single-phase AC induction
× Power Factor × Efficiency motor

To limit errors, it is important to see that the factor 3 results


from correcting from line-to-line voltage to line-to-neutral 11.10.1 Capacitor Start
voltage and from multiplying by three to sum the average
shaft power from all three of the phases. A common design is a “capacitor start” single-phase AC
motor. A capacitor start motor has an auxiliary set of wind-
3
3= ings which is used to start the motor with the main windings,
3 Figs.  11.44 and 11.45. The starting windings are in series
with a large capacitor creating a phase shift between the start-
ing windings and the main windings. The starting windings
11.10 Single-Phase AC Motors are turned off by a centrifugal switch which opens when the
rotor has spun up to speed.
A single stator winding excited by AC power can produce a
magnetic field which pulses in amplitude sinusoidally but can-
not produce a rotating magnetic vector. The challenge in the 11.10.2 Series Wound
design of single-phase AC motors is to get them started. Once
there is rotation, then the interaction of the stator winding with A series wound single-phase AC motor is a high speed design.
the rotor can create enough phase shift between the stator’s The series connection between the stator and rotor windings
and rotor’s magnetic vectors to keep the motor running. requires brushes to conduct current to the rotor windings.
The motor behaves as if it were a DC motor because the
sinusoidal oscillation reverses the direction of the stator and
rotor fields simultaneously, leaving them in the same relative
orientation, Fig. 11.46.

Fig. 11.45   Cross-section of Main Winding


a single-phase AC motor with
auxiliary starting windings

Auxiliary Starting Winding

Rotor Winding
656 11  AC Circuits and Motors

Fig. 11.46   Cross-section of a


series wound single-phase AC
motor
“Salient” pole

“Field” winding
on the stator

The rotor and stator


φstator
(field) windings are
connected in series φrotor

Complex impedances, Z, simplify the analyses and can be


11.11 Mechanical Design Considerations used symbolically and graphically. The inverse of complex
impedance is complex admittance, Y. The dynamic attri-
11.11.1 NEMA butes of an AC circuit or machine are commonly described
by its driving point impedance.
NEMA is a US electrical equipment standards body. Phasors are vectors which represent the variables of AC
NEMA stands for National Electrical Manufacturers As- analyses, voltage and current. They are complex exponen-
sociation. NEMA standards exist for mechanical design tials with RMS magnitudes and evaluated at time t = 0, which
considerations, including motor performance, power sup- is set at the beginning of any cycle of the input voltage in
ply requirements, frame sizes, mounting bolt patterns, elec- steady-state. Complex impedance represented as a vector
trical enclosure designs, and safety. The induction motor is termed a phasor operator. Phasor operators operate on
classification based on the shape of its speed–torque curve phasors. The complex plane of a phasor operator (complex
is a NEMA standard. impedance) is the resistance–reactance plane, where resis-
tance is the real axis and reactance is the imaginary axis.
Reactance represents the energy temporarily stored in the
11.11.2 US Department of Energy Efficiency AC system during its cycle. The energy stored is returned to
Data the grid. The power is not used, but the electric utility must
provide the current.
Electric motors consume a large proportion of the electri- AC power’s complex plane’s real axis is “active” power.
cal energy generated. Small improvement in the overall ef- The imaginary plane is “reactive” power. Power drawn as
ficiency of the electric motors has both economic and en- shaft work from a motor is active power. The power factor of
vironmental benefits. The US Department of Energy has a motor is 100 cos(φ ), where φ is angle between the phasor
created a database of the specifications and efficiencies of operator and the positive real axis.
commercially available electric motors and offers free soft-
ware to help engineers identify the most efficient motors for
particular design requirements. Problems

11.1 A single-phase AC motor with the driving point imped-


Summary ance Z = 0.5 Ω + j 0.08 Ω is connected to a 120 VAC
voltage source. Determine the current drawn by the
AC power in North America is generated at 60 Hz or motor and the apparent, reactive, and active power of
377 rad/sec. Europe and most of the rest of the world use the motor.
50 Hz, which is less efficient. Polyphase power is two phase 11.2 Two single-phase AC motors with impedances Z1 and
for residences and three phase for industry. AC circuit analy- Z2 are connected in parallel to a 120 VAC voltage
sis is a specialized application of frequency response analy- source Fig. P11.2.
sis. AC analyses are steady-state analyses. The magnitudes Determine the total current drawn by the motors and
of voltage and current are root-mean-square (RMS) values. the apparent, reactive, and active power of the each
Conversion from the true, DC amplitude, V0, of sinusoidal motor.
voltage to the RMS value is 11.3 A three-phase, delta connected AC motor with the driv-
ing point impedance Z = 3 Ω + j1.2 Ω is connected to a
V0
VRMS = = 0.707 V0
2
Chapter 11 Appendix 657

Fig. P11.2  a Motor circuit, a b


b Motor impedances R2 L2
Z2 = 5 Ω + 4.2 jΩ
jΩ
R1 L1

+
Z 1= 3.5 Ω + 1.6 jΩ

-

Fig. P11.5   An RL Fig. P11.7   An RL AC R1


AC circuit
R1 L circuit

R2 L R2

Fig. P11.6   An RC AC Fig. P11.8   An RLC AC C


R1 C circuit L
circuit

R2 R1 R2

240 VAC voltage source. Determine the current drawn 11.7 The RL circuit shown in Fig. P11.7 is powered by
by the motor and the apparent, reactive, and active 120 VAC. Use complex impedances to determine the
power of the motor. driving point impedance for R1 = 3 Ω, R 2 = 20 Ω,
11.4 A three-phase, wye connected AC motor with the driv- and L = 10 mH. Draw the phasor operator of the
ing point impedance Z = 3 Ω + j1.2 Ω is connected to a driving point impedance. Determine the current
240 VAC voltage source. Determine the current drawn drawn by the circuit and the apparent, reactive, and
by the motor and the apparent, reactive, and active active power.
power of the motor. 11.8 The RLC circuit shown in Fig. P11.8 is powered by
11.5 The RL circuit shown in Fig. P11.5 is powered by 120 VAC. Use complex impedance to determine the
120 VAC. Determine the driving point impedance for driving point impedance for R1 = 3 Ω, R 2 = 20 Ω, and
R1 = 3 Ω, R 2 = 20 Ω, and L = 10 mH. Draw the phasor L = 10 mH. Draw the phasor operator. Determine the
operator of the driving point impedance. Determine the current drawn by the circuit and the apparent, reactive,
current drawn by the circuit and the apparent, reactive, and active power.
and active power
11.6 The RC circuit shown in Fig. P11.6 is powered by 120
VAC. Use complex impedances to determine the driv- Chapter 11 Appendix
ing point impedance of the RC circuit for R1 = 3 Ω,
R 2 = 30 Ω, and C = 40 mF. Draw the phasor operator Evaluation of the Root-Mean-Squared Integral
of the driving point impedance. Determine the current
drawn by the circuit and the apparent, reactive, and This integral is evaluated twice, first in its present form and
active power. then with the sine function expressed in terms of complex
exponentials using Euler’s formulas.
658 11  AC Circuits and Motors

T T We now evaluate integral expressed in terms of complex


∫ (V sin (w t )) dt = V02 ∫ sin 2 (w t ) dt
2
0 e­ xponentials. Recall Euler’s sine and cosine formulas.
0 0

e jθ − e − jθ e jθ + e − jθ
sin(θ ) = cos(θ ) =
Consulting a table of integrals for the integral of sine squared 2j 2
yields,
If we integrate over a full cycle, it makes no differ-
1 1
∫ sin (wt ) dt = 2 t − 2w cos (wt ) sin (wt ) ence if we use cosine or sine. Using a cosine function
2

v(t ) = V0 cos(w t + φ ), express it using complex ­exponentials.


1 1
∫ sin (wt ) dt = 2 t − 4w sin ( 2wt )
2

 e j (wt + φ ) + e − j (wt + φ ) 
v(t ) = V0 cos(w t + φ ) = V0  
 2 
Thus
T
T 1
(V0 sin(w t + φ ) ) dt
T
1 1  VRMS = ∫
2
V ∫ sin (w t ) dt = V  t −
2 2
cos (w t ) sin (w t )
2
T 0
0
0
 2 2w 0
 0

2
or 1
T
 e j (wt + φ ) + e − j (wt + φ ) 
T
VRMS = ∫ V0 
T 0  2  dt

T
1 1 
V02 ∫ sin 2 (w t ) dt = V02  t − sin ( 2 w t )
0
 2 4w  0 T
V02
∫ (e ) dt
2
j (w t + φ )
VRMS = + e − j (w t + φ )
Either form will do since the integral is evaluated over a full 4T 0

cycle of period T where T = and
w Expanding the integral

 2π 
cos (w t ) t = 0w = cos  w − cos ( 0) = 1 − 1 = 0
t=
T
w  ∫ (e ) dt
2
j (w t + φ )
 + e − j (wt + φ )
0

 2π  T
sin (w t ) t = 0w − sin ( 0) = 0 − 0 = 0
t=
= sin  w
 w  (
= ∫ e j 2(wt + φ ) + 2e j (wt + φ ) e − j (wt + φ ) + e − j 2(w t + φ ) dt
0
)

 2π  T T
sin ( 2w t ) t = 0w = sin  2 w − sin ( 0)
t=

∫ (e ) dt = ∫ (e )
2
j (w t + φ )
 w  + e − j (wt + φ ) j 2 (w t + φ )
+ 2e0 + e − j 2(wt + φ ) dt
0 0

sin ( 2w t ) t = 0w = sin ( 4π ) − sin ( 0) = 0 − 0 = 0
t= T T T T

∫ (e ) dt = ∫ e
2
j wt + φ )
(
+ e − j (wt + φ ) j 2 (w t + φ )
dt + 2∫ dt + ∫ e − j 2(wt + φ ) dt
0 0 0 0

due to the symmetry of sine and cosine about zero. Hence


du deu
Evaluating the first integral term using eu = and
T T dx dx
1 T
∫ (V 0 sin (w t )) dt = V0
22
t = V02 2π
0
2 0 2 recognizing that the angular frequency w =
T
Substituting into the expression for root-mean-square,
T  j 2 T T + φ 0 + φ 
T  2π   2π   2π 
j 2 t + φ j 2

∫e
 T   T 
dt =  e − e 
1
T
1 2 T 4π  
(V0 sin (w t )) dt =
2

0
VRMS = V0
T 0 T 2
T  2π 
T j ( 4π + 2φ )
( )
j 2 t + φ

∫e − e j 2φ
 T 
V0 V dt = e
VRMS = = 0 4π
2 1.414 0
Reference and Suggested Reading 659

The complex exponentials e j (4π + 2φ ) and e j 2φ are the same Hence,


complex number since 0, 2π, 4π, and any other integer mul-
T
V02
∫ (e ) dt
tiple of 2π are the same angle. Therefore VRMS = j (w t + φ )
+ e − j (wt + φ )
2

4T 0
T  2π 
j 2 t + φ T j (4π + 2φ ) T j 2φ
∫e − e j 2φ ) = (e − e j 2φ ) = 0
 T 
dt = (e
4π 4π V02 V2 V
0 VRMS = ( 2T ) = 0 = 0
4T 2 2
and
T  2π 
T − j ( 4π + 2φ )
( )
− j 2 t + φ

∫e − e − j 2φ
 T 
dt = e
0
4π Reference and Suggested Reading
T  2π 
T − j 2φ
( )
− j 2 t + φ
Chapman SJ (2005) Electric machinery fundamental, 4th edn. McGraw-
∫e − e − j 2φ = 0
 T 
dt = e
4π Hill, New York
0
Fitzgerald AE, Higginbotham DE, Grabel A (1981) Basic electrical
engineering. McGraw-Hill, New York
leaving the only term in the integral to be Fitzgerald AE, Kingsley C, Umans SD (2003) Electric machinery, 6th
edn. McGraw-Hill, New York
T T

∫ (e )
+ 2e0 + e − j 2(wt + φ ) dt = ∫ 2e0 dt = 2 (T − 0) = 2T
j 2 (w t + φ )

0 0
Index

A Bond graphs, 123
AC induction motors, 234 Boolean control, 3, 4
Across variable, 123, 124, 126, 127, 130, 132, 135, 136, 147, 174, Boolean variables, 499
181, 182 Branches, 19, 20, 24, 25, 35
difference, 125, 126 Branch point, 529
Active power, 653 Brushes, 351
Address, 495 Brushless DC motor, 649
Adjoint of a matrix, 421 Byte, 495
Alias, 68
Alternating current (AC), 635 C
Analysis, 26, 31 Cam and follower, 362
Angular frequency, 49, 69, 88 Cantilevered beam, 201
Angular velocity, 222 Capacitor start, 655
Antilog, 596 Cartesian coordinates, 65
Apparent power, 652 Cartesian form, 63, 66
Area-moment of inertia, 224 Cascaded blocks, 528
Attenuation, 599 Cause and effect, 148
Automatic control theory, 3, 4 Cavitation, 347
Automotive lead-acid storage battery, 47 Chain drive, 339
Characteristic equation, 54, 55, 57, 70, 71, 79, 545
B Characteristic function, 55, 57, 100, 159, 163, 164, 521, 524
Back-EMF, 352, 354 Characteristic value see Eigenvalue, 60
Base conversion, 496 Check valve, 483
Beat, 586 Circular pitch, 342
Belt drive, 339, 395 Closed-loop feedback control, 3, 4
multiport, 367, 384 Closed-loop negative feedback control, 531
Bending, 200 Closed-loop system, 533
Bevel gear, 344 Closed-loop transfer function, 532
Binary, 495 Cluster gear, 346
Binary place values, 496 Coast-down test, 244
Bingham fluid, 483 Coefficient matrix, 522
Bit, 495 Coenergy, 12, 13, 14, 22
Block diagram, 527 Cogging, 351
algebra, 529 Column matrix, 412
differential equation, 535 Commutator, 351
state equations, 542 Compatibility, 23, 28, 41, 125, 130
Bode plot Compatibility equations, 20, 21, 23–26, 29, 31, 36, 37, 123
critically damped phase-angle, 603 Complex exponential, 60, 64, 67, 73, 83, 158
first-order log-magnitude, 598 counter-rotating, 71
first-order phase-angle, 603 form, 65–68, 80
phase angle, 602, 620 imaginary component, 70
second-order log-magnitude, 598 real component, 70
underdamped phase-angle, 603 rotating, 70
Bode plot asymptotic approximation, 607 unit vector, 64
first-order factors, 608 Complex impedance, 639
gain, 614 Complex number, 56, 60, 61, 64, 65, 68
integrator and differentiator, 613 addition and subtraction, 61
sketching, 615 argument, 63
transfer function form, 607 cartesian form, 60
underdamped second-order factor, 610 complex conjugate, 60
K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1, © Springer Science+Business Media New York 2014 661
662 Index

magnitude, 61, 63, 65, 83 E


modulus, 61, 63, 83 Eddy currents, 351
multiplication and division, 62 Effective mass, 205
polar form, 63 cantilervered beam, 206
principal angle, 64 spring, 206
Complex variable, 54 Eigen equation, 57
Compliance, 199 Euler’s sine and cosine formula, 68, 71, 73
Compound gear, 346, 388 Eigenvalue, 57, 60, 69, 70, 75, 84, 155, 158, 179, 545
Computational model, 47 complex conjugate, 60, 71, 79, 83, 88, 101, 159, 163
Continuity, 21, 28, 32, 37, 125, 130 real, 84
Continuity equations, 24–26, 29, 31, 36, 43, 123, 126, 134, 174, 182 real component, 89
Corner frequency, 600 underdamped systems, 247
first-order, 598 Eigenvector, 444
second-order, 600 Elastic beam theory, 200
Controlled variable, 531 Elastic-perfectly plastic, 5, 9
Coulomb friction, 211 Electrical
torque on shaft, 243 capacitance, 30
Crank and slider, 362 schematics, 35
Creep, 235 transformers, 644
Critically damped, 88 Element
damping, 122
D energetic, 4, 20, 36, 118
D’Alembert’s force, 122 energy storage, 11, 30, 36, 127
Damped frequency, 585 equation, 36
Damped natural frequency, 549 equations, 126
Damping coefficient, 120, 122, 157 ideal, 27
Damping ratio, 103, 550 linear, 49
Dashpot, 217 mass, 118
DC generator, 354 mechanical, 4, 7
DC motor, 350, 353, 354 Elemental models, 32
DC servomotor, 242 Energetic
Decade, 597 attribute, 118, 135, 137, 174, 182
Decay, 549 models, 137
envelope, 88 network, 136
exponential, 143, 147 property, 1, 8, 9, 20, 36
rate, 70 schematic, 133, 154
Decaying sinusoid, 98 systems, 39
Decibel, 596 Energy, 1, 8, 13–15, 39, 157
Decision block, 500 density, 14, 24
Decomposition, 252 dissipation, 118, 157, 168
Delta connection, 649 elastic strain, 10, 11, 32, 129
Design practice, 27 equations, 37, 127
Determinant, 418 gravitational potential, 15, 27
Diagonal matrix, 420, 444 kinetic, 15, 27, 121, 141
Differential elemental equation, 140 mechanical, 39
Differential equations, 45, 46, 57, 59, 75, 127 potential, 11, 16, 20
first order, 142 storage, 30, 45, 46, 56, 135, 139
Differential system equation, 3, 16, 23, 37, 38, 42, 53, 93, 120, 121, storage equations, 127
125, 127, 132, 135, 137, 139, 149, 182 strain, 141
Diode, 483 Engineering
Direct pass-through matrix, 427 analysis, 5, 27, 39
Disassemble, 26 computations, 5, 8
Discrete time, 469 design, 26, 39
Disturbed system, 54 mathematics, 3, 4
Drag cup, 232 models, 5, 39
Driving point impedance, 642 Entangled loops, 534
Driving potential, 20, 21, 23, 30, 35, 36 Equations
Dynamical systems, 1, 2 energetic, 130
electric circuits, 3, 4 Equilibrium
Dynamic calculation, 70 dynamic, 123
Dynamic equilibrium, 29 Equivalent element, 255, 374
Dynamic range, 86, 144 dampers in parallel, 256
Dynamic system, 2, 3, 4, 8, 31, 45, 46, 52, 135, 142 dampers in series, 256
mathematics of, 3, 5 masses in parallel, 258
Dynamic tests, 235 springs in parallel, 257
springs in series, 257
Index 663

Error, 531 H
Euler method, 468, 469, 481 Harmonics, 586, 593
second-order system, 471 Heaviside step
Euler’s equations, 64 function, 142
Euler’s cosine formula, 64 input transitions, 182
Euler’s sine formula, 65, 99 Heaviside unit step function, 46, 47, 49, 74, 76, 78, 91, 92, 136–138,
Euler’s equations EulersEqs, 65 150
Euler’s number, 143 superposition of, 148
Euler’s sine formula, 158 time shifted, 148
Executable block, 500 Helical gear, 343
Expansion by minors, 419 Hexadecimal, 495, 496
Exponential, 54 Homogeneous
decay, 55, 89 equation, 51, 54, 55, 59, 71, 75, 77, 79, 155
properties of, 66 solution, 55, 60, 75, 77, 79, 81, 82, 142, 144, 155, 157
real, 84 Homogenous
Exponentiation, 596 response, 51, 56
solution, 58
F Hooke’s law, 29
Feedback, 531 Hybrid system
Feedforward loop, 530 electrical, rotational, and translational, 391
Feedforward transfer function, 531 rotational translational, 377
Feed-through matrix, 522 Hydraulic
Final value theorem, 93 actuator, 356
Finite difference, 467 motor, 356
First order ordinary differential equation, 74 piston/cylinder, 356
First order step response, 87 Hydrodynamic bearings, 233
First-order step response, 149
Floating, 368 I
mechanical sources, 368 Ideal viscous friction, 208
transducer or transformer, 373 Identity matrix, 97, 102, 523
Floating point variable, 499 Idler, 384
Flow chart, 499 Idler gear, 346
Fluid coupling, 233 Imaginary number, 60, 62, 84
Flux, 644 Impulse function, 76
Flux linkage, 636, 644 Impulse-momentum theorem, 582
Force Impulse response
source of, 133 overdamped, 247
Force-source-mass-damper system, 145 underdamped, 249
Force-source-spring-damper system, 140, 151, 162 Incrementally linear, 210, 521
linear graph of, 130, 132 Independent energy storage element, 56
Forcing function, 57, 58, 79, 80 Inductance, 636
Forward stepping, 468 Induction motors, 648
Fourier series, 46 Initial conditions, 59, 76, 78, 80, 138, 140, 142, 155
Fourier transformation, 49 energized systems, 161
Fourth-order Runge–Kutta Initial value, 146
two state equations, 479 method, 165
Frequency response theorem, 93
equation, 590 Input
relationship, 638 function, 46, 52, 152, 153
Fundamental frequency, 586, 593 impulse, 138
pulse, 148
G pulse function, 148
Gain, 599 unit pulse, 148
Gaussian elimination, 96 Input matrix, 522
Gear ratio, 346 Input pulse function, 151
Gear set, 334, 341, 384 Integral causality, 594
multiport, 365 Integrator, 536
Gear set, 334, 341, 348 Invariant, 24, 42
multiport, 365 Inverse mathematical operations, 144
Gear train see Gear set, 344 Inverse of a matrix, 523
General solution, 59, 76, 78, 82, 155, 157 Inverse tangent, 83
Geometric compatibility, 6, 32 algorithm, 63
Gibbs phenomenon, 594 function, 65
Graphical integration, 17, 31 Inverted, 529
Graphical representation, 27 Inverter, 529
Ground voltage, 36
664 Index

J Mass-moment of inertia, 222
Jenkins-Traub algorithm, 57 cylinder, 223
parallel-axis theorem, 224
K superposition, 224
Kinetic energy, 144, 198 Mathcad, 106
Kirkchoff’s circuit laws, 123 assignment statement, 106
Heaviside unit step function, 190
L
plot, 106, 190
Lag, 584, 587
superposition, 190
Lagrange’s and Hamilton’s energy methods, 4, 7
Mathcad’s Heaviside step function, 138
Laplace-domain, 90, 519, 521, 522, 526
Mathematically independent, algebraic equations, 21, 38
response function, 95
MATLAB
Laplace transform
absolute value function abs, 633
pair, 97, 103, 163
angle function, 633
tables, 95
arithmetic operators, 505
Laplace transformation, 3, 31, 38, 56, 71, 75, 90, 92, 93, 94, 99, 100,
array, 504
155, 157, 163, 178, 521
array and matrix operators, 507
initial condition terms, 56
array variables, 110
inverse Laplace transformation, 92, 94, 95, 98, 100
assignment statement, 110, 508
inverse transformation, 164
Bode plots, 632
transform integral, 90
colon operator, 513
transform pairs, 95, 101
comments, 510
Laplace variable, 54, 55, 90, 94, 95, 96, 521
control tool box, 515
Lead, 587
editor, 110
Leading, 584
elseif statement, 509
Lenz’s law, 352
else statement, 509
Lever, 335, 338
environment, 109, 503
multiport, 363, 379
flow chart, 112, 193
Linear algebra, 96, 98, 101
for-end loop, 110
Linear differential system equation, 521
for loop, 509
Linear elemental equations, 183
formatting plots, 511
Linear equations, 149
function imag, 634
Linear graph, 121, 124, 135, 154, 181
function real, 634
drawing, 218
if statement, 190, 508
nodes, 217
logical operators, 507
Linear graph method, 16, 30, 122, 129
nested loops, 191
Linear graph symbol
Nyquist plots, 634
transformer, 335, 337
plot, 111
Linearization, 208
plot statement, 510
Linearize, 53
plotting and figures, 511
Linear mathematical operators, 50
reading and writing to a text file, 513
Linear operator, 53, 519, 527
reading and writing to Excel worksheet, 513
Linear systems, 3, 5
reading and writing to files, 512
Line to line, 651
relational operators, 191, 507
Logarithmic
script, 111, 112
scale, 597
semilogx, 632
spiral, 70
square root function sqrt., 633
Logarithms
time shift, 191
common, 595
user written functions, 475
properties, 595
while loop, 509
Log-decrement formula, 249
workspace, 111
Logical variables, 499
MATLAB code, 470, 472, 474–476, 478, 479, 481, 483
Loop, 22, 39
Matrix
equation, 23, 41, 126
addition, 412
M cofactors, 418, 420
Magnetic flux, 352 differentiation, 414
Magnetic hysteresis, 351 inversion, 416
Magnetic reluctance, 352 multiplication, 413
Magnetic vector, 647 subtraction, 412
Magnitude, 588 Mechanical power, 211
Magnitude ratio, 581 Mechanical system
Mapping, 57 rotational, 20
Mass, 9, 14, 25, 197 translational, 20
conservation, 15, 26 Mechanical transformers, 334
flow rate, 20, 37 Method of undetermined coefficients, 46, 51, 74
Mass-damper system, 139, 161 Miter gears, 344
Index 665

Model, 27 condition, 119
abstract, 5, 118 point, 209
elastic-perfectly plastic, 9, 50 range, 10, 16, 47, 209
element, 29 Operator, 527
energetic, 8, 13, 118 Ordinary differential equation, 53, 77
engineering, 2, 4, 7, 26, 27, 31, 32 second order, 79
graphical, 28 Orienting, 25, 44
input, 137 Orthogonality, 419
linear, 50, 53, 120 Oscillations, 1, 16, 28, 56, 71
linear element, 50 decaying of, 69
linear material, 9, 15 period, 88
lumped parameter, 28 Oscillatory
material, 6, 10 second order step response, 88
mathematical, 6, 7, 28, 84 second order system, 99
network, 20, 36 Output matrix, 522
Newtonian, 8, 13 Overdamped, 86, 154
nonlinear, 7, 11 Overshoot, 70
nonlinear stress-strain, 7, 11
physical, 31 P
Modulus, 590 Parallelogram rule, 61
Motor action, 350 Partial fraction expansion, 92, 95, 560
Multiplicative operator, 93, 522 Particular solution, 51, 57, 78, 80, 142
Passive elements, 124, 125
N Path, 19, 35, 125
Natural fractions, 498 equation, 23, 41, 126, 136
Natural frequency, 249 equations, 126
undamped, 103 Period, 581, 583, 593
Natural response, 2, 3 Phase angle, 69, 82, 84, 635
NEMA, 649 Phase shift, 69, 81, 103
Nested loops, 533 Phase voltage, 651
Network, 19, 35 Phasor operator, 643
Network topology, 24, 42 Phasors, 642
Newtonian fluid, 119 Physical systems, 58
Newtonian formulation, 121, 122 Piece-wise continuous function, 89
Newtonian mechanics, 14, 26 Pitch circle, 342
Newton’s second law, 15, 26 Plant, 532
Node, 10, 19, 20, 21, 25, 35, 38, 121, 124, 125, 132, 147 Polar form, 588
equation, 125, 136 Pole, 547
ground, 132 Pole-zero form, 588
subscript, 125 Pole-zero plot, 551
subscripts, 22, 39, 122 Polyphase, 637
velocity, 122 Positive feedback, 532
Non-causal, 151 Post-multiplication, 97
Non-inverted, 529 Potential, 20, 36
Non-linear damping, 209, 473 Power, 8, 12, 15, 16, 20, 130, 147
Nonlinear differential equation, 53 flow, 15, 26, 124
Non-linearity ideal source, 47
backlash, 482 mechanical, 12, 17, 21, 122
clipping, 482 source, 16, 25, 29, 46
directional dependency, 483 variable, 46
inflection point, 482 variables, 1, 15, 16, 26, 143
saturation, 482 Power law viscosity, 474
threshold, 483 Power screw, 361
Nonlinear physical relationship, 51 Power sources, 212
Non-oscillatory see Overdamped, 154 rotational systems, 234
Normalization, 144 Power variables, 127
Numerical differentiation, 538 Practical pulse-width-modulated amplifiers, 148
Numerical instability, 90 Pre-multiplication, 97
Numerical methods, 47 Pressure, 20, 21, 36
Nyquist polar plot, 623 Proportional controller, 533
Proportional damping, 208
O Pulse response, 148
Observable canonical state-space form, 539 function, 153
Octal, 459 Pulse train, 148
Open-ended design problems, 4, 6 Pulse-width-modulation, 148
Open-loop, 532 Pump, 354, 396
Operating
666 Index

axial flow, 356 Signal, 527


gear, 355 Sign convention, 16, 28
positive displacement, 354 fluid transducers, 355
reciprocating piston, 355 linear graph, 215
turbopump, 356 mechanical systems, 214
Pythagorean theorem, 61, 62, 63, 81, 83 motors, 352
transducers and transformers, 346
Q Significant, 28
Quadratic equation, 77 Signs, 145
Simply supported beam, 201
R
Singular matrix, 420
Rack and pinion, 360
Sink, 214
Radix, 496
Sinusoid, 58
Reactance, 643
Sinusoidal inputs, 49
Reactive power, 653
Sinusoidal output, 588
Real numbers, 60
Sinusoidal response
Recursive calculation, 468
transient, 579
Reference signal, 531
Slip, 648
Relaxation, 235
Slug, 198
Relaxation time, 53
Snubbers, 209
Relevant physical truths, 28
Solenoid, 352
Resistive energy loss, 636
Source
Resonance, 584
force, 122
Resonant, 586
power, 122
Resonant frequency, 585
Spectrum, 598
Resonant peak, 585
Spring
effect of damping ratio, 601
translational, 129
Response
Spring back, 14, 25
steady-state, 143
Spring constant, 13, 23, 129, 199
Response function, 89, 146, 172, 173
Spring rate see Spring constant, 13, 23
pulse, 153
Spur gear, 342
Reversible systems, 16, 28
Square wave, 594
Rigid, 12, 20
Stability, 547
Rolling-contact bearings, 232
Stable, 84
Root-mean-square (RMS), 637
exponential growth, 76, 86
Rotating magnetic field, 648
Standard form, 53, 74, 77
Rotational damper, 233
State equations, 522
Rotational hydraulic motor
standard vector-matrix form, 522
gear, 359
vector-matrix form, 426
gerotor, 358
State-space, 423, 437
piston, 359
higher-order system equation, 438, 440
rotor, 358
State-space representation, 149
vane, 358
State variables, 139, 146, 147, 153
Rotational mechanical systems, 221
energy storage variables, 544
Row matrix, 412
without physical meaning, 544
Runge-Kutta algorithms, 89
State vector, 522
Runge-Kutta method, 45
Steady-state, 54, 58, 85, 147, 152, 161
three state equations, 480
response, 59, 89, 142
S value, 146
Saint-Venant’s principle, 9, 15 Step function
Schematic, 124 Heaviside’s definition of, 138
Second order Step response, 85
spring-mass-damper system, 156 function, 49
underdamped unit step response, 165 unit, 151
Second order step responses, 86 Strain energy, 199
non-oscillatory, 86 Summation junction, 529
overdamped, 251 Superposition, 46, 49, 51, 52
Second-order step responses Synchronous motor, 648
superposition of, 154 Synthesis, 26
Self-locking, 362 System, 19, 35
Series wound, 655 boundaries, 45
Serpentine belt, 384 energetic, 20, 36
Shear modulus, 230 fluid, 21, 38
Shear strain, 231 identification, 84
rate, 119 mass-damper, 128
Shock absorber, 210 mechanical, 12, 20
Index 667

nonlinear, 151 Underdamped system, 79


physical, 30 Unit exponential decay, 85
spring-damper, 130 Unit impulse function, 46, 48
System dynamics, 4, 6 Unit ramp function, 46, 48
System equation, 121, 134 Units, 53, 128
Unit stable exponential growth, 85
T Unit step response, 152
Thermal systems, 30 first order, 153
Thermodynamics, 16, 28 function, 46, 153, 178
Three phase, 637 Unstable, 548
Time constant, 51, 53, 84, 85, 86, 89, 121, 142, 144, 146, 151, 152, exponential growth, 69
153 system, 69
form, 121 U.S. customary units, 198
Time-domain block diagram, 536 mass moment of inertia, 226
Time-domain expression, 98
Time shift, 18, 32, 147, 150, 153 V
Time step, 89 Variable frequency motor, 649
Torque, 222 Vector algebra, 411
Torque constant, 352 Vector-matrix algebra, 411
Torque curve, 234 Vector of inputs, 427
Torsion spring, 230 Vector of output variables, 427
Transducer, 336, 350 Vector of transfer functions, 526
Transfer function, 67, 93, 94, 99, 162, 521, 524, 527 Velocity
Transformation matrix, 443 source, 129, 132, 212
Transformed state vector, 523 Velocity source-damper-mass-spring system, 174
Transformer, 333 Viscoelasticity, 220
block model, 336 Viscoelastic model
Transformers and transducers Kelvin–Voight, 220
multiport, 362 Maxwell, 220
Transient period, 51, 54, 161 Zener, 220
Transient response Viscosity, 119
first-order system, sinusoidal input, 582
Translational mechanical system, 129 W
Translational spring, 199 Weight, 197
Translational velocity, 129 Word, 495
Transpose of a matrix, 415 Work, 8, 11, 12, 13
Trapezoidal approximation, 148 Worm drive, 344
Trapezoidal integration, 484 Wye connection, 649
Trigonometric identities, 71
Y
Trivial equations, 33
Young’s modulus, 9, 15
U
Z
Underdamped, 154
Zero, 551
Underdamped pulse response
second order, 165

Das könnte Ihnen auch gefallen